Analog Design Centering and Sizing 2007
Analog Design Centering and Sizing 2007
Graeb
Analog Design
Centering
and Sizing
Analog Design Centering and Sizing
ANALOG DESIGN CENTERING
AND SIZING
HELMUT E. GRAEB
Institute for Electronic Design Automation
Technische Universitaet Muenchen, Germany
A C.I.P. Catalogue record for this book is available from the Library of Congress.
Published by Springer,
P.O. Box 17, 3300 AA Dordrecht, The Netherlands.
www.springer.com
List of Figures xi
List of Tables xix
Preface xxi
1. INTRODUCTION 1
1.1 Integrated Circuits 1
1.2 Analog Circuits 4
1.3 Analog Design 6
1.4 Analog Sizing 13
2. TOLERANCE DESIGN: EXAMPLE 19
2.1 RC Circuit 19
2.2 Performance Evaluation 20
2.3 Performance-Specification Features 20
2.4 Nominal Design/Performance Optimization 20
2.5 Yield Optimization/Design Centering 22
3. PARAMETERS & TOLERANCES,
PERFORMANCE & SPECIFICATION 27
3.1 Parameters 27
3.2 Parameter Tolerances 28
3.3 Range-Parameter Tolerances 29
3.4 Statistical Parameter Distribution 31
3.5 Univariate Normal Distribution 32
3.6 Multivariate Normal Distribution 34
3.7 Transformation of Statistical Distributions 39
3.8 Generation of Normally Distributed Sample Elements 43
vi ANALOG DESIGN CENTERING AND SIZING
INTRODUCTION
Figure 1. Trends in process and design technology. Data have been taken from the Intel web
site and from the International Technology Roadmap for Semiconductors (ITRS) web site.
at nearly constant percentage of ICs that pass the production test (yield) and
with a reasonable life time (reliability).
Design technology contributes to achieving the exponential growth of IC com-
plexity by developing the ability to design ICs with ever increasing number of
functions by
bottom-up library development methods for complex components and
top-down system design methodologies
for quickly emerging system classes.
As a result it has been possible to produce ICs with an exponentially grow-
ing number of transistors and functions while keeping the increase in the cost
of design, manufacturing and test negligible or moderate. Figure 1 illustrates
the technology and design development. The dashed line shows the increase in
number of transistors on Intel processors, which exhibits a doubling of this num-
ber every 24 months for 35 years. Similar developments have been observed for
other circuit classes like memory ICs or signal processors. The solid line shows
the development of design cost for system-on-chip ICs. If no progress in design
technology and electronic design automation (EDA) had been achieved since
Introduction 3
Figure 3. A CMOS operational transconductance amplifier (OTA, left) and a CMOS folded-
cascode operational amplifier (right).
circuit design for decades, therefore become more and more important for dig-
ital design as well.
Figure 4. An OTA-C biquad filter with OTAs from Figure 3 as building blocks [57].
Figure 5. A phase-locked loop (PLL) with digital components, phase frequency detector (PFD),
divider (DIV), and analog components, charge pump (CP), loop filter (LF), voltage-controlled
oscillator (VCO).
in Figure 5 flat on circuit level reaches hours or days. The usual design process
that requires frequent numerical simulations becomes infeasible in this case.
A remedy offers the transition from modeling on circuit level to modeling on
architecture level using behavioral models. By modeling the whole PLL on
architecture level or by combining behavioral models of some blocks with
transistor models of other blocks, the PLL simulation time reduces to minutes
and the usual design process remains feasible.
Typical analog design levels are system level, architecture level, circuit level,
device level and process level. An example of an analog system is a receiver
frontend, whose architecture consists of different types of blocks like mixers,
low-noise amplifiers (LNAs) and ADCs. On circuit level, design is typically
concerned with a transistor netlist and applies compact models of components
like transistors. These compact transistor models are determined on device
level requiring dopant profiles, which in turn are determined on process level,
where the integrated manufacturing process is simulated.
The design partitioning concerns the hierarchical decomposition of the sys-
tem into subblocks. Design partitioning is required to handle the design com-
plexity and refers to important design decisions with respect to the partitioning
between hardware and software, between analog and digital parts, and between
different functions and blocks. The design partitioning yields graphs of design
tasks that represent design aspects like block interdependencies or like critical
paths for the design project management.
During the design process, a refinement of the system and an enrichment
with more and more details concerning the realization occurs. Along with it,
the system consideration traverses the design levels from the system level over
the architecture level down to the circuit level. Moreover, the design process
switches among three different design views, which are the behavioral view, the
structural view and the geometric view (Table 2). The design level rather refers
to the position of the considered system component within the hierarchical
system (de)composition. The design view on the other hand rather refers to
the modeling method used to describe the considered system component. The
behavioral view refers to a modeling with differential equations, hardware and
behavior description languages, transfer functions, or signal flow graphs. It is
oriented towards a modeling of system components with less realization details.
The structural view refers to a modeling with transistor netlists or architecture
schematics. It is oriented towards a modeling of system components with
more realization details. The geometric view refers to the layout modeling.
On the lowest design level, it represents the ultimate realization details for IC
production, the so-called mask plans. During the design process, not only a
8 ANALOG DESIGN CENTERING AND SIZING
Table 2. Some modeling approaches for analog circuits in behavioral and structural view.
dv
dt + 1
RC v + IS v
C (exp( VT ) − 1) − 1
C i0 =0
Figure 7. Nonlinear analog circuit. Structural model in form of a circuit netlist, behavioral
model in form of a differential equation. Concerning the output voltage, both models are
equivalent.
refinement towards lower design levels takes place, but also a trend from the
behavioral view towards the structural and geometric view.
In analog design, often the structural view is to the fore, as transistor netlists
on circuit level or block netlists are prevalent in the design. But the behavioral
view is immediately included, for instance by the implicit algebraic differential
equations behind the transistor netlist or by program code. Behavioral and
structural models can be equivalent as in the two examples in Figures 7 and 8
but they do not have to be equivalent.
The behavioral model and the structural model in Figure 8 represent more
abstract models of the OTA circuit in Figure 3. While Figure 3 represents the
OTA on circuit level using detailed transistor models, Figure 8 represents the
OTA on architecture level to be used as a component of the filter in Figure 4.
The architecture-level structural and behavioral models are less complex, less
accurate and faster to simulate than the circuit-level models.
The relationship between design levels and design view in the design process
has been illustrated by the so-called Y-chart, in which each branch corresponds
to one of the three design views and in which the design levels correspond to
Introduction 9
Figure 8. Simplified models of the OTA in Figure 3. Behavioral model in form of a description
language code, Structural model in form of a circuit netlist. Concerning the OTA ports, both
models are equivalent and approximate to the transistor netlist in Figure 3.
concentric circles around the center of the Y [47]. From the design automation
point of view, the change in design view during the design process from the
behavioral view towards the structural and geometric view is of special interest.
The term synthesis as the generic term for an automatic design process has
therefore been assigned to the transition from the behavioral view towards the
structural view, and from the structural view to the geometric view, rather than to
the transitions in design level or design partitioning [84]. Figure 9 illustrates this
concept, where synthesis and optimization denote the transition from behavioral
to structural and geometric view, and where analysis and simulation denote the
inverse transition.
10 ANALOG DESIGN CENTERING AND SIZING
Figure 10. Design flow of a PLL. As in Figure 9, boxes denote the structural view, rounded
boxes the behavioral view.
Next, the design level switches to the circuit level, where the same process
from the architecture level is repeated for all blocks simultaneously, as indicated
by three parallel arrows in Figure 10. A transistor netlist capable of satisfying
the block specification is determined by structural synthesis. The circuit-level
parameter values, as the transistor widths and lengths, are computed by para-
metric synthesis.
The transition from the architecture level to the circuit level includes a tran-
sition from the behavioral view of the three analog blocks “under design” to the
structural view. Here a transformation of the architectural parameter values to
a circuit-level performance specification of each individual block is required.
Adequate behavioral block models do not only consider the input and output
behavior but prepare the required transformation by including model parameters
that have a physical meaning as performances of a block on circuit level, as for
instance the gain of the VCO.
12 ANALOG DESIGN CENTERING AND SIZING
After the synthesis process, the obtained real block performance values on
circuit level can be used to verify the PLL performance.
If we illustrate the synthesis process just described in a plane with design
level and design view as coordinates, we obtain a synthesis path as shown in
Figure 11.
This figure reflects that both structural and parametric synthesis generally
rely on function evaluations, which happen in a behavioral view. This leads to
periodical switches to the behavioral view within the synthesis flow that can be
illustrated with a “fir”-like path.
A top-down design process as just described requires on each level adequate
models from the lower design levels. These models are computed beforehand
or during the design process. Ideally, the results of the top-down process can
be verified. But as the underlying models may not be accurate enough or
incomplete, for instance for reasons of computational cost, verification may
result in repeated design loops.
The design of the PLL was illustrated starting on the architecture level. On
system level, signal flow graphs, which represent a yet non-electric information
flow, are often applied. Switching to the architecture level then involves the
important and difficult transition to electrical signals.
Introduction 13
The term design centering for yield optimization emanates from such a spa-
tial interpretation of the optimization problem. Section 7.1.2 will show that a
maximum yield is obtained when the center of gravity of the original probability
density function hits the center of gravity of the truncated probability density
function. Yield optimization as finding an equilibrium of the center of gravity of
the manufactured “mass” concerning the performance specifications becomes
design centering in this interpretation. This interpretation suggests that the
terms yield optimization and design centering should be used as synonyms.
As indicated in Figure 13, we emphasize the distinction between the analysis
of an objective and the optimization of this objective. We will see that yield
optimization approaches result immediately from the two ways of either statis-
tically estimating or geometrically approximating the yield. For that reason, the
terms “statistical” and “geometric” will be used to distinguish yield optimiza-
tion/design centering approaches. We will derive the first- and second-order
gradient of the yield based on the results of a Monte-Carlo analysis in Section
7.1. This leads to the formulation of a deterministic optimization method for the
statistically analyzed yield, which could be denoted as deterministic statistical-
yield optimization approach. We will derive the gradients of the worst-case
distances, which are geometric measures of yield, in Section 7.2.1, which leads
to a deterministic geometric-yield optimization approach.
Another important issue about yield optimization/design centering is the
difficulty to maximize yield values beyond 99.9%, which are described by
terms like six-sigma design. We will explain how to deal with these cases in
statistical-yield optimization in Section 4.8.7. Geometric-yield optimization
inherently covers these cases, as illustrated in Section 6.3.9.
Analog sizing is a systematic, iterative process that applies mathematical
methods of numerical optimization. The solution methods can be characterized
in different ways, which are summarized in Table 3.
16 ANALOG DESIGN CENTERING AND SIZING
1 2 3 4 5
Performance Specification Initial After nominal design After yield optimiza-
feature feature tion/design centering
Gain ≥ 80dB 67dB 100dB 100dB
Transit frequency ≥ 10M Hz 5M Hz 20M Hz 18M Hz
Phase margin ≥ 60◦ 75◦ 68◦ 72◦
Slew rate ≥ 10V /µs 4V /µs 12V /µs 12V /µs
DC power ≤ 50µW 122µW 38µW 39µW
Yield 0% 89% 99.9%
2.1 RC Circuit
Using an elementary RC circuit, we will illustrate the tasks of nominal design
and tolerance design. Figure 14 shows the circuit netlist and circuit variables.
Two performance features, the time constant τ and the area A, and two
circuit parameters, the resistor value R and the capacitor value C, are consid-
ered. Please note that R and C are normalized, dimensionless quantities. This
example is well suited as we can explicitly calculate the circuit performance in
dependence of the circuit parameters and as we can visualize the design situa-
tion in a two-dimensional performance space and a two-dimensional parameter
space.
τ = R·C
A = R+C
Figure 14. Elementary RC circuit with performance features time constant τ and area A as a
function of resistor value R and capacitor value C.
20 ANALOG DESIGN CENTERING AND SIZING
Figure 15 illustrates the situation of the RC circuit after nominal design. Figure
15(a) sketches the two-dimensional space spanned by the performance fea-
tures “area” and “time constant,” Figure 15(b) shows the corresponding two-
dimensional space spanned by the parameters R and C.
Tolerance Design: Example 21
Figure 15. RC circuit. (a) Performance values of RC circuit after nominal design. (b) Parameter
values of RC circuit after nominal design.
We can also see that although we have centered the performance values, we
have not centered the parameter values: R and C are closer to the double-
slashed curve than to the triple-slashed curve. This difference is due to the
nonlinearity of the time constant in the parameters. If we had centered the
parameter values instead of the performance values, using R = C = x, we
would not have centered x2 between 2, i.e. x2 = 0.5 · (0.5
√0.5 and √ √ + 2),√but
we would have centered x between 0.5 and 2, i.e. x = 0.5 · ( 0.5 + 2).
As a result, we would have designed the following parameter and performance
values: √
R = C = 0.75 · 2 ≈ 1.061
τ = 1.125 (4)
√
A = 1.5 · 2 ≈ 2.121
This difference between the situation in the performance space and the para-
meter space leads us to yield optimization/design centering.
Figure 16. Parameter values and parameter tolerances of the RC circuit after nominal design.
ellipses have decreasing line thickness. Had the standard deviations of R and
C been equal, the level curves would have been circles. σC being larger than
σR corresponds to a higher dilation of the level curves in the direction of C than
in the direction of R. The distribution of the resistor values hence is less broad
than that of the capacitor values. Figure 16 shows that parameter vectors exist
that lead to a violation of the performance specification. The volume under
the probability density function over the whole parameter space is 1, which
corresponds to 100% of the manufactured circuits. We can see that a certain
percentage of circuits will violate the performance specification. The percent-
age of manufactured circuits that satisfy the performance specification is called
parametric yield Y . Parametric variations refer to continuous variations within
a die or between dies of the manufacturing process. The resulting performance
variations can be within the acceptable limits or violate them. In addition, the
manufacturing process is confronted with catastrophic variations that result for
instance from spot defects on the wafer causing total malfunction and resulting
in the so-called catastrophic yield loss. The yield is an important factor of the
manufacturing cost: the smaller the yield, the higher the manufacturing cost.
In the rest of the paper, the term yield will refer to the parametric yield of a
circuit.
The estimated yield after nominal design in this example is:
Y = Y (R , C ) = 58.76% (6)
Chapter 6 will deal with methods to estimate the parametric yield. Figure 17
extends Figure 16 with a third axis for the value of the probability density
function to a three-dimensional view.
24 ANALOG DESIGN CENTERING AND SIZING
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0
4
1
3
2
2
3 1
4 0
Figure 17. Probability density function after nominal design of the RC circuit, truncated by the
performance specification.
Only that part of the probability density function that corresponds to accept-
able values of R and C is given. We can think of the probability density function
as a bell-shaped “cake” that describes the variation of the actual parameter vec-
tors due to the manufacturing process. A portion of varying parameter vectors
are cut away by the performance specification. An optimal design will lead to a
maximum remaining volume of the “cake,” which is equivalent to a maximum
parametric yield Ymax . The part of the design process that targets at maximum
yield is equally called yield maximization or design centering.
Figures 16 and 17 illustrate that the result of nominal design does not lead
to a maximum yield. The first reason for that is the mentioned nonlinearity of
the performance with respect to the parameters, which leads to skewing equal
safety margins of the performance with respect to its lower and upper bounds
in the parameter space. The second reason is the parameter distribution, which
Tolerance Design: Example 25
1 1
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0
4 4
1 1
3 3
2 2
2 2
3 1 3 1
4 0 4 0
Figure 18. (left) Parameter values and truncated probability density function. after nominal
design of the RC circuit. (right) Parameter values and truncated probability density function
after yield optimization/design centering of the RC circuit.
3.1 Parameters
In the literature, the term parameter usually denotes a variable quantity that
is related to the circuit structure and the circuit behavior. In this book, the term
parameters shall be used in a more specific sense for those quantities that are
input to the numerical simulation. Quantities related to the output of numerical
simulation shall be denoted as performance features. This distinction is owing
to the fact that the mapping of parameter values onto values of performance
features by means of numerical simulation takes the – computationally very
expensive – center stage of a systematic analog design process.
The manufactured ICs are arranged by quality classes, for instance of faster or
slower ICs, and usually sold at different prices. The IC that is actually bought
therefore can have a quite different performance from what is specified as its
nominal performance in the data sheet. In addition, its performance will dynam-
ically vary with altering operating conditions. These dynamic variations during
operation may also be larger or smaller, depending on the quality of the actual
IC. Operating conditions are usually modeled as intervals of range parameters.
Manufacturing fluctuations are modeled as statistical parameter distributions
on the regarded design level. The calculation of the statistical parameter distri-
bution is a difficult task that requires specific optimization algorithms and the
consistent combination of measurements that belong to different design levels
and design views [40, 29, 109, 87, 98].
polytope region
ellipsoid region
Tr,E = {xr (xr − xr,0 )T · Aellips · (xr − xr,0 ) ≤ b2ellips
(10)
Aellips ∈ Rnxr ×nxr , Aellips symmetric, positive definite
nonlinear region
Figure 19. (a) Box tolerance region. (b) Polytope tolerance region. (c) Ellipsoid tolerance
region. (d) Nonlinear tolerance region.
Figure 19 illustrates the four types of tolerance regions. The box region accord-
ing to (8), illustrated in Figure 19(a), is the prevalent type of range-parameter
tolerances. It results if minimum and maximum values for each range para-
meter are specified. The dashed line in Figure 19(a) for instance corresponds
to the upper bound of parameter xr,1 ≤ xr,U,1 . In many cases, the tolerance
region Tr will be closed, that means a finite range of values will be given for
each range parameter. This reflects practical design situations where neither
process parameters nor operating parameters should attain excessive values,
be it for physical limitations or limited resources. The oxide thickness as well
as the channel length for example are process quantities that cannot go below
certain values. On the other hand, the channel length will be bounded above
due to the limited chip area. For the operating range in turn, reasonable finite
ranges have to be negotiated between manufacturer and customer. The size of
this range is related to the product price: an IC that works between −40◦ C and
125◦ C will be more expensive than the same IC designed to work between 0◦ C
and 70◦ C.
If the range parameters’ bounds are determined in a linear dependence of
each other, the tolerance region will have the shape of a polytope according
to (9), as illustrated in Figure 19(b). (9) describes a set of npoly inequalities
of the form aTµ· · xr ≤ bpoly,µ , where aTµ· = [aµ,1 aµ,2 aµ,3 . . .]T is the µ-th
row of the matrix Apoly . Every inequality corresponds to a bounding plane
as indicated by the dashed line in Figure 19(b), determining a halfspace of
acceptable parameter vectors. The tolerance region Tr is the intersection of
Parameters & Tolerances, Performance & Specification 31
all these halfspaces. Inequalities may be redundant, leaving them out would
not alter the tolerance region Tr . The bounding line at the top of Figure 19(b)
corresponds to such a redundant inequality.
(10) defines a tolerance region that is shaped as an ellipsoid, as illustrated by
the ellipse in Figure 19(c). It will be detailed in the following Section 3.4 that
the level curves of a Gaussian probability density function are ellipsoids. If the
shape of a parameter region corresponds to the shape of the level contours of a
probability density function, we are able to complement the parameter region
with probability values for the occurrence. In this way, we allow a transition
between range parameters and statistical parameters.
(11), illustrated by Figure 19(d), finally provides the most general descrip-
tion of a nonlinear tolerance region described by single nonlinear inequalities,
ϕnonlin,µ (xr ) ≥ 0, as nonlinear bounding surfaces, as indicated by the dashed
line in Figure 19(d).
xs,1 xs,nxs
cdf(xs ) = ... pdf(t) · dt , dt = dt1 · dt2 · . . . · dtnxs (13)
−∞ −∞
∂ nxs
pdf(xs ) = cdf(xs ) (14)
∂xs,1 · ∂xs,2 · . . . · ∂xs,nxs
xs ∼ N (xs,0 , σ 2 ) (20)
1 Speaking of a random variable x, we immediately denote the VALUE x of a random variable X. This
value x refers e.g. to
a random number generated according to the statistical distribution of the random variable,
the argument in the probability density pdf(x),
x+dx
the infinitesimal range x + dx having a probability of occurrence prob(X ≤ x) = x pdf(t)dt =
cdf(x + dx) − cdf(x).
x
The probability that a random variable X has a value less or equal x is prob(X ≤ x) = −∞ pdf(t)dt =
cdf(x). Please note that the probability that a random variable X has a value of x is prob(X = x) =
x
x pdf(t)dt = 0!
Parameters & Tolerances, Performance & Specification 33
Table 5. Selected cumulative distribution function values of the univariate normal distribution.
Figure 20. Probability density function pdf and cumulative distribution function cdf of a uni-
variate normal distribution.
The relationship between tolerance ranges and yield values becomes more
complicated for multivariate distributions.
1
pdfN (xs ) = √ nxs √ · exp −0.5 · β 2 (xs ) (23)
2π · det C
β 2 (xs ) = (xs − xs,0 )T · C−1 · (xs − xs,0 ) (24)
β ≥ 0
Parameters & Tolerances, Performance & Specification 35
0.16
0.14 f(x,y)
0.15
0.12 0.1
0.05
0.1
0.08
0.06
0.04
0.02
0
-4
-3
-2
-1
0
1
2 3 4
3 1 2
-1 0
4 -4 -3 -2
Figure 21. Multivariate normal distribution for two parameters according to (23), (24) or (34),
respectively, with mean values xs,0,1 = xs,0,2 = 0, correlation = 0.5, and variances σ1 = 2,
σ2 = 1.
In (24), the vector xs,0 = [xs,0,1 . . . xs,0,k . . . xs,0,nx ]T denotes the vector of
mean values xs,0,k , i = 1, . . . , nxs , of the multivariate normal distribution of
the parameters xs,k , k = 1, . . . , nxs . Background material about expectation
values can be found in Appendix A. Figure 21 illustrates the probability density
function of a multivariate normal distribution for two parameters.
It shows the typical bell shape of the probability density function of two para-
meters and the quadratic form of the equidensity contours determined by (24).
The matrix C denotes the variance/covariance matrix, or covariance matrix,
of the parameter vector xs .
A random vector xs (see Note 1 in Section 3.5) that originates from a normal
distribution with mean value xs,0 and covariance matrix C is also denoted as:
xs ∼ N (xs,0 , C) (25)
C = Σ·R·Σ (26)
36 ANALOG DESIGN CENTERING AND SIZING
⎡ ⎤
σ1 0 ... 0
⎢ .. ⎥
⎢ 0 σ2 . ⎥
Σ = ⎢ . ⎥ (27)
⎣ .. ..
. 0 ⎦
0 ... 0 σnxs
⎡ ⎤
1 1,2 ... 1,nxs
⎢ .. ⎥
⎢ 1 . ⎥
R = ⎢ 2,1
.. ⎥ (28)
⎣ .
..
. nxs −1,nxs ⎦
nxs ,1 ... nxs ,nxs −1 1
⎡ ⎤
σ12 σ1 1,2 σ2 ... σ1 1,nsx σnxs
⎢ .. ⎥
⎢ σ σ1 σ22 . ⎥
C = ⎢ 2 2,1
.. ⎥ (29)
⎣ .
..
. σnxs −1 nxs −1,nxs σnxs ⎦
nxs ,1 ... nxs ,nxs −1 σn2 xs
σk2 is the variance of parameter xs,k . σk k,l σl is the covariance of parameters
xs,k , xs,l . For the variance, we have:
σk > 0 (30)
For the correlation, we have:
k,l = l,k (31)
−1 ≤ k,l ≤ +1 (32)
From (26)–(32) follows that the covariance matrix C is symmetric and pos-
itive semidefinite.
If k,l = 0, this is denoted as the two parameters xs,k , xs,l being uncorrelated.
In the case of a multivariate normal distribution, “uncorrelated” is equivalent
to “statistically independent.” In general, “statistically independent” implies
“uncorrelated.”
If two parameters xs,k , xs,l are perfectly correlated, then |k,l | = 1. Perfect
correlation refers to a functional relationship between the two parameters. For
each value of one parameter a unique value of the other parameter corresponds,
i.e. the statistical uncertainty between the two parameters disappears. In this
case, the covariance matrix has a zero eigenvalue and the statistical distribution
is singular, i.e. (23) cannot be written. In the following, we will assume that
|k,l | < 1 (33)
This means that C is positive definite and that the multivariate normal distribu-
tion is nonsingular. Asingular distribution can be transformed into a nonsingular
one [5].
Parameters & Tolerances, Performance & Specification 37
For two parameters, (23) and (24) have the following form:
xs,1 xs,0,1 σ12 σ1 σ2
xs = ∼ N xs,0 = , C=
xs,2 xs,0,2 σ1 σ2 σ22
1 1 (xs,1 − xs,0,1 )2
pdfN (xs,1 , xs,2 ) = · exp − ·
2πσ1 σ2 1 − 2 2(1 − 2 ) σ12
xs,1 − xs,0,1 xs,2 − xs,0,2 (xs,2 − xs,0,2 )2
−2 + (34)
σ1 σ2 σ22
If the two parameters are uncorrelated, i.e. = 0, and if their mean values
are xs,0,1 = xs,0,2 = 0, and if their variance values are σ1 = σ2 = 1, (34)
becomes:
1 1 2 2 1 1 T
pdfN (xs,1 , xs,2 ) = · exp − xs,1 + xs,2 = √ nxs · exp − xs xs (36)
2π 2 2π 2
The level curves of the probability density function of a normal distribution, i.e.
the equidensity curves, are determined by the quadratic form (24). Therefore,
the level curves of pdfN are ellipses for nxs = 2, ellipsoids for nxs = 3 (Figure
21) or hyper-ellipsoids for nxs ≥ 3. In the rest of this book, three-dimensional
terms like “ellipsoid” will be used regardless of the dimension.
Figure 22 illustrates different shapes of the equidensity curves of a normal
probability density function in dependence of variances and correlations. Each
equidensity curve is determined by the respective value of β in (24). For
β = 0, the maximum value of pdfN , given by √2π nxs1·√det C , is obtained at
the mean value xs,0 . Increasing values β > 0 determine concentric ellipsoids
centered around the mean value xs,0 with increasing value of β and decreasing
corresponding value of pdfN .
Figure 22(a) illustrates the situation if the two parameters are positively cor-
related. The major axes of the ellipsoids then have a tilt in the parameter space.
Note that the equidensity ellipsoids denote a tolerance region corresponding
to (10). If the two parameters are uncorrelated, then the major axes of the
ellipsoids correspond to the coordinate axes of the parameters, as illustrated
in Figure 22(b). If uncorrelated parameters have equal variances, then the
ellipsoids become spheres, as illustrated in Figure 22(c).
38 ANALOG DESIGN CENTERING AND SIZING
Figure 22. Different sets of level contours β 2 = (xs − xs,0 )T · C−1 · (xs − xs,0 ) (24) of
a two-dimensional normal probability density function pdfN . (a) Varying β, constant positive
correlation , constant unequal variances σ1 , σ2 . (b) Varying β, constant zero correlation,
constant unequal variances. (c) Varying β, constant zero correlation, constant equal variances.
(d) Varying correlation, constant β, constant unequal variances.
Parameters & Tolerances, Performance & Specification 39
While the left set in (37) describes an ellipsoid tolerance region according
to (10), the right set in (37) describes a box tolerance region according to
(8). The bounding-box-of-ellipsoids property (37) says that the union of all
ellipsoid tolerance regions that are obtained by varying the correlation value of
two parameters for a constant β yields exactly a box tolerance region ±βσk/l
around the mean value. A motivation of (37) is given in Appendix C.7.
The cumulative distribution function cdfN (xs ) of the multivariate normal
distribution can be formulated according to (13). A more specific form analo-
gous to (18) and (19) of the univariate normal distribution cannot be obtained.
Only if the correlations are zero, then the probability density function of the
multivariate normal distribution can be formulated as a product of the univari-
ate probability density functions of the individual parameters (see (35)) and we
obtain:
nxs
xs,k
nxs
R = I ⇒ cdfN (xs ) = pdfN (t)dt = cdfN (xs,k ) (38)
i=1 −∞ k=1
Therefore, we will assume that the statistical parameters are normally distributed.
But parameters are not normally distributed in general. For instance para-
meters like oxide thickness that cannot attain values below zero naturally have a
skew distribution unlike the symmetric normal distribution. These parameters
have to be transformed into statistical parameters whose distribution is normal.
The resulting transformed parameters of course have no longer their original
physical meaning. This is in fact not necessary for the problem formulation
of yield optimization/design centering. The reader is asked to keep in mind
that the normally distributed statistical parameters that are assumed in the rest
of this book may originate from physical parameter that are distributed with
another distribution.
Given two random vector variables, y ∈ Rny and z ∈ Rnz with ny = nz ,
which can be mapped onto each other with a bijective function2 , i.e. z(y) and
y(z), the probability density function pdfz of random variable z is obtained
from the probability density function pdfy of random variable y by:
∂y
pdfz (z) = pdfy (y(z)) · det (39)
∂zT
∂y
pdfz (z) = pdfy (y(z)) ·
∂z
2 z(y) is meant to be read as “z evaluated at y.” This implies that z is a function of y, i.e. z = ϕ(y). The
y = cdf−1
y (z) (43)
Example 2. This example shows how the probability density function pdfz
of random variable z can be derived if the probability density function pdfy of
random variable y and the transformation function y → z are given.
Given as an example are a random variable y originating from a normal
distribution (44) and an exponential function (45) that maps y onto z:
2
1 −1
y−y0
pdfy (y) = √ e 2 σ
(44)
2π · σ
z = ey (45)
We are looking for the probability density function pdfz (z). As the function
z = ey is bijective w.r.t. R → R+ , (39) is applied:
∂y ∂y 1
pdfz (z) = pdfy (y(z)) · with y = lnz and = yields
∂z ∂z z
2
1 −1
lnz−y0
pdfz (z) = √ e 2 σ
(46)
2π · σ · z
(46) denotes the pdf of a lognormal distribution. As can be seen from the
calculations above, the logarithm y = lnx of a lognormally distributed random
42 ANALOG DESIGN CENTERING AND SIZING
1 1 2
pdfy (y) = √ e− 2 y (47)
2π
z = y2 (48)
We are looking for the probability density function pdfz (z). The function
z = y 2 is not bijective, but surjective w.r.t. R → R+ . In this case we can add
the probabilities of all infinitesimal intervals dy that contribute
√ to the probability
√
of an infinitesimal interval dz. Two values, y (1) = + z and y (2) = − z, lead
to the same value z and hence two infinitesimal intervals in y contribute to the
probability of dz:
pdfz (z) · dz = pdfy y (1) · dy (1) + pdfy y (2) · dy (2) (49)
3A sample element is represented by a vector of parameter values x. It refers to the stochastic event of an
infinitesimal range of values dx around x.
44 ANALOG DESIGN CENTERING AND SIZING
Comparison of (52) and (53) with (54) and (55) finally yields
In the transformation (57) and (58), no specific assumption about the form
of the matrix A is included. Therefore, not only a Cholesky decomposition,
which yields a triangular matrix, can be applied, but also an eigenvalue
decomposition. As C is positive definite, we obtain:
1 1
C = VΛVT = VΛ 2 Λ 2 VT , (59)
1
where V is the matrix of eigenvectors and where Λ 2 is the diagonal matrix
of the roots of the eigenvalues. A comparison with (56) leads to:
1
A = VΛ 2 (60)
Note that the actual transistor parameter at the interface to the circuit simulator
may be a sum of a deterministic design parameter component xd,k , of a global
statistical parameter component xs,glob,l , and of a local statistical parameter
component xs,loc,m .
x → f (62)
where t denotes the time, x the parameters, un the node voltages, iZ the currents
in Z-branches, q the capacitor charges, φ the inductor fluxes, and ˙ the derivative
with respect to time. Performance evaluation in general is a map of a certain
parameter vector x onto the resulting performance feature vector f , which is
carried out in two stages. First, numerical simulation [97, 75, 121, 101, 89]
by means of numerical integration methods maps a parameter vector onto the
node voltages and Z-branch currents:
x → un , iZ (64)
un , iZ → f (65)
and the calculation of each function element costs minutes to hours CPU
time.
Algorithms for analog optimization must therefore very carefully spend numer-
ical simulations, just as oil exploration has to carefully spend the expensive test
drillings.
In order to avoid the high computational cost of numerical simulation, per-
formance models or macromodels for a cheaper function evaluation would be
welcomed. Performance models can be obtained through numerical methods
called response surface modeling [31, 39, 3, 76, 37, 124, 16, 125, 127, 43, 64,
78, 17, 4, 126, 14, 29] or through analytical methods like symbolic analysis.
Macromodels can be obtained by simplified DAEs [22], by symbolic analysis
[108, 112, 113, 30, 107, 106, 32, 117, 67, 120, 119, 44, 18, 114, 15, 51, 50, 53,
52], by equivalent circuits [49, 20], or by behavioral models [128, 54, 65, 71,
72, 82]. However, the generation of such models requires a large number of
numerical simulations. Only if the resulting models are applied often enough
the setup cost are worthwhile. We are not aware of investigations of the break-
even where pre-optimization modeling undercuts the cost of optimization on
direct numerical simulation.
ϕP SF,µ (f ) ≥ 0 (66)
ϕP SF,µ (f ) ≥ 0, µ = 1, . . . , nP SF (67)
48 ANALOG DESIGN CENTERING AND SIZING
Af = {f | fL ≤ f ≤ fU }
= {f | fi ≥ fL,i ∧ fi ≤ fU,i , i = 1, . . . , nf } (70)
nx
∆f = S · ∆x = ∇f (xk ) · ∆xk (72)
k=1
1 Please note that the Nabla symbol does not denote the Nabla operator here. In the abbreviated notation of
the partial derivative with the Nabla sign the parameters in brackets denote the parameters with regard to
which the partial derivative is calculated. Other parameters, which are held constant at certain values, are
not mentioned explicitly in this notation, but they determine the result value as well.
Analog Sizing Tasks 51
If we assume the same alteration ∆x for all parameters, then a ranking of the
parameters according to their significance is obtained by comparing the lengths
of the performance gradients with respect to the individual parameters:
xk more significant than xl ⇔ ∇f (xk ) > ∇f (xl ) (76)
4.1.6 Exercise
Given is the sensitivity matrix for two performance features f1 , f2 and two
parameters x1 , x2 at a certain point x0 (Figure 23):
1 2
S= (81)
−1 1
Analyze the design situation regarding similarity, significance, adjustability and
multiple-objective behavior.
Analog Sizing Tasks 53
The formulas to compute ∇un (xT ) and ∇iZ (xT ) require the corresponding
parts in the transistor models. This increases the transistor modeling effort,
as not only the transistor equations but also their first derivatives have to be
provided.
The formulas to compute f (uTn ) and ∇f (iTZ ) are part of the postprocessing
operations that have to include the corresponding first derivatives.
The main effort is the a-priori calculation of the formulas for the first deriva-
tives. This is done once for each transistor class of the underlying production
technology and once for the considered circuit class.
The simulation overhead for sensitivity computation remains relatively small.
It amounts to about 10% of the CPU time of one simulation for the computation
of the sensitivity of all performance features with respect to one parameter.
x → x , f → f (84)
Examples for reference points are lower bounds, upper bounds or initial values:
The scaled variables according to (87) will be in the interval between 0 and
1: xk , fi ∈ [0, 1]. The values xRP,L,k , xRP,U,k , fRP,L,i , fRP,U,i have to be
chosen very carefully. The performance specification (70) or tolerances range
of parameters (8) may be helpful at this point. If only an upper or a lower
bound of a variable is available, the proper choice of the opposite bound is very
difficult. Initial guesses are used in these cases, which have to be tightened or
relaxed according to the progress of the optimization process.
Statistical parameters scaled according to (88) are independent from each other
and have most probably values between −3 and 3. The resulting covariance
matrix is the unity matrix and has the best condition number, i.e. 1.
∆f , S · ∆x = ∆f → ∆x (89)
The solution of (89) requires the solution of a system of linear equations that
may be rectangular and rank-deficient. The computational accuracy of the
solution is determined by the condition number of the system matrix, cond(S) =
S−1 ·S. Gradients that differ very much cause an ill-conditioned sensitivity
matrix. An improvement is to equalize the sensitivities, which in turn leads to
a corresponding scaling of performance features and parameters.
There are two approaches to equalize the sensitivity matrix S.
The first approach is to equalize the rows of S to an equal length of 1:
f > f ∗ ⇔ f ≥ f ∗ ∧ f
= f ∗ ⇔ ∀i f i ≥ f ∗i ∧ ∃i f
i
= fi∗ (94)
f < f ∗ ⇔ f ≤ f ∗ ∧ f
= f ∗ ⇔ ∀i f i ≤ f ∗i ∧ ∃i f
i
= fi∗ (95)
(94) and (95) define that a vector is less (greater) than another vector if all its
components are less (greater) or equal and if it differs from the other vector.
58 ANALOG DESIGN CENTERING AND SIZING
Figure 24. (a) Set M > (f ∗ ) of all performance vectors that are greater than f ∗ , i.e. inferior to
f ∗ with regard to multiple-criteria minimization according to (92) and (93). (b) Set M < (f ∗ ) of
all performance vectors that are less than f ∗ , i.e. superior to f ∗ with regard to multiple-criteria
minimization.
This equivalently means that a vector is less (greater) than another vector if at
least one of its components is less (greater) than the corresponding component
of the other vector. In Figure 24(a), the shaded area indicates all performance
vectors that are greater than a given performance vector f ∗ for a two-dimensional
performance space.
These form the set M > (f ∗ ):
The solid line at the border of the shaded region is meant as a part of M > (f ∗ ).
The small kink of the solid line at f ∗ indicates that f ∗ is not part of M > (f ∗ )
according to the definition in (94). The solid line represents those performance
vectors for which one performance feature is equal to the corresponding com-
ponent of f ∗ and the other one is greater. In the shaded area in Figure 24(a),
both performance features have values greater than the corresponding star val-
ues. With respect to a multiple-objective minimization problem, performance
vectors in M > (f ∗ ) are inferior to f ∗ . Figure 24(b) illustrates the analogous
situation for all performance vectors that are less than, i.e. superior to, a given
performance vector f ∗ for a two-dimensional performance space. These form
the set M < (f ∗ ):
M < (f ∗ ) = { f | f < f ∗ } (97)
From visual inspection of Figure 24(b), we can see that the border of the set of
superior performance vectors M < (f ∗ ) to f ∗ corresponds to the level contour of
Analog Sizing Tasks 59
the l∞ -norm of a vector, which is defined as the max operation on the absolute
values of its components. The set of superior performance vectors M < (f ∗ )
to f ∗ can therefore be formulated based on the max-operation or the l∞ -norm,
if the performance vectors are assumed to be scaled such that they only have
n
positive values, f ∈ R+f :
nf < ∗ ∗
f ∈ R+ : M (f ) = f max (fi /fi ) ≤ 1 ∧ f
= f ∗
(98)
i
= {f | [ . . . fi /fi∗ . . . ] ∞ ≤ 1 ∧ f
= f ∗ } (99)
Based on (97)–(99), Pareto optimality is defined by excluding the existence of
any superior performance vector to a Pareto-optimal performance vector:
f ∗ is Pareto optimal ⇔ M < (f ∗ ) = { } (100)
Figure 25. (a) Pareto front P F (f1 , f2 ) of a feasible performance region of two performance
features. Different Pareto points addressed through different weight vectors from reference point
fRP , which is determined by the individual minima of the performance features. (b) Pareto front
P F (f1 , f2 , f3 ) of three performance features with boundary B = {P F (f1 , f2 ), P F (f1 , f3 ),
P F (f2 , f3 )}.
the set of all performance vectors that are achievable under consideration of the
constraints of the optimization problem. The solid line at the left lower border
of the region of achievable performance values represents the set of Pareto
points, i.e. the Pareto front. Each point on the front is characterized in that
no point superior to it exists, i.e. M < (f ∗ ) is empty. Three Pareto points are
illustrated and the corresponding empty sets M < (f ∗ ) are indicated. We can as
well see that each Pareto point represents the minimum weighted l∞ -norm with
respect to the reference point fRP with components according to (102) that is
achievable for a given weight vector w. Depending on the weight vector w, a
specific Pareto point on a ray from the reference point fRP in the direction w
is determined. The Pareto front P F and its boundary B can be defined as:
P F (f ) = {f ∗ (w) | wi ≥ 0 , i wi = 1 }
(103)
B(P F (f1 , . . . , fnf )) = i P F (f1 , . . . , fi−1 , fi+1 . . . fnf )
A Pareto front for three performance features and its boundary is shown in
Figure 25(b). Starting from the individual minima, the Pareto front can be
formulated hierarchically by adding single performance features according
to (103).
Pareto fronts may exhibit different shapes and may be discontinuous. Figure
26 shows three example shapes of Pareto fronts in relation to the convexity
Analog Sizing Tasks 61
∗
Figure 26. (a) Continuous Pareto front of a convex feasible performance region. fI,2 is mono-
∗ ∗
tone in fI,1 . (b) Continuous Pareto front of a nonconvex feasible performance region. fI,2 is
∗
monotone in fI,1 . (c) Discontinuous Pareto front of a nonconvex feasible performance region.
∗ ∗
fI,2 is nonmonotone in fI,1 .
of the feasible performance region and the monotonicity of the explicit Pareto
front function of one performance feature.
Figure 27. (a) Level contours of the weighted l1 -norm. (b) Level contours of the weighted
l2 -norm. (c) Level contours of the weighted l∞ -norm.
weighted l2 -norm
√ !
W=diag(wi )
h2,w = hT ·W·h = wi2 · h2i (106)
i
weighted l∞ -norm
h∞,w = max |wi | · |hi | (107)
i
i = 1, . . . , nf
If a lower bound fL,i is specified or required for a performance feature fi , the
worst-case represents the maximal deviation of the performance feature from
its nominal value in negative direction. This value is obtained by computing
the minimum value of the performance feature that occurs over the statistical
and range parameters xs , xr within their given tolerance regions Ts , Tr .
Analog Sizing Tasks 65
In each worst-case parameter vector xs,W L,i , xs,W U,i , where the performance
feature fi has its worst-case value fW L,i , fW U,i , the values of the other perfor-
mance features add to the worst-case performance vectors f (xW L,i ), f (xW U,i ).
A worst-case performance vector f (xW L,i ), f (xW U,i ) hence describes the situ-
ation of all performance features, if performance feature fi is at its individual
lower or upper worst-case. Figure 28 shows a graphical overview of the input
and output of this mapping.
Figure 29 illustrates the input and output of a worst-case analysis in the
spaces of design, statistical and range parameters and of performance fea-
tures. For graphical reasons, each of these spaces is two-dimensional in this
example. Figure 29(a) shows the input situation with nominal values for design
parameters xd , statistical parameters xs,0 , and range parameters xr , and with
tolerance regions of the range parameters Tr and of the statistical parameters Ts .
The tolerance regions have the typical shape of a box for the range parameters
according to (8) and of an ellipsoid for the statistical parameters according to
(10), (23) and (24). The tolerance region of statistical parameters has been
determined to meet a certain yield requirement YL according to Section 6.2.
Note that the design parameters do not change their values during worst-case
analysis. A worst-case analysis computes lower and upper worst-case values
for each of the two performance features. The results are illustrated in Figure
29(b) and consist of:
Analog Sizing Tasks 67
Figure 29. (a) Input of a worst-case analysis in the parameter and performance space: nominal
parameter vector, tolerance regions. (b) Output of a worst-case analysis in the parameter and
performance space: worst-case parameter vectors, worst-case performance values.
68 ANALOG DESIGN CENTERING AND SIZING
4.8.1 Yield
The yield Y can be defined as the probability that a manufactured circuit
satisfies the performance specification under all operating conditions:
Y = prob{ ∀
xr ∈ Tr
fL ≤ f (xd , xs , xr ) ≤ fU } (114)
0 ≤ Y ≤ 1 ≡ 0% ≤ Y ≤ 100% (115)
On the one hand, the yield is determined by the probability density function of
the statistical parameters (13), (14). If no performance specification was given,
then the yield would be 100% according to (15). Figure 31(a) shows a normal
probability density function with its ellipsoidal equidensity contours, which is
unaffected by any performance specification.
Analog Sizing Tasks 71
0.16
0.14 f(x,y)
0.15
0.12 0.1
0.05
0.1
0.08
0.06
0.04
0.02
0
-4
-3
-2
-1
0
1
2 3 4
3 1 2
-1 0
4 -4 -3 -2
Figure 31. (a) The volume under a probability density function of statistical parameters, which
has ellipsoid equidensity contours, corresponds to 100% yield. (b) The performance specification
defines the acceptance region Af , i.e. the region of performance values of circuits that are in
full working order. The yield is the portion of circuits in full working order. It is determined
by the volume under the probability density function truncated by the corresponding parameter
acceptance region As .
The volume under this probability density function is 1, which refers to 100%
yield. On the other hand, the yield is determined by the performance specifica-
tion (70), (67) and the range parameters’ tolerance region (8)–(11). These lead
to a yield loss because a certain percentage of the statistically varying para-
meters will violate the performance specification for some operating condition.
Figure 31(b) illustrates how the ellipsoidal equidensity contours of the proba-
bility density function of statistical parameters are truncated by the parameter
acceptance region As , which corresponds to the performance acceptance region
Af defined by the performance specification. Figure 17 earlier illustrated the
72 ANALOG DESIGN CENTERING AND SIZING
Af = {f | fL ≤ f ≤ fU } (118)
Af,L,i = {f | fi ≥ fL,i } (119)
Af,U,i = {f | fi ≤ fU,i } (120)
)
Af = Af,L,i ∩ Af,U,i (121)
i=1,...,nf
Analog Sizing Tasks 73
* +
As,L,i =
xs ∀ f (x , x ) ≥ fL,i
xr ∈ Tr i s r
(123)
* +
As,U,i =
xs ∀ f (x , x ) ≤ fU,i
xr ∈ Tr i s r
(124)
)
As = As,L,i ∩ As,U,i (125)
i=1,...,nf
The definitions of the parameter acceptance region and its partitions include
that the respective performance-specification features have to be satisfied for all
range-parameter vectors within their tolerance region. This more complicated
formulation is due to the fact that the tolerance region of range parameters has
the form of a performance specification but is an input to circuit simulation.
Figure 32 illustrates the parameter acceptance region and its partitioning
according to performance-specification features. This partitioning can be of
use for an approximation of As in (116). In Figure 32, the complementary
“non-acceptance” region partitions are shown as well. The performance non-
acceptance region Āf and its partitions are defined as:
* +
Āf =
f
i i
∃ f < fL,i ∨ fi < fU,i (126)
The parameter non-acceptance region Ās and its partitions are defined as:
* +
Ās =
xs ∃ ∃ fi (xs , xr ) < fL,i ∨
xr ∈ Tr i ∨fi (xs , xr ) > fU,i
(130)
* +
Ās,L,i =
xs ∃ f (x , x ) < fL,i
xr ∈ Tr i s r
(131)
74 ANALOG DESIGN CENTERING AND SIZING
Figure 32. Parameter acceptance region As partitioned into parameter acceptance region par-
titions, As,L,1 , As,U,1 , As,L,2 , As,U,2 , for four performance-specification features, f1 ≥ fL,1 ,
f1 ≤ fU,1 , f2 ≥ fL,2 , f2 ≤ fU,2 . As results from the intersection of the parameter acceptance
region partitions.
Analog Sizing Tasks 75
* +
Ās,U,i =
xs ∃ f (x , x ) > fU,i
xr ∈ Tr i s r
(132)
,
Ās = Ās,L,i ∪ Ās,U,i (133)
i=1,...,nf
1 , xs ∈ As 1 , f ∈ Af
δ(xs ) = = (134)
0 , xs ∈ Ās 0 , f ∈ Āf
1 , xs ∈ As,L,i
δL,i (xs ) = (135)
0 , xs ∈ Ās,L,i
1 , xs ∈ As,U,i
δU,i (xs ) = (136)
0 , xs ∈ Ās,U,i
the yield can be formulated as the expectation value of the acceptance function
according to Appendix A:
+∞ +∞
Y = ... δ(xs ) · pdf(xs ) · dxs = E {δ(xs )} (137)
−∞ −∞
Y represents the overall yield, YL,i and YU,i represent the yield partitions
of the respective performance-specification features. An estimation of the
yield according to (137)–(139) is based on statistical estimators according to
Appendix B.
The yield partitions allow a ranking of the individual performance-
specification features concerning their impact on the circuit robustness. Figure
32 illustrates that each added performance-specification feature usually leads
to another yield loss. The smallest specification-feature yield value hence is an
upper bound for the overall yield Y .
Performance-specification features
fL,i , fU,i , i = 1, . . . , nf Yield Y
Tolerance region Tr → Yield Partitions
Nominal design YL,i , YU,i , i = 1, . . . , nf
x = [ xTd xTs,0 xTr ]T
(140)
This mapping is illustrated in Figure 33.
The yield analysis can be related to a worst-case analysis, which was
illustrated in Figure 28. The worst-case performance values from a worst-case
analysis become upper or lower performance-feature bounds as input of a yield
analysis. The yield value that is obtained as an output of yield analysis becomes
an input yield requirement of the worst-case analysis where it is transformed to
a tolerance region of statistical parameters according to Section 6.2.
Figure 34. Yield optimization/design centering determines a selected point of Pareto front of
performance features.
Figure 36. (a) Initial situation of yield optimization/design centering by tuning of design para-
meters xd that are disjunct from statistical parameters. (b) After yield optimization/design
centering by tuning of design parameters xd that are disjunct from statistical parameters. The
parameter acceptance region As depends on the values of design parameters xd . The equidensity
contours of a normal probability density function are ellipsoids according to (24).
Note that maximum yield does not equally mean a maximum tolerance region
inside the acceptance region.
The picture looks different if the design parameter space and the statistical
parameter space are identical. In that case, the parameter acceptance region
As will be constant. Yield optimization/design centering then happens through
tuning of the statistical parameter distribution, which basically concerns the
mean values, variances, correlations or higher-order moments. The first choice
of yield optimization/design centering in this case is to tune the mean value
xs,0 :
Figure 37. (a) Yield optimization/design centering by tuning of the mean values of statistical
parameters xs,0 . Parameter acceptance region As is constant. (b) Yield optimization/design
centering by tuning of the mean values, variances and correlations of statistical parameters xs,0 ,
C (tolerance assignment). Parameter acceptance region As is constant. Level contours of the
normal probability density function (24) change their shape due to changes in the covariance
matrix C.
From the definition of the parameter acceptance region and its partitions
in (122)–(124) follows that the yield depends on the range-parameter bounds
xr,L,k , xr,U,k , k = 1, . . . , nxr as well.
Therefore, a yield sensitivity and yield improvement can be formulated with
regard to the range-parameter bounds as another type of tolerance assignment.
Tolerance assignment plays a role in process tuning.
Tolerance assignment is also applied for the selection of discrete parameters
with different tolerance intervals, like for instance resistors with ±1% tolerances
or ±10% tolerances. Here the goal is to select the largest possible tolerance
intervals without affecting the yield in order to save production costs.
Note that the yield also depends on the specified performance-feature bounds
fL,i , fU,i , i = 1, . . . , nf , which can become a subject of yield optimiza-
tion/design centering. Here the goal is to select the best possible performance
specification that can be guaranteed with a certain yield.
Figure 38. To go beyond 99.9% yield for maximum robustness, yield optimization/design
centering requires specific measures.
value. It is intuitively clear that this will find a solution that can go beyond a
yield of 99.9%.
A corresponding formulation of (146) for yield optimization/design center-
ing, which only adjusts the nominal parameter vector and leaves the second-
order moments unchanged, is:
c(xs,0 ) ≥ 0
max det(a · C) s.t. (147)
xs,0 , a Y (xs,0 , a · C) = const ≡ 50%
The solution of (147) will lead to a picture as in the lower right half of
Figure 38.
A yield optimization/design centering is only complete if tries to go beyond
yield values of 99.9% as described above. The geometric approach to yield opti-
mization/design centering, described in Section 7.2, does not have the property
of a weak optimum illustrated in Figure 38. It therefore leads to an optimum
as illustrated on the lower right side of this figure without further ado.
Chapter 5
WORST-CASE ANALYSIS
In (150) and (151) the index i denoting the ith performance feature has been
left out. (150) and (151) can be itemized concerning any performance feature
and any type or subset of parameters.
Figure 39 illustrates the classical worst-case analysis problem in a two-
dimensional parameter space for one performance feature. The gray area is
the box tolerance region defined by a lower and upper bound of each para-
meter, xr,L,1 , xr,U,1 , xr,L,2 , xr,U,2 . The dotted lines are the level contours
according to the gradient ∇f (xr,0 ) of a linear performance model, which are
equidistant planes. Each parameter vector on such a plane corresponds to a cer-
tain performance value. The plane through the nominal point xr,0 corresponds
to the nominal performance value f0 . As the gradient points in the direction of
steepest ascent, the upper and lower worst-case parameter vectors xr,W U and
Worst-Case Analysis 87
xr,W L can readily be marked in. They correspond to the level contours of f
that touch the tolerance region furthest away from the nominal level contour.
This will usually happen in a corner of the tolerance box. These level contours
through the worst-case parameter vectors represent the worst-case performance
values fW U and fW L . The problem formulations (150) and (151) describe a
special case of a linear programming problem that can be solved analytically.
∇f (xr,0 ) − λL + λU = 0 (153)
λL,k · (xr,k − xr,L,k ) = 0 , k = 1, . . . , nxr (154)
λU,k · (xr,U,k − xr,k ) = 0 , k = 1, . . . , nxr (155)
88 ANALOG DESIGN CENTERING AND SIZING
(153) results from the condition that ∇L(xr ) = 0 must hold for a stationary
point of the optimization problem. (154) and (155) represent the complemen-
tarity condition of the optimization problem.
As either only the lower bound xr,L,k or the upper bound xr,U,k of a parameter
xr,k can be active but not both of them, we obtain from (153) and the property
that a Lagrange factor at the solution xr,W L is positive:
It depends on the sign of the gradient if (156) or (157) holds. Inserting either
(156) in (154) or (157) in (155) leads to the formula of an element of a worst-case
parameter vector for a worst-case lower performance value xr,W L,k . Analo-
gously an element of a worst-case parameter vector for a worst-case upper
performance value xr,W U,k can be derived.
⎧
⎪
⎨ xr,U,k , ∇f (xr,0,k ) > 0
xr,W U,k = xr,L,k , ∇f (xr,0,k ) < 0 (159)
⎪
⎩ undefined, ∇f (xr,0,k ) = 0
Figure 40. Classical worst-case analysis with undefined elements xr,W L,1 , xr,W U,1 , of worst-
case parameter vectors xr,W L , xr,W U .
The index L/U means that the corresponding formula holds once for a lower
performance-feature bound and once for an upper performance-feature bound.
Figure 41. Normal probability density function of a manufactured component splits into trun-
cated probability density functions after test according to different quality classes.
this is done. The dotted lines are the level contours according to the gradient
∇f (xs,0 ) of a linear performance model, which are equidistant planes. Each
parameter vector on such a plane corresponds to a certain performance value.
The plane through the nominal point xs,0 corresponds to the nominal perfor-
mance value f0 .
As the gradient points in the direction of steepest ascent, the upper and lower
worst-case parameter vectors xs,W U and xs,W L can readily be marked in Figure
42. They correspond to the level contours of f that touch the tolerance region
furthest away from the nominal level contour. This happens somewhere on the
border of the ellipsoid. The level contours through the worst-case parameter
vectors represent the worst-case performance values fW U and fW L .
The problem formulations (163) and (164) describe a special programming
problem with a linear objective function and a quadratic inequality constraint.
= f0 − βW · σf¯ (177)
.
fW U = f0 + βW · ∇f (xs,0 )T · C · ∇f (xs,0 ) (178)
= f0 + βW · σf¯ (179)
in Section 3.8.
f¯i ∼ N f0,i , σf2¯i (180)
The mean value of the probability density function is the nominal performance
value in (162), the variance is given by (175).
From (177), (179) and (180), and applying (17), a unique relationship between
the yield of a performance feature and the value of βW , which characterizes the
ellipsoid tolerance region, can be established:
2
f¯i −f0,i
1 − 12
pdff¯i (f¯i ) = √
σf¯
e i (181)
2π · σf¯i
In (186) and (187) the index i denoting the ith performance feature has been
left out. (186) and (187) can be itemized concerning any performance feature
and any type or subset of parameters.
Figure 43 illustrates the general worst-case analysis problem in a two-
dimensional parameter space for one performance feature. The gray area is
the ellipsoid tolerance region defined by βW , which is determined such that
the worst-case represents a given yield requirement. This is done according
to (182)–(184). For a three-sigma design for instance, βW = 3 would be
chosen. Section 5.5 motivates why (182)–(184) are well-suited in the general
nonlinear case.
The dotted lines are the level contours of the nonlinear performance feature.
Each parameter vector on such a level contour corresponds to a certain perfor-
mance value. The contour line through the nominal point xs,0 corresponds to
the nominal performance value f0 . In this example, the performance is uni-
modal in the parameters and its minimum value is outside of the parameter
tolerance ellipsoid.
The upper and lower worst-case parameter vectors xs,W U and xs,W L can
readily be marked in. They correspond to the level contours of f that touch the
tolerance region furthest away from the nominal level contour. This happens
somewhere on the border of the ellipsoid in this case. These level contours
through the worst-case parameter vectors represent the worst-case performance
values fW U and fW L .
The property that performance level contour and tolerance region border
touch each other in the worst-case parameter vector corresponds to a plane
tangential both to the performance level contour and the tolerance region border.
This tangent can be interpreted by a linearized performance model f¯ in the
worst-case parameter vector. The linearization is based on the gradient of the
performance in the worst-case parameter vector.
96 ANALOG DESIGN CENTERING AND SIZING
Figure 43. General worst-case analysis, in this case the solution is on the border of the tolerance
region.
(190) results from the condition that ∇L(xs ) = 0 must hold for a station-
ary point of the optimization problem. (191) represents the complementarity
condition of the optimization problem.
The second-order optimality condition holds because ∇2 L(xs ) = 2 · λ · C−1
is positive definite, as C−1 is positive definite and as λW L ≥ 0.
We assume that the solution is on the border of the ellipsoid tolerance region.
This happens for instance if the performance function f (xs ) is unimodal and
if its maximum or minimum is outside of the parameters tolerance region. The
constraint in (189) is therefore active:
λW L > 0 (192)
(xs,W L − xs,0 )T · C−1 · (xs,W L − xs,0 ) = βW
2
(193)
(190), (192) and (193) have the same form as the first-order optimality con-
dition of the realistic worst-case analysis (166)–(168). The general worst-case
parameter vectors and general worst-case performance values therefore also
have the same form as those of the realistic worst-case analysis. The only dif-
ference is in the gradient of the performance. One overall performance gradient
appears in the realistic worst-case analysis, whereas an individual performance
gradient at each worst-case parameter vector appears in the general worst-case
analysis.
βW
xs,W U − xs,0 = + - · C · ∇f (xs,W U ) (196)
∇f (xs,W U )T · C · ∇f (xs,W U )
βW
= + · C · ∇f (xs,W U ) (197)
σf¯(W U )
In (195) and (197) the variance σf2¯(W L/U ) of a performance function f¯(W L/U )
that is linear (188) in normally distributed parameters (25) has been inserted:
σf2¯(W L/U ) = ∇f (xs,W L/U )T · C · ∇f (xs,W L/U ) (198)
The mean value of the probability density function is the value of the linearized
performance function according to (199) in the nominal parameter vector, the
variance is given by (198).
From (201), (203) and (204), and applying (17), a unique relationship
between the yield of the performance feature linearized at the worst-case para-
meter vector and the value of βW , which characterizes the ellipsoid tolerance
region can be established:
⎛ ⎞2
(W L/U ) ¯(W L/U )
f¯ −f
1⎝ i 0,i ⎠
−2
(W L/U ) 1 σ (W L/U )
f¯
pdff¯(W L/U ) (f¯i )= √ e i (205)
i 2π · σf¯(W L/U )
i
fW U,i
(W U ) (W U )
ȲU,i = pdff¯(W U ) (f¯i ) · df¯i
−∞ i
(W U )
fW U,i −f¯
0,i
σ (W U ) 1 1 2
= f¯
i √ e− 2 t · dt
−∞ 2π
βW
1 1 2
= √ e− 2 t · dt (206)
−∞ 2π
100 ANALOG DESIGN CENTERING AND SIZING
= ȲL,i (207)
ȲL,i and ȲU,i are approximated yield partition values regarding the lower and
upper worst-case performance value fW L/U,i of performance feature fi .
This equation opens up the possibility of a technical interpretation of βW as
the measure that relates an ellipsoid tolerance region to an approximated yield.
We call βW the worst-case distance between the nominal value and the worst-
case value of a linearized performance feature according to (188). While the
worst-case value of the real performance feature and the linearized performance
feature are equal, the nominal value of the linearized performance feature differs
from the real nominal value according to (199). According to (201) and (203),
the worst-case distance is measured in the unit “variance of the linearized per-
formance”: the worst-case performance value is βW times the variance away
from the nominal linearized performance value. This variance is determined
by (198).
βW = 3 therefore refers to a three-sigma safety margin (three-sigma design)
of a performance feature, and βW = 6 refers to a six-sigma safety margin
(six-sigma design).
As a multiple of a performance variance, it can immediately be translated into
an approximated yield value for one performance feature according to (206) and
(207) and Table 5. Vice versa, a yield value can be translated approximately into
a value of βW , which determines an ellipsoid tolerance region in a multivariate
parameter space according to (191).
(188) has been applied to get from the yield definition in the performance space
(208), (210) to the yield definition in the parameter space (209), (211).
pdfN is the normal probability density function according to (23).
On the other hand, the true yield partition values, YL,i with regard to a lower
worst-case performance-feature value fW L,i , and YU,i with regard to an upper
worst-case performance-feature value fW U,i , can be formulated as:
YU,i = ... pdfN (xs ) · dxs (212)
fi (xs )≤fW U,i
YL,i = ... pdfN (xs ) · dxs (213)
fi (xs )≥fW L,i
(209), (211), (212) and (213) show that the error in the yield approximation
results from the approximation of the parameter acceptance region concerning
a performance-specification feature.
Figure 44 illustrates the situation for the example in Figure 43. In Figure
44(a), the upper worst-case has been picked. The ellipsoid tolerance region
corresponds to the given value of βW . The real level contour of all parameter
vectors that lead to the worst-case performance value fW U is shown by as a
dotted curve. It represents the border of the acceptance region that determines
the true yield partition value YU . This acceptance region is shaded in gray.
In addition, the level contour of all parameter vectors that lead to fW U
according to the linear performance model f¯(W U ) is shown as a solid line. It is
tangential to the true level contour in the worst-case parameter vector xs,W U ,
because f¯(W U ) has been specifically established in xs,W U by the general worst-
case analysis. It represents the border of the approximate acceptance region
that determines the approximate yield partition value ȲU . This region is filled
with a linen pattern.
The integration of the probability density function over the difference region
between the linen-pattern-filled area of the approximate parameter acceptance
region partition and the gray area of the true parameter acceptance region par-
tition determines the error of the yield approximation. In Figure 44(a), the
true yield will be overestimated. In Figure 44(b), the lower worst-case has
102 ANALOG DESIGN CENTERING AND SIZING
Figure 44. (a) Comparison between true parameter acceptance region partitions (gray areas)
and approximate parameter acceptance region partitions (linen-pattern-filled areas) of a general
worst-case analysis of an upper worst-case performance value. (b) Lower worst-case perfor-
mance value. (c) Lower and upper worst-case performance value.
Worst-Case Analysis 103
Figure 45. Duality principle in minimum norm problems. Shown are two acceptance regions
(gray), the respective nominal points within acceptance regions and two points on the border of
each acceptance region. Points (a) and (d) are the worst-case points.
been picked. Here, the true yield will be underestimated. Figure 44(c) shows
the overall approximate acceptance region if the lower and upper worst-case
parameter vectors and their linearizations are combined.
By inspection, the following statements concerning the approximation error
result:
Considering the decreasing values of the probability density function orthog-
onal to its ellipsoid level contours, the yield approximation error becomes
more critical in regions that correspond to ellipsoids closer to the center.
The chosen linearization is very well adapted to this property. It is exact in
the worst-case parameter vector, where the value of the probability density
function is maximum among the difference region. It loses precision in
modeling the exact level contour proportionally to the decrease in the corre-
sponding probability density function values. Apparently the approximation
is very accurate therefore.
Practical experience shows that the approximation error usually is about
1%–3% absolute yield error.
The duality principle in minimum norm problems says that the minimum
distance of a point to a convex set is equivalent to the maximum distance of
the point to all planes that separate the point from the convex set. Figure 45
illustrates the duality principle. In this example, we have assumed without
loss of generality that the variances are equal and that the correlations are
zero. Then the level contours are spheres.
On the left side of Figure 45, point (a) is the worst-case point for the given
nominal point inside the gray acceptance region. The duality principle in
this case applies for the distance of the nominal point to the complement of
104 ANALOG DESIGN CENTERING AND SIZING
the acceptance region, which is convex. It says that point (a) provides the
linearization among all possible points on the border of the acceptance region
from which the nominal point has maximum distance. This is illustrated by
point (b), which leads to a linearization with a smaller distance. We can
also see that any tangent on the acceptance region’s border will lead to
an underestimation of the true yield. Therefore the worst-case distance
obtained due to point (a) is a greatest lower bound among all tangential
approximations of the acceptance region.
On the right side Figure 45, the acceptance region itself is convex now.
We can see that point (c) leads to a larger distance between the nominal
point and the tangential approximation of the acceptance region’s border
than the worst-case point (d). Now the duality principle says that the worst-
case distance will be the smallest value among all possible distances of the
nominal point to tangents of the border of the acceptance region. At the
same time, it can be seen that the yield value will be overestimated by such
a tangent. Therefore the worst-case distance obtained due to point (d) is a
least upper bound among all tangential approximations of the acceptance
region.
In summary, the linearization at a worst-case parameter vector provides
the best yield approximation among all tangential planes of an acceptance
region.
5.6 Exercise
Given is a single performance function of two parameters:
f = xs,1 · xs,2 (214)
f could be for instance the time constant of the RC circuit in Chapter 2. The
nominal parameter vector is:
xs,0 = [ 1 1 ]T (215)
The parameters are normally distributed with the covariance matrix:
0.22 0
C= (216)
0 0.82
Two performance-specification features are given:
f ≥ fL ≡ 0.5 (217)
f ≤ fU ≡ 2.0 (218)
Perform a classical worst-case analysis in the three-sigma tolerance box of
parameters based on a linear performance model established at the nominal
parameter vector. Calculate the worst-case parameter vectors and corre-
sponding worst-case performance values. Compare the worst-case perfor-
mance values from the linearized performance model with the “simulated”
values at the worst-case parameter vectors. Check if the performance spe-
cification is satisfied in the worst-case.
106 ANALOG DESIGN CENTERING AND SIZING
YIELD ANALYSIS
In this chapter, the yield analysis problem formulated in Section 4.8 will
be further developed. The integral in the yield definitions (116) and (117), as
well as in (137), (138) and (139) cannot be solved analytically, but have to be
solved numerically. Two approaches to solve this task will be described in the
following.
The first approach is based on statistical estimation by sampling according
to the parameter distribution. It leads to the so-called Monte-Carlo analysis,
which is a statistical technique for the numerical computation of integrals. It
will be explained that the accuracy of the statistical estimation does not depend
on the number of parameters and performance features, but on the size of the
sample. An acceptable accuracy requires a large number of simulations.
The second approach is based on the partitioning of the performance speci-
fication into the individual performance-specification features and a geometrical
approximation of the integration problem. The resulting problem has the same
form as the general-worst case analysis described in Section 5.4. This approach
is in practice more efficient than a Monte-Carlo technique, but its complexity
grows proportionally to the number of parameters.
s ∼ D (pdf(xs )) , i = 1, . . . , nM C
x(µ)
Figure 46. Statistical yield estimation (Monte-Carlo analysis) consists of generating a sam-
ple according to the underlying statistical distribution, simulating of each sample element and
flagging of elements satisfying the performance specification.
1, h1 (x) ≥ 0
hA (x) =
0, h1 (x) < 0
110 ANALOG DESIGN CENTERING AND SIZING
h(x)
I/ = /
E hA (x) ·
pdfM C (x)
pdfM C (x)
n
1 MC
h(x(µ) )
= hA (x(µ) ) (222)
nM C
µ=1
pdfM C (x(µ) )
The choice of the probability density function is important for the estimation.
But even if it is predetermined as in a yield analysis, another distribution may
be better for the yield estimation with a Monte Carlo analysis. This is called
importance sampling.
∞ ∞
Y = E {δ(xs )} = δ(xs ) · pdf(xs ) · dxs
...
pdf(xs ) −∞ −∞
∞ ∞
pdf(xs )
= ... δ(xs ) · · pdfM C (xs ) · dxs
−∞ −∞ pdfM C (xs )
Yield Analysis 111
pdf(xs )
= E δ(xs ) · (223)
pdfM C (xs ) pdfM C (xs )
pdf(xs )
Y/ = /
E δ(xs ) ·
pdfM C (xs )
pdfM C (xs )
n
MC (µ)
1 pdf(xs )
= δ(x(µ)
s ) (µ)
(224)
nM C pdfM C (xs )
µ=1
x(µ)
s ∼ D (pdfM C (xs )) , i = 1, . . . , nM C
The change of the original distribution, for which the expectation value of a
function is to be computed, to another sampling distribution leads to a weighting
of the function with the ratio of the two distributions.
Importance sampling allows to use arbitrary sampling distributions for the
yield estimation in order to improve the quality of the yield estimation.
A measure of the yield estimation quality is the variance of the yield estima-
tion value.
(B.15) 1 / {δ(xs )}
= ·V
nM C
1 nM C
(B.18)
= · · E{δ / 2 {δ(xs )}
/ 2 (xs )} − E
nM C nM C − 1
δ 2 =δ 1
= · (Y/ − Y/ 2 )
nM C − 1
Y/ · (1 − Y/ )
= (226)
nM C − 1
(225) and (226) show that the accuracy of a statistical yield estimation by
Monte-Carlo analysis primarily depends on the size of the sample, but also on
the yield value itself.
Obviously, the variance of the yield estimator σ 2
is quadratical in the yield
Y
value. From the first and second-order optimality conditions (Appendix C),
∂σ 2
1
Y
= (1 − 2Y ) ≡ 0 (227)
∂Y nM C
∂ 2 σ 2
2
Y
= − < 0 (228)
∂Y 2 nM C
follows that the variance of the yield estimator has a maximum for a yield of
50%:
arg max σY
= 50% (229)
Y
Table 7. Standard deviation of the yield estimator if the yield is 85% for different sample sizes
evaluated according to (225).
/Y
) − cdf(Y/ − kγ · σ
γ = cdf(Y/ + kγ · σ /Y
) → kγ (230)
kγ is determined without knowing the yield estimate and yield estimate variance
through the corresponding normalized univariate normal distribution.
For instance, a confidence level of γ = 90% denotes a probability of 90% for
the yield estimate to be within an interval of ±1.645/ σY
around the estimated
value Y/ . A confidence level of γ = 95% denotes a probability of 95% for
the yield estimate to be within an interval of ±1.960/ σY
around the estimated
value Y/ . γ = 99% corresponds to ±2.576/ σY
, and γ = 99.9% corresponds to
±3.291/ σY
.
114 ANALOG DESIGN CENTERING AND SIZING
Table 8. Required sample size for an estimation of a yield of 85% for different confidence
intervals and confidence levels according to (232).
∆Y = kγ · σ
/Y
(231)
Inserting the obtained value kγ and (231) in (225), the required sample size
nM C for the given confidence level γ → kγ and interval Y/ ± ∆Y can be
computed:
Y · (1 − Y ) · kγ2
nM C ≈ (232)
∆Y 2
For a yield of 85%, Table 8 evaluates (232) for three different confidence in-
tervals, 75% . . . 95%, 80% . . . 90%, and 84% . . . 86%, and for four different
confidence levels, 90%, 95%, 99%, and 99.9%.
Table 8 shows for instance that a Monte Carlo analysis with 8,461 sample
elements is required if we want have a 99% probability that the yield is within
±1% around its estimated value of 85%. It can be seen that thousands of sample
elements are required for a sufficient accuracy of the statistical yield estimation.
According to (225), to increase the accuracy by a factor of F , the sample size
has to be increased by a factor of F 2 . This is illustrated in Table 8, where
the increase in accuracy by a factor of 10 from ∆Y = 10% to ∆Y = 1%
corresponds to an increase in the sample size by a factor of 100.
Each sample element has to be evaluated by simulation. Simulation is com-
putationally very expensive and exceeds by far the computational cost of the
remaining operations. The computational cost of a Monte Carlo analysis is
therefore mainly determined by the number of simulations, i.e. the sample
size.
Yield Analysis 115
√
Accuracy ∼ nM C
Complexity ∼ nM C
Interestingly, the accuracy does not depend on any other quantity like for
instance the nonlinearity of the performance, and the complexity does not
depend on any other quantity like for instance the number of parameters.
Table 9. Tolerance intervals Ts,I and corresponding yield values YI according to (235).
Ts,I = [βa σ, βb σ] YI
[ −∞ , −1σ ] 15.9%
Ts,I = [βa σ, βb σ] YI
[ −∞ , 0 ] 50.0%
[ −1σ , 1σ ] 68.3% [ −∞ , 1σ ] 84.1%
[ −2σ , 2σ ] 95.5% [ −∞ , 2σ ] 97.7%
[ −3σ , 3σ ] 99.7% [ −∞ , 3σ ] 99.9%
Figure 47. Tolerance box Ts,B and normal probability density function pdfN,0R with zero
mean and unity variance.
ak − xs,0,k bk − xs,0,k
βak = , βbk = (239)
σk σk
pdfN denotes the probability density function of the multivariate normal distri-
bution as defined in (23)–(29). Using the variable transformation (238) and the
decomposition (26) of the covariance matrix, we obtain parameters t that are
normally distributed each with a mean value of 0 and a variance of 1, and that
are mutually correlated with the correlation matrix R. The probability density
function of the resulting normal distribution N (0, R) is denoted as pdfN,0R .
The variable transformation (238) also transforms the integration bounds ak
and bk into their distances to their nominal values as multiples of their standard
deviations, βak and βbk .
(237) represents an integral over the probability density function pdfN,0R in
a box Ts,B . Figure 47 illustrates level contours of pdfN,0R and Ts,B for two
parameters. In this example, βak = −βbk = −βW has been chosen. The
resulting box is symmetrical around the origin.
(237) has to be solved numerically. Only if the parameters are uncorrelated,
(237) can be evaluated by using the cumulative distribution function values of
the univariate normal distribution:
nxs βb
k 1 1 2
R=I : YB = √ e− 2 t · dt (240)
k=1 βak
2π
118 ANALOG DESIGN CENTERING AND SIZING
Table 10. Yield values YB for a tolerance box Ts,B with ∀k βak = −βbk = −βW = −3 in
dependence of the number of parameters nxs and of the correlation. ∀k=l k,l = 0.0 according
to (241), ∀k=l k,l = 0.8 according to (237).
3 99.1% 99.3%
4 98.8% 99.2%
5 98.5% 99.1%
6 98.2% 99%
7 97.9% 98.9%
8 97.6% 98.85%
9 97.3% 98.8%
10 97.0% 98.7%
⎫
R=I ⎪
⎬ βb nxs
1 1 2
∀k=l βak = βal = βa : YB = √ e− 2 t · dt (241)
⎪
⎭ βa 2π
∀k=l βbk = βbl = βb
Table 10 shows the yield value YB corresponding to a tolerance box Ts,B with
∀k βak = −βbk = −βW = −3. YB is given for different numbers of
parameters nxs and two different correlation values. For = 0.0, (241) is
applied to compute YB , for k,l = 0.8, (237) is applied. The given correlation
value holds for all parameter pairs.
It can be seen that the yield YB that corresponds to a tolerance box depends
on the dimension of the parameter space and on the correlations between para-
meters. This dependence is more pronounced for smaller tolerance boxes than
the one with βW = 3 in Table 10.
From Figure 22 follows that a tolerance box represents the worst-case if
correlations are unknown. Then, a tolerance box and corresponding yield value
according to (237) seems to be appropriate. Usually, at least a qualitative
knowledge about parameters being rather uncorrelated or strongly correlated is
available. In that case, we are interested to use a tolerance class that does not
depend on the number of parameters and does not depend on the correlation
values. As the parameters or transformed parameters can be assumed to be
normally distributed, it is promising to apply the tolerance region that results
from the level contours of the probability density function, i.e. ellipsoids.
Figure 48. Ellipsoid tolerance region Ts,E and equidensity contours of normal probability
density function pdfN .
the quadratic form (24). Defining a tolerance class according to such an ellip-
soid, the ellipsoid tolerance region Ts,E is:
Figure 48 illustrates the ellipsoid tolerance region Ts,E for two parameters.
Ellipsoid tolerance regions are also the given tolerance regions in the realistic
worst-case analysis described in Section 5.2 and the general worst-case analysis
describe in Section 5.4.
In order to get to the corresponding yield YE , we apply the variable trans-
formation (58),
Table 11. Yield values YE for an ellipsoid tolerance region Ts,E with βW = 3 in dependence
of the number of parameters nxs according to (247) and (248).
nxs YE
2 98.9%
3 97.1%
4 93.9%
5 89.1%
6 82.6%
7 74.7%
8 65.8%
9 56.3%
10 46.8%
.. ..
. .
15 12.3%
of freedom:
nxs
β2 = t2k ∼ χ2nxs (246)
k=1
From (247) and (248) we can see that the yield YE that corresponds to an
ellipsoid tolerance region is independent of the correlation values of the nor-
mal distribution of the parameters, contrary to a box tolerance region. But
(247) and (248) also show that YE depends on the dimension of the parameter
space, nxs .
Table 11 shows the yield value YE corresponding to a tolerance box Ts,E with
βW = 3 for different number of parameters. Obviously the yield YE within
a tolerance ellipsoid determined by βW strongly decreases with an increasing
number of parameters. This property is advantageous for the sampling proper-
ties of the normal distribution, as it adapts the spreading of sample elements to
Yield Analysis 121
the dimension of the scanned parameter space. But this property is disadvan-
tageous for a tolerance class, where it is unwanted if a tolerance region refers
to different yield values depending on the number of parameters.
Figure 49. Single-plane-bounded tolerance region Ts,SP,U and equidensity contours of normal
probability density function pdfN .
σf2 = gT · C · g (253)
In Figure 49, xs,W U being an upper bound means that the gradient g points
from the border of Ts,SP,U away from the tolerance region Ts,SP,U .
We can formulate an equivalent form of the single-plane-bounded tolerance
region Ts,SP,U and corresponding yield partition YSP,U to (249) and (250) in
the space of the single performance feature f :
pdff is a normal probability density function with mean value and covariance
as given in (252) and (253).
Let us express the difference between the given bound fU and the perfor-
mance value at xs,0 , f0 = f (xs,0 ), as a multiple of the performance variance
σf2 (253):
+ βW U · σf , f0 ≤ fU
fU − f0 = g · (xs,W U − xs,0 ) ≡
T
(257)
− βW U · σf , f0 ≥ fU
(257) considers that the nominal parameter vector xs,0 can be inside or outside
of the tolerance region Ts,SP,U according to (249). If xs,0 is outside of Ts,SP,U ,
then f0 > fU and the difference is negative.
(256) can be reformulated for a standardized normal distribution with zero
mean and unity variance using (257):
⎧ βW U
⎪
⎪ 1 − 12 t2
⎨ −∞ √2π e
⎪ · dt , f0 ≤ fU
YSP,U = (258)
⎪ − βW U 1
⎪
⎪
⎩ − 1 2
√ e 2 · dt , f0 ≥ fU
t
−∞ 2π
The equivalence of (249) and (255) means that a single-plane-bounded tolerance
region Ts,SP,U corresponds to the acceptance region of a single performance-
specification feature, i.e. a single bound on a performance feature.
The equivalence of (250) and (258) means that the corresponding yield parti-
tion YSP,U can be evaluated based on the standardized normal distribution with
zero mean and unity variance.
The same considerations can be done for a lower bound on a performance
feature, starting from the corresponding formulation of the tolerance region,
Ts,SP,L = {xs | gT · (xs − xs,W L ) ≥ 0}. The resulting yield partition is:
⎧ βW L
⎪
⎪ 1 1 2
√ e− 2 t · dt , f0 ≥ fL
⎪
⎨ −∞ 2π
YSP,L = − βW L (259)
⎪
⎪ 1
⎪
⎩ − 1 2
√ e 2 · dt , f0 ≤ fL
t
−∞ 2π
From (258) and (259) follows that the worst-case distance gets a positive sign,
if the performance-specification feature is satisfied at the nominal parameter
vector, and gets a negative sign, if the performance-specification feature is
violated at the nominal parameter vector. Figure 50 illustrates the four cases
124 ANALOG DESIGN CENTERING AND SIZING
Figure 50. Single-plane-bounded tolerance regions Ts,SP,L , Ts,SP,U for a single performance
feature with either an upper bound (first row) or a lower bound (second row), which is either
satisfied (first column) or violated at the nominal parameter vector (second column).
Table 12. Single-plane-bounded tolerance region Ts,SP and corresponding yield partition val-
ues YSP according to (258) or (259).
Ts,SP = [ −∞ , βW σf ] YSP
[ −∞ , −1σf ] 15.9%
[ −∞ , 0 ] 50.0%
[ −∞ , 1σf ] 84.1%
[ −∞ , 2σf ] 97.7%
[ −∞ , 3σf ] 99.9%
Figure 51. Single-plane-bounded tolerance region Ts,SP with corresponding worst-case para-
meter vector xW,SP and worst-case distance βW for two parameters. βW denotes a ±βW times
the covariances tolerance ellipsoid. Tolerance box Ts,B determined by ±βW times the covari-
ances with corresponding worst-case parameter vector xW,B and worst-case distance βW,B .
Table 13. Exaggerated robustness βW,B represented by a corner worst-case parameter vector
of a classical worst-case analysis, for different correlation ∀k=l k,l = among the parameters,
and for different numbers of parameters nxs .
Figure 52. Geometric yield analysis for a lower bound on a performance feature, f > fL ,
which is satisfied at the nominal parameter vector. Worst-case parameter vector xs,W L has the
smallest distance from the nominal statistical parameter vector xs,0 measured according to the
equidensity contours among all parameter vectors outside of or on the border of the acceptance
region partition As,L . The tangential plane to the tolerance ellipsoid through xs,W L as well as
to the border of As,L at xs,W L determines a single-plane-bounded tolerance region As,L .
Generally, (264) and (265) may have several solutions. In practice, a unique
solution as illustrated in Figure 52 can be observed for most performance fea-
tures. Exceptions are for instance mismatch-sensitive performance features for
which (264) and (265) often lead to two worst-case parameter vectors for each
mismatch-producing transistor pair, which are nearly symmetrical on both sides
of the nominal parameter vector. This special case requires special solution
algorithms.
Due to the presence of range-parameter tolerances, the constraints in (264)
and (265) have to be worked out in more detail. According to (123) and (124),
130 ANALOG DESIGN CENTERING AND SIZING
(266)–(269) are inserted into (264) and (265) to produce the problem formu-
lation of geometric yield analysis for the four cases of a lower/upper bound
that is satisfied/violated at the nominal statistical parameter vector. At the same
time, we replace the maximization of the probability density function by the
equivalent problem of minimizing β, which determines the equidensity contour
according to (24).
f ≥ fL and xs,0 ∈ As,L : min β 2 (xs ) s.t. min f (xs , xr ) ≤ fL (270)
xs ,xr xr ∈Tr
As mentioned, we have assumed box constraints for the range parameters (8).
(275) applies in (270) and (271), (276) applies in (272) and (273). The negative
132 ANALOG DESIGN CENTERING AND SIZING
∇LI,L (xr,W L )
λW L · (∇f (xr,W L ) − λ1,W L + λ2,W L ) = 0 (284)
k = 1, . . . , nxr : λ1,W L,k · (xr,W L,k − xr,L,k ) = 0 (285)
k = 1, . . . , nxr : λ2,W L,k · (xr,U,k − xr,W L,k ) = 0 (286)
xr,L ≤ xr,W L ≤ xr,U (287)
Yield Analysis 133
(281) results from the condition that ∇L(xs ) = 0 must hold for a stationary
point of the outer optimization problem. (284) results from the condition that
∇L(xr ) = 0 must hold for a stationary point of the inner optimization problem.
(282) represents the complementarity condition of the outer optimization
problem, (285) and (286) represent the complementarity condition of the inner
optimization problem.
(283) is the constraint of the outer optimization problem, (287) is the con-
straint of the inner optimization
problem.
∂f ∂f
∇f (xs,W L ) = ∂xs and ∇f (xr,W L ) = ∂x hold.
xs,W L ,xr,W L r xs,W L ,xr,W L
Note that (281) and (282) are very similar to the first-oder optimality condi-
tion of the general worst-case analysis (190) and (191). And that (284)–(286)
have nearly the same form as the first-oder optimality condition of the classical
worst-case analysis (153) and (155).
We can verify by contradiction that the constraint (283) is active at the
solution, i.e. f (xs,W L , xr,W L ) = fL . If this constraint would be inactive
at the solution, then λW L = 0. Using (281), then xs,W L = xs,0 would be
the solution. It follows that f (xs,W L , xr,W L ) = f (xs,0 , xr,W L ) < fL would
hold, which is a contradiction to the initial situation that the nominal statistical
parameter vector is satisfying the performance-feature bound. Therefore, the
constraint (283) is active at the solution in all cases (277)–(280), i.e.:
2
Nonnegative curvature of ∂∂xL2 at the worst-case parameter vector for uncon-
r
strained directions describes that the performance function is bounded below
with respect to the range parameters. This is required for the existence of a
border of the acceptance region As as defined in (123).
134 ANALOG DESIGN CENTERING AND SIZING
Figure 53. Two stationary points xs,A and xs,W of (277), which satisfy the first-order opti-
mality condition. Only xs,W satisfies the second-order optimality condition and is therefore a
solution of (270).
2
Nonnegative curvature of ∂∂xL2 at the worst-case parameter vector for
s
unconstrained directions corresponds to the relation between the curvature of
the tolerance ellipsoid of the statistical distribution and the curvature of the
border of the acceptance region. For xs,W L to be a minimum of (270), the cur-
vature of the tolerance ellipsoid has to be stronger than the curvature of the
border of the acceptance region as in the example in Figure 52.
Figure 53 shows another example. Both parameter vectors xs,A and xs,W L
satisfy the first-order optimality condition, but only xs,W L satisfies the second-
order optimality condition as well. In xs,A , the curvature of the border of the
acceptance region is stronger than the curvature of the corresponding tolerance
region. Therefore, this tolerance region is not a subset of the acceptance region,
and there are points inside this tolerance region but outside of the acceptance
region for which the probability density value or the corresponding β value is
larger than at xs,A . Therefore, xs,A is not a solution of (270).
Note that there is a third stationary point of (277) in Figure 53, where the
thin circle touches the border of the acceptance region. This point satisfies the
second-order optimality condition and is a local minimum of (270). A similar
situation will occur for locally varying mismatch-producing parameters. From
a deterministic optimizer starting from xs,0 can be expected that it will find
the global minimum xs,W L , as xs,W L is closer to xs,0 due to the problem
formulation. But as there is no guaranty for that and as we might be interested
to know all local minima, suitable measures have to be taken.
Yield Analysis 135
λ2W L
· ∇f (xs,W L )T · C · ∇f (xs,W L ) = βW
2
L
4
λW L βW L
=- (294)
2 ∇f (xs,W L )T · C · ∇f (xs,W L )
Inserting (294) in (293) yields the formulation of the statistical worst-case para-
meter vector for the case of a lower performance-feature bound that is satisfied
136 ANALOG DESIGN CENTERING AND SIZING
at the nominal parameter vector. The other cases are obtained analogously,
starting from the Lagrangian functions (277)–(280).
Note that (295), (296), which describe the worst-case statistical parameter vec-
tors of a geometric yield analysis, are identical to (194), (196), which describe
the worst-case statistical parameter vectors of a general worst-case analysis.
f¯(W L/U ) (xs ) = fL/U + ∇f (xs,W L/U )T · (xs − xs,W L/U ) (298)
−∇f (xs,W L/U )T · (xs,W L/U − xs,0 ) = f¯(W L/U ) (xs,0 ) − fL/U (299)
Note that (300) and (301), which describe the worst-case distance from a
geometric yield analysis, are identical to the worst-case distance from a single-
plane-bounded tolerance region (257), and are identical to the worst-case
distance from a general worst-case analysis (200)–(203).
In all cases, a worst-case distance is defined as a multiple of a performance
standard deviation σf¯(W L) . Specifically, it is the standard deviation of the
linearized performance at the worst-case parameter vector.
Moreover, (300) and (301) show that a change in the worst-case distance
consists of two parts. On the one hand, the distance between the performance-
feature bound and the nominal performance value has to be changed according
to the nominator in (300) and (301). This corresponds to performance centering
as described in Sections 2.4 and 2.5. On the other hand, the performance
sensitivity with regard to the statistical parameters has to be changed according
to the denominator in (300) and (301). The appropriate combination of both
parts constitutes yield optimization/design centering. Note that we are aiming
at increasing the worst-case distance if the nominal parameter vector is inside
the parameter acceptance region partition, and that we are aiming at decreasing
the worst-case distance if the nominal parameter vector is outside the parameter
acceptance region partition.
Figure 54. Worst-case distances and approximate yield values from a geometric yield analysis
of the operational amplifier from Table 14.
140 ANALOG DESIGN CENTERING AND SIZING
Figure 55. Parameter acceptance region As (gray area) originating from four performance-
specification features, f1 ≥ fL,1 , f1 ≤ fU,1 , f2 ≥ fL,2 , f2 ≤ fU,2 (Figure 32). A geometric
yield analysis leads to four worst-case parameter vectors xW L,1 , xW U,1 , xW L,2 , xW U,2 and
four single-plane-bounded tolerance regions. The intersection of these single-plane-bounded
tolerance regions forms the approximate parameter acceptance region As (linen-pattern-filled
area).
Figure 56. General worst-case analysis and geometric yield analysis as inverse mappings
exchanging input and output.
6.4 Exercise
Given is the example of Section 5.6 with a single performance function of
two parameters:
f = xs,1 · xs,2 (305)
f could be for instance the time constant of the RC circuit in Chapter 2. The
nominal parameter vector is:
xs,0 = [ 1 1 ]T (306)
The parameters are normally distributed with the covariance matrix:
0.22 0
C= (307)
0 0.82
Two performance-specification features are given:
f ≥ fL ≡ 0.5 (308)
f ≤ fU ≡ 2.0 (309)
Perform a geometric yield analysis for the two performance-specification
features. Apply the optimality conditions (Appendix C) to calculate a
solution.
Given is the following single performance function of two parameters:
1
f = x2s,1 · x2s,2 (310)
4
The nominal parameter vector is:
xs,0 = [ 0 0 ]T (311)
The parameters are normally distributed with the covariance matrix:
1 0
C= (312)
0 1
One performance-specification feature is given:
f ≤ fU ≡ 1.0 (313)
Perform a geometric yield analysis. Apply the optimality conditions
(Appendix C) to calculate a solution. Check the positive definiteness of
∇2 L(xr,W L ) and the second-order optimality condition (290) to verify the
solution.
Chapter 7
1
pdfδ (xs ) = · δ(xs ) · pdf(xs ) (314)
Y
146 ANALOG DESIGN CENTERING AND SIZING
∞ ∞
1
xs,0,δ = E {xs } = · ··· xs · δ(xs ) · pdf(xs ) · dxs (315)
pdf Y −∞ −∞
δ
1
nM C
/s,0,δ =
x s ) · xs
δ(x(µ) (µ)
(317)
nok
µ=1
n
MC
/δ = 1
C s )(xs −/
δ(x(µ) (µ)
xs,0,δ )(x(µ)
s −/xs,0,δ )T (318)
nok − 1
µ=1
n
MC
/
nok = s ) = Y · nM C ≈ Y · nM C
δ(x(µ) (319)
µ=1
xs,0,δ denotes the mean value of the truncated probability density function pdfδ
according to (315).
From (324), the first-order optimality condition for a yield maximum Y ∗ =
Y (x∗s,0 ) follows immediately:
(325) says that the optimal yield is achieved when the mean value of the trun-
cated probability density function equals that of the original probability density
function.
The mean value of a probability density function can be interpreted as the
center of gravity of the mass represented by the volume under the probability
density function. The first-order optimality condition (325) can therefore be
interpreted in the sense, that the truncation of the probability density function
due to the performance specification does not change the center of gravity
in the optimum. This is the motivation for the term design centering. In the
optimum nominal statistical parameter vector x∗s,0,δ , the design is centered with
regard to the performance specification in the sense that the center of gravity
of the probability density function is not affected by the truncations due to
the performance specification. Design centering means to find a sizing which
represents an equilibrium concerning the center of gravity of the manufactured
and tested “mass.”
Note that a centering of the performance-feature values between their bounds
is not a centered design according to this interpretation. Note also that geomet-
rically inscribing a maximum tolerance ellipsoid in the parameter acceptance
region neither is a centered design according to this interpretation.
Figure 57 illustrates the situation before and after having reached the equi-
librium concerning the centers of gravity of original and truncated probability
density function for two statistical parameters. Those parts of the equidensity
contours that belong to the truncated probability density function are drawn
in bold.
148 ANALOG DESIGN CENTERING AND SIZING
Figure 57. (a) Statistical-yield optimization before having reached the optimum. Center of
gravity xs,0,δ of the probability density function truncated due to the performance specification
(remaining parts drawn as bold line) differs from the center of gravity of the original probability
density function. A nonzero yield gradient ∇Y (xs,0 ) results. (b) After statistical-yield opti-
mization having reached the optimum. Centers of gravity of original and truncated probability
density function are identical.
We insert (328) into (327) and extend the terms xs − xs,0 to (xs − xs,0,δ ) +
(xs,0,δ − xs,0 ):
∇2 Y (xs,0 ) =
+∞
C−1 · · · δ(xs )[(xs −xs,0,δ )+(xs,0,δ −xs,0 )][”]T pdfN (xs )dxs − Y C C−1 (329)
−∞
Using (315), (316) and (A.12)–(A.14), we obtain from (329) the Hessian matrix
of the yield with regard to the nominal values of statistical parameters:
∇2 Y (xs,0 ) = Y C−1 Cδ + (xs,0,δ − xs,0 )(xs,0,δ − xs,0 )T − C C−1 (330)
xs,0,δ denotes the mean value of the truncated probability density function pdfδ
according to (315), and Cδ denotes the covariance matrix of the truncated
probability density function pdfδ according to (316).
xs,0,δ and Cδ can be estimated as part of a Monte-Carlo analysis using (317)
and (318).
From (330) and the first-order optimality condition (325), the necessary
second-order optimality condition for a yield maximum Y ∗ = Y (x∗s,0 ) follows
immediately:
Cδ − C is negative semidefinite (331)
(331) says that the yield is maximum if the variability of the truncated prob-
ability density function is smaller than that of the original probability density
function. This expresses the property that the performance specification cuts
away a part of the volume under the probability density function, and does this
still in the optimum.
Computing the statistical yield gradient with regard to deterministic design para-
meters according to (337) requires nxd Monte-Carlo analyses. The resulting
computational costs are prohibitive if numerical simulation is applied.
An alternative can be developed based on selecting a single statistical
parameter xs,k ,
⎡ ⎤
xs,1 ⎡ ⎤
xs,1
⎢ .. ⎥
⎢ . ⎥ ⎢ .. ⎥
⎢ ⎥ ⎢ . ⎥
⎢ xs,k−1 ⎥ ⎢ ⎥
⎢ ⎥ ⎢ x ⎥
xs = ⎢ xs,k ⎥ ∈ Rnxs −→ xs = ⎢ s,k−1 ⎥ ∈ Rnxs −1 , xs,k (338)
⎢ ⎥ ⎢ xs,k+1 ⎥
⎢ xs,k+1 ⎥ ⎢ .. ⎥
⎢ .. ⎥ ⎣ ⎦
⎣ . ⎦ .
xs,nxs
xs,nxs
and formulating the yield (116) via the marginal distribution of the remaining
statistical parameter xs :
+∞ +∞ 0 xs,k,U (xs ) 1
Y = ... pdfN (xs,k ) · dxs,k pdfN (xs ) · dxs (339)
−∞ −∞ xs,k,L (xs )
= E cdfN (xs,k,U (xs )) − cdfN (xs,k,L (xs )) (340)
pdfN (xs )
Here, we have assumed that the parameters have been transformed into stan-
dardized normally distributed random variables xs with zero mean and unity
covariance matrix, i.e. xs ∼ N (0, I). pdfN and cdfN are the probability
density function and cumulative distribution function according to (16), (17)
and (23).
xs,k,U and xs,k,L represent the borders of the parameter acceptance region
As (122) projected onto the axis of parameter xs,k
A statistical yield estimator based on (340) is determined using (B.1):
n
MC
1
Y/ = (cdfN (xs,k,U (xs(µ) )) − cdfN (xs,k,L (xs(µ) ))) (341)
nM C
µ=1
1 1
nxs nM C
/
Y/ = (cdfN (xs,k,U (xs(µ) )) − cdfN (xs,k,L (xs(µ) ))) (342)
nxs nM C
k=1 µ=1
(341) and (342) are evaluated based on a Monte-Carlo analysis. This requires
(µ) (µ) (µ)
the computation of xs,k,U (xs ) and xs,k,L (xs ) in each sample element xs
by solving the following optimization problems, if the nominal statistical para-
meter vector is inside the acceptance region, xs,0 ∈ AsL/U :
⎧
⎪ (µ)
⎪
⎪ min fi (xs , xs,k , xr ) ≤ fL,i
⎪
⎪ x ∈ T
⎪
⎨
r r
(µ)
min (xs,k − xs,k,0 ) s.t.
2 max fi (xs , xs,k , xr ) ≥ fU,i (343)
⎪ x ∈ T
⎪
xs,k ,xr r r
⎪
⎪ i = 1, . . . , nf
⎪
⎪
⎩ x ≤x
s,k s,k,0
⎧
⎪ (µ)
⎪
⎪ min fi (xs , xs,k , xr ) ≤ fL,i
⎪
⎪ x ∈ T
⎪
⎨
r r
(µ)
min (xs,k − xs,k,0 ) s.t.
2 max fi (xs , xs,k , xr ) ≥ fU,i (344)
⎪ x ∈ T
⎪
xs,k ,xr r r
⎪
⎪ i = 1, . . . , nf
⎪
⎪
⎩ x ≥x
s,k s,k,0
(343) and (344) compare to the geometric yield analysis (264). The difference
is that all performance-specification features are considered simultaneously and
that only one parameter is considered in the objective function.
The solution of (343) and (344) becomes a line search along the parameter
xs,k if the worst-case range-parameter vector can be predetermined as described
in Section 6.3.5.
The statistical yield gradient with regard to deterministic design parameters
is formulated starting from (340):
Here fi and fj are the performance features whose bounds are active in the
solution of (343) and (344).
In addition to the nM C simulations of a Monte-Carlo analysis, the com-
putation of the statistical yield gradient with respect to deterministic design
Yield Optimization/Design Centering 153
parameters requires at least 2nM C line searches to solve (343) and (344) plus
2nM C sensitivity analyses to solve (346).
The resulting simulation costs may still be prohibitive in practice. Methods
based on statistical estimation of the yield gradient for deterministic design
parameters like [43, 63] therefore fall back on response surface models.
xd,µ :
From (349) and (350), the worst-case-distance gradients with respect to the
mean values of statistical parameters xs,0 and with respect to (deterministic)
design parameters xd follow:
+1
∇βW L/U (xd )= . ·∇f (xd,µ ) (352)
∇f (xs,W L/U )T ·C· ∇f (xs,W L/U )
−1
∇βW L/U (xd )= . ·∇f (xd,µ ) (354)
∇f (xs,W L/U )T ·C· ∇f (xs,W L/U )
Yield Optimization/Design Centering 155
(351)–(354) show that the worst-case-distance gradient has the same form con-
cerning statistical parameters and (deterministic) design parameters.
The worst-case-distance gradient corresponds to the performance gradient at
the worst-case parameter vector. Its length is scaled according to the variance
of the linearized performance (198), its direction depends on whether a lower
or upper performance-feature bound is specified and whether it is satisfied or
violated.
The equivalence of statistical parameters and deterministic design parameters
in the worst-case distance (349) and (350) and in the worst-case-distance gra-
dient (351)–(354) can also be interpreted using Figure 52.
The worst-case distance is increased by increasing the difference between the
performance-feature value at the nominal statistical parameter vector and the
performance-feature bound. In a linear performance model, this increase can
be achieved by shifting any parameter of any kind. In Figure 52, which shows
the subspace of parameters that are statistical, the performance gradient at the
worst-case parameter vector shows the direction in which the nominal statistical
parameter vector xs,0 has to be shifted for a steepest ascent in the worst-case
distance according to (351). The same effect is achieved by a change in the
(deterministic) design parameter vector xd according to (352), which shifts the
boundary of As,L away from xs,0 according to (348).
∇βW
2
L/U () = λW L/U (355)
With (294) and based on differentiating (356) and (357) with respect to xd , we
obtain (352) and (354).
β W L/U = [ . . . αi · βW L/U,i . . . ]T
158 ANALOG DESIGN CENTERING AND SIZING
Figure 58. Worst-case distances and approximate yield partition values before and after
geometric-yield optimization of the operational amplifier from Table 15.
which has a quadratic objective function due to the last term in (363). The
stationary point of (364) is calculated as:
a level contour of the objective ¯F (r)2 . Figure 59 illustrates optimum steps
for some trust regions.
Due to the quadratic nature of the objective ¯F (r)2 , the amount of
additional decrease in the objective that can be obtained by an increase of
the allowed step length r ≤ ∆ is decreasing. This is even more so with a
worsening problem condition of the Jacobian matrix J.
Therefore, the Pareto front of objectives ¯F (r∗ )2 versus the step r∗
according to (366) acquires a typical shape as illustrated in Figure 60.
A point in the bend of this curve is preferable, because it leads to a grand
progress towards the target worst-case distances at a small step length. The
additional progress towards the target worst-case distances beyond the bend
is rather small. Additionally, the required step length for additional progress
beyond the bend becomes large, and the step will be more likely to leave the
region of validity of the linearized model (362). Therefore, a step in the bend
will be selected. Additional simulations have to be spent to verify that the
linearized model holds in the corresponding trust region.
Note that the described approach is a part of a complete optimization algo-
rithm. Other algorithmic components are required that deal with updating the
target values of the optimization objectives, or with determining bends in the
Pareto curve.
162 ANALOG DESIGN CENTERING AND SIZING
A.2 Moments
The moment of order κ, m(κ) , of a single random variable z results from
setting h(z) = z κ in (A.1):
m(κ) = E {z κ } (A.2)
A.5 Variance
The variance of a single random variable z is defined as the central moment
of order 2 of z and denoted by σ 2 or V {z}:
c(2) = σ 2 = V {z} = E (z − µ)2 (A.6)
σ denotes the standard deviation of a single random variable z:
-
σ = V {z} (A.7)
A.6 Covariance
The covariance, cov {zk , zl }, of two random variables zk and zl is defined
as a mixed central moment of order 2:
cov {zk , zl } = E {(zk − mk ) · (zl − ml )} (A.8)
A.7 Correlation
The correlation, corr {zk , zl }, of two random variables zk and zl is defined as
their covariance normalized with respect to their individual standard deviations:
cov {zk , zl }
corr {zk , zl } = (A.9)
σk · σl
2
V {h(z)} = E (h(z) − a) · (h(z) − a)T
− (E {h(z)} − a) · (E {h(z)} − a)T (A.14)
2
V {h(z)} = E (h(z) − a)2 − (E {h(z)} − a)2
V {h(z)} = E h(z) · hT (z) − E {h(z)} · E hT (z)
V {h(z)} = E h2 (z) − (E {h(z)})2
(A.12) is the linear transformation formula for expectation values. It says that
the expectation value of the linear transformation of a random vector equals the
corresponding linear transformation of the vector’s expectation value.
(A.13) is the linear transformation formula for variances.
(A.14) is the translation formula for variances. It relates the variance as a
second-order central moment to the second-order moment and the quadratic
expectation value.
From (A.13) follows the Gaussian error propagation:
V{aT · z + b} = aT · V {z} · a = aT · C · a
k,l =0 2 2
= ak al σk k,l σl = ak σk (A.15)
k,l k
168 ANALOG DESIGN CENTERING AND SIZING
A.11 Exercises
1. Prove (A.12). (Apply (A.1), (15).)
2. Prove (A.13). (Apply (A.11), (A.12), (A.1).)
3. Prove (A.14). (Apply (A.11), (A.12).)
4. Prove that z according to (A.16) has a mean value of zero and a variance
of one. (Apply calculation formulas from Section A.9.)
5. Show that (52) holds for a multinormal distribution according to (23) and
(24). (Insert (23) and (24) in (A.1). Use (57) and (58) for a variable sub-
stitution. Consider that the corresponding integral over an even function is
zero.)
6. Show that (53) holds for a multinormal distribution according to (23) and
(24). (Insert (23) and (24) in (A.10). Use (57) and (58) for a variable
substitution.)
Appendix B
Statistical Estimation of Expectation Values
x(µ) ∼ D (pdf(x)) , µ = 1, . . . , nM C
A statistical estimation is based on a sample of sample elements.
A widehat is used to denote an estimator function Φ(x/ (1) , . . . , x(nM C ) ) for
a function Φ(x).
Sample elements x(µ) are independently and identically distributed accord-
ing to the given statistical distribution D with the probability density function
pdf(x). Therefore:
2
E h(x(µ) ) = E {h(x} = mh (B.2)
2
V [h(x(1) ) . . . h(x(nM C ) )]T = diag V{h(x(1) )} . . .
V{h(x(nM C ) )} (B.3)
2
V h(x(µ) ) = V {h(x)} (B.4)
x(µ) ∼ D (pdf(x)) , µ = 1, . . . , nM C
170 ANALOG DESIGN CENTERING AND SIZING
From (B.10) follows that the quality measure Q is the estimator variance, i.e.
the variance of the estimation function, for an unbiased estimator:
2
bΦ
= 0 : QΦ
= V Φ / (B.11)
with h(µ) = h(x(µ) ). Applying (A.13), (B.3) and (B.4) finally leads to:
2 1
/
Qµ
h = V E {h(x)} = · V {h(x)} (B.14)
nM C
Using (B.16) and (B.17) in the above proof leads to the corresponding formula
for an estimator of the quality:
2
/ µ
= V
Q / E / {h(x)} = 1 · V / {h(x)} (B.15)
h
nM C
/ {A · h(z) + b} = A · E
E / {h(z)} + b (B.16)
/ {A · h(z) + b} = A · V
V / {h(z)} · AT (B.17)
/ {h(z)} = nM C
V / h(z) · hT (z) − E
E / {h(z)} · E
/ hT (z) (B.18)
nM C − 1
B.7 Exercises
1. Prove that the expectation-value estimator according to (B.1) is unbiased.
(Check if (B.8) is true. Apply (A.12) and (B.2).)
2. Prove the second part of (B.10). (Express Φ by using (B.7) and insert in
(A.14). Combine terms with Φ / as h.)
Starting point is the Taylor expansion of the objective function f (x) at the
optimum solution x∗ and f ∗ = f (x∗ ) :
1 T
f (x∗ + r) = f ∗ + gT · r + · r · H · r + ... (C.3)
2
174 ANALOG DESIGN CENTERING AND SIZING
Figure C1. Descent directions and steepest descent direction of a function of two parameters.
∀ gT · r ≥ 0 (C.6)
r = 0
Appendix C: Optimality Conditions of Nonlinear Optimization Problems 175
Necessary Condition. The only way to satisfy (C.6) is that the gradient g is
zero. That represents the necessary first-order optimality condition for a local
minimum solution x∗ with f ∗ = f (x∗ ):
∇f (x∗ ) = 0 (C.7)
x**2+y**2 -x**2-y**2
200 -20
180 -40
160 -60
140 -80
120 -100
100 -120
200 80 0 -140
180 60 -20 -160
160 40 -40 -180
140 20 -60
120 -80
100 -100
80 -120
60 -140
40 -160
20 -180
0 -200
10 10
5 5
-10 0 -10 0
-5 -5
0 -5 0 -5
5 5
10 -10 10 -10
x**2 x**2-y**2
100 100
90 80
80 60
70 40
60 20
50 0
100 40 100 -20
30 -40
80 20 -60
10 50 -80
60
0
40
20 -50
0 -100
10 10
5 5
-10 0 -10 0
-5 -5
0 -5 0 -5
5 5
10 -10 10 -10
Figure C2. Different types of definiteness of the second-derivative of a function of two para-
meters: positive definite (upper left), positive semidefinite (upper right), indefinite (lower left),
negative definite (lower right).
Figure C3. Descent directions and unconstrained directions of a function of two parameters
with one active constraint.
C3, this corresponds to the region where the two gray block sectors overlap and
is illustrated with a dark gray block sector.
Obviously, each active constraint reduces the range of unconstrained descent
directions.
case. Then, we would formulate the requirement that the first-order derivative
of the gradient of the Lagrangian function is zero:
∇L(x) ≡ 0 : ∇f (x∗ ) − λ∗ · ∇c(x∗ ) = 0 (C.16)
cµ (x∗ ) = 0, µ ∈ EC (C.18)
cµ (x∗ ) ≥ 0, µ ∈ IC (C.19)
λ∗µ ≥ 0, µ ∈ IC (C.20)
λ∗µ · cµ (x∗ ) = 0, µ ∈ EC ∪ IC (C.21)
(C.17) and (C.20) are explained through the condition that no unconstrained
descent direction exists in the minimum. The restrictions in (C.17) are defined
for all active constraints A(x∗ ).
No statement about the sign of the Lagrange factor can be made for an
equality constraint.
(C.18) and (C.19) formulate that the constraint must not be violated in the
minimum.
(C.21) is the so-called complementarity condition. It expresses that in the
minimum either a constraint is 0 (that means “active”) or the corresponding
Lagrange factor is 0. If both are zero at the same time, this corresponds to a
minimum of the objective function where the constraint just got active. Deleting
this constraint would not change the solution of the optimization problem.
A Lagrange factor of zero for an inactive constraint corresponds to deleting
it from the Lagrange function. Therefore we have that:
L(x∗ , λ∗ ) = f (x∗ ) (C.22)
(C.17)–(C.21) are known as Karush-Kuhn-Tucker(KKT) conditions.
1
= L(x∗ , λ∗ ) + ∇L(x∗ )T · r + rT ·∇2 L(x∗ )·r + . . . (C.23)
2
0 (C.17)
∇2 L(x∗ ) = ∇2 f (x∗ ) − λ∗µ · ∇2 cµ (x∗ ) (C.24)
µ∈A(x∗ )
∗ T
∀ ∗
rT · ∇2 L(x∗ ) · r ≥ 0 (C.25)
∇cµ (x ) · r = 0 , µ ∈ A+ (x )
∇cµ (x∗ )T · r ≥ 0, µ ∈ A(x∗ ∗
) \ A+ (x )
A+ (x∗ ) = EC ∪ µ ∈ IC cµ (x∗ ) = 0 ∧ λ∗µ > 0
Note that this is a weaker requirement than positive definiteness, because not
all directions r are included in (C.25).
(C.22):
(C.27)
∇f ∗ (µ ) = ∇L∗ (µ ) = λ∗µ (C.28)
(C.28) says that the sensitivity of the minimum objective with respect to an
increase in the constraint boundary is equal to the corresponding Lagrange
factor.
C−1 = C / ⇔C=C
/T · C / −1 · C
/ −T , (C.33)
which can be obtained by a Cholesky decomposition or an eigenvalue decom-
position, (C.31) can be transformed into:
/ · x∗ = 1 · C
C / −T · ej (C.34)
λ∗
Applying (C.33) and two times (C.34) in (C.32) leads to
σk = |λ∗ | · β (C.35)
Applying (C.33) and one time (C.34) in (C.32) leads to
x∗k = λ∗ · β 2 (C.36)
182 ANALOG DESIGN CENTERING AND SIZING
From (C.35) and (C.36) and the Hessian matrix of the Lagrangian function
(C.30), ∇2 L(x) = −λ · C−1 , follows that the Lagrangian function (C.30) has
a maximum and a minimum with the absolute value |x∗k | = β · σj .
(C.29) therefore says that any ellipsoid with any correlation value leads to a
minimum and maximum value of x∗k = ±β · σk . This shows the lub property
of (37). For a complete proof we additionally have to show that in direction
of other parameters all values on the bounding box are reached by varying the
correlation. Instead we refer to the visual inspection of Figure 22(d).
References
[1] H. Abdel-Malek and A. Hassan. The ellipsoidal technique for design centering and
region approximation. IEEE Transactions on Computer-Aided Design of Circuits and
Systems, 10:1006–1013, 1991.
[2] D. Agnew. Improved minimax optimization for circuit design. IEEE Transactions on
Circuits and Systems CAS, 28:791–803, 1981.
[4] Antonio R. Alvarez, Behrooz L. Abdi, Dennis L. Young, Harrison D. Weed, Jim Teplik,
and Eric R. Herald. Application of statistical design and response surface methods to
computer-aided VLSI device design. IEEE Transactions on Computer-Aided Design of
Circuits and Systems, 7(2):272–288, February 1988.
[6] K. Antreich, H. Graeb, and C. Wieser. Circuit analysis and optimization driven by worst-
case distances. IEEE Transactions on Computer-Aided Design of Circuits and Systems,
13(1):57–71, January 1994.
[7] K. Antreich and S. Huss. An interactive optimization technique for the nominal design
of integrated circuits. IEEE Transactions on Circuits and Systems CAS, 31:203–212,
1984.
[8] K. Antreich and R. Koblitz. Design centering by yield prediction. IEEE Transactions
on Circuits and Systems CAS, 29:88–95, 1982.
[10] Kurt J. Antreich, Helmut E. Graeb, and Rudolf K. Koblitz. Advanced Yield Optimization
Techniques, Volume 8 (Statistical Approach to VLSI) of Advances in CAD for VLSI.
Elsevier Science Publishers, Amsterdam, 1994.
184 REFERENCES
[12] J. Bandler and S. Chen. Circuit optimization: The state of the art. IEEE Transactions
on Microwaves Theory Techniques (MTT), 36:424–442, 1988.
[13] J. Bandler, S. Chen, S. Daijavad, and K. Madsen. Efficient optimization with integrated
gradient approximation. IEEE Transactions on Microwaves Theory Techniques (MTT),
36:444–455, 1988.
[15] Kamel Benboudjema, Mounir Boukadoum, Gabriel Vasilescu, and Georges Alquie.
Symbolic analysis of linear microwave circuits by extension of the polynomial interpo-
lation method. IEEE Transactions on Circuits and Systems I: Fundamental Theory and
Applications, 45(9):936, 1998.
[16] M. Bernardo, R. Buck, L. Liu, W. Nazaret, J. Sacks, and W. Welch. Integrated circuit
design optimization using a sequential strategy. IEEE Transactions on Computer-Aided
Design of Circuits and Systems, 11:361–372, 1992.
[19] Stephen Boyd and Lieven Vandenberghe. Convex Optimization. Cambridge University
Press, 2004.
[20] Graeme R. Boyle, Barry M Cohn, Danald O. Pederson, and James E. Solomon. Macro-
modeling of integrated operational amplifiers. IEEE Journal of Solid-State Circuits SC,
9(6):353–364, December 1974.
[23] R. Chadha, K. Singhal, J. Vlach, and E. Christen. WATOPT - an optimizer for circuit
applications. IEEE Transactions on Computer-Aided Design of Circuits and Systems,
6:472–479, 1987.
[25] E. Christensen and J. Vlach. NETOPT – a program for multiobjective design of linear
networks. IEEE Transactions on Computer-Aided Design of Circuits and Systems,
7:567–577, 1988.
[26] M. Chu and D. J. Allstot. Elitist nondominated sorting genetic algorithm based rf ic
optimizer. IEEE Transactions on Circuits and Systems CAS, 52(3):535–545, March
2005.
[27] L. Chua. Global optimization: a naive approach. IEEE Transactions on Circuits and
Systems CAS, 37:966–969, 1990.
[28] Andrew R. Conn, Paula K. Coulman, Ruud A. Haring, Gregory L. Morill, Chandu
Visweswariah, and Chai Wah Wu. JiffyTune: Circuit optimization using time-domain
sensitivities. IEEE Transactions on Computer-Aided Design of Circuits and Systems,
17(12):1292–1309, December 1998.
[29] P. Cox, P. Yang, S. Mahant-Shetti, and P. Chatterjee. Statistical modeling for efficient
parametric yield estimation of MOS VLSI circuits. IEEE Transactions on Electron
Devices ED, 32:471–478, 1985.
[30] Walter Daems, Georges Gielen, and Willy Sansen. Circuit simplification for the symbolic
analysis of analog integrated circuits. IEEE Transactions on Computer-Aided Design of
Circuits and Systems, 21(4):395–407, April 2002.
[31] Walter Daems, Georges Gielen, and Willy Sansen. Simulation-based generation of
posynomial performance models for the sizing of analog integrated circuits. IEEE
Transactions on Computer-Aided Design of Circuits and Systems, 22(5):517–534, May
2003.
[32] Walter Daems, Wim Verhaegen, Piet Wambacq, Georges Gielen, and Willy Sansen.
Evaluation of error-control strategies for the linear symbolic analysis of analog inte-
grated circuits. IEEE Transactions on Circuits and Systems I: Fundamental Theory and
Applications, 46(5):594–606, May 1999.
[34] Indraneel Das and J. E. Dennis. Normal-boundary intersection: A new method for
generating the Pareto surface in nonlinear multicriteria optimization problems. SIAM
Journal on Optimization, 8(3):631–657, August 1998.
[35] Bart De Smedt and Georges G. E. Gielen. WATSON: Design space boundary exploration
and model generation for analog and RF IC design. IEEE Transactions on Computer-
Aided Design of Circuits and Systems, 22(2):213–223, February 2003.
[37] Maria del Mar Hershenson, Stephen P. Boyd, and Thomas H. Lee. Optimal design of a
CMOS Op-Amp via geometric programming. IEEE Transactions on Computer-Aided
Design of Circuits and Systems, 20(1):1–21, January 2001.
[38] S. Director and G. Hachtel. The simplicial approximation approach to design centering.
IEEE Transactions on Circuits and Systems CAS, 24:363–372, 1977.
186 REFERENCES
[39] Alex Doboli and Ranga Vemuri. Behavioral modeling for high-level synthesis of ana-
log and mixed-signal systems from vhdl-ams. IEEE Transactions on Computer-Aided
Design of Circuits and Systems, 2003.
[40] K. Doganis and D. Scharfetter. General optimization and extraction of IC device model
parameters. IEEE Transactions on Electron Devices ED, 30:1219–1228, 1983.
[41] Hans Eschenauer, Juhani Koski, and Andrzej Osyczka. Multicriteria design optimiza-
tion: procedures and applications. Springer-Verlag, 1990.
[42] Mounir Fares and Bozena Kaminska. FPAD: A fuzzy nonlinear programming approach
to analog circuit design. IEEE Transactions on Computer-Aided Design of Circuits and
Systems, 14(7):785–793, July 1995.
[43] P. Feldmann and S. Director. Integrated circuit quality optimization using surface inte-
grals. IEEE Transactions on Computer-Aided Design of Circuits and Systems, 12:1868–
1879, 1993.
[45] Roger Fletcher. Practical Methods of Optimization. John Wiley & Sons, 1987.
[46] Kenneth Francken and Georges G. E. Gielen. A high-level simulation and synthesis
environment for sigma delta modulators. IEEE Transactions on Computer-Aided Design
of Circuits and Systems, 22(8):1049–1061, August 2003.
[47] D.D. Gajski and R.H. Kuhn. Guest editor’s introduction: New VLSI tools. ieeecomputer,
16:11–14, 1983.
[48] Floyd W. Gembicki and Yacov Y. Haimes. Approach to performance and sensitivity mul-
tiobjective optimization: The goal attainment method. IEEE Transactions on Automatic
Control, 20(6):769–771, December 1975.
[50] G. Gielen and W. Sansen. Symbolic Analysis for Automated Design of Analog Integrated
Circuits. Kluwer Academic Publishers, Dordrecht, 1991.
[51] G. Gielen, P. Wacambacq, and W. Sansen. Symbolic analysis methods and applications
for analog circuits: A tutorial overview. Proceedings of the IEEE, 82, 1994.
[52] G. Gielen, H. C. Walscharts, and W. C. Sansen. ISAAC: A symbolic simulation for analog
integrated circuits. IEEE Journal of Solid-State Circuits SC, 24:1587–1597, December
1989.
[53] G. Gielen, H. C. Walscharts, and W. C. Sansen. Analog circuit design optimization based
on symbolic simulation and simulated annealing. IEEE Journal of Solid-State Circuits
SC, 25:707–713, June 1990.
References 187
[54] Georges G. E. Gielen, Kenneth Francken, Ewout Martens, and Martin Vogels. An analyt-
ical integration method for the simulation of continuous-time delta-sigma modulators.
IEEE Transactions on Computer-Aided Design of Circuits and Systems, 2004.
[55] Georges G. E. Gielen and Rob A. Rutenbar. Computer-aided design of analog and mixed-
signal integrated circuits. Proceedings of the IEEE, 88(12):1825–1852, December 2000.
[56] Philip E. Gill, Walter Murray, and Margaret H. Wright. Practical Optimization. Aca-
demic Press. Inc., London, 1981.
[58] H. Graeb, S. Zizala, J. Eckmueller, and K. Antreich. The sizing rules method for analog
integrated circuit design. In IEEE/ACM International Conference on Computer-Aided
Design (ICCAD), pages 343–349, 2001.
[59] A. Groch, L. Vidigal, and S. Director. A new global optimization method for electronic
circuit design. IEEE Transactions on Circuits and Systems CAS, 32:160–169, 1985.
[60] G. D. Hachtel and P. Zug. APLSTAP – circuit design and optimization system – user’s
guide. Technical report, IBM Yorktown Research Facility, Yorktown, New York, 1981.
[61] R. Hanson and C. Lawson. Solving Least Squares Problems. Prentice-Hall, New Jersey,
1974.
[62] R. Harjani, R. Rutenbar, and L. Carley. OASYS: A framework for analog circuit syn-
thesis. IEEE Transactions on Computer-Aided Design of Circuits and Systems, 8:1247–
1266, 1989.
[63] D. Hocevar, P. Cox, and P. Yang. Parametric yield optimization for MOS circuit blocks.
IEEE Transactions on Computer-Aided Design of Circuits and Systems, 7:645–658,
1988.
[65] Xiaoling Huang, Chris S. Gathercole, and H. Alan Mantooth. Modeling nonlinear
dynamics in analog circuits via root localization. IEEE Transactions on Computer-
Aided Design of Circuits and Systems, 2003.
[66] Ching-Lai Hwang and Abu Syed Md. Masud. Multiple Objective Decision Making.
Springer, 1979.
[67] Jacob Katzenelson and Aharon Unikovski. Symbolic-numeric circuit analysis or sym-
bolic ciruit analysis with online approximations. IEEE Transactions on Circuits and
Systems I: Fundamental Theory and Applications, 46(1):197–207, January 1999.
[69] G. Kjellstroem and L. Taxen. Stochastic optimization in system design. IEEE Transac-
tions on Circuits and Systems CAS, 28:702–715, 1981.
[70] Ken Kundert, Henry Chang, Dan Jefferies, Gilles Lamant, Enrico Malavasi, and Fred
Sendig. Design of mixed-signal systems-on-a-chip. IEEE Transactions on Computer-
Aided Design of Circuits and Systems, 19(12):1561–1571, December 2000.
[71] Francky Leyn, Georges Gielen, and Willy Sansen. Analog small-signal modeling – part
I: Behavioral signal path modeling for analog integrated circuits. IEEE Transactions
on Circuits and Systems II: Analog and Digital Signal Processing, 48(7):701–711, July
2001.
[72] Francky Leyn, Georges Gielen, and Willy Sansen. Analog small-signal modeling –
part II: Elementary transistor stages analyzed with behavioral signal path modeling.
IEEE Transactions on Circuits and Systems II: Analog and Digital Signal Processing,
48(7):712–721, July 2001.
[73] M. Lightner, T. Trick, and R. Zug. Circuit optimization and design. Circuit Analysis,
Simulation and Design, Part 2 (A. Ruehli). Advances in CAD for VLSI 3, pages 333–391,
1987.
[74] M. R. Lightner and S. W. Director. Multiple criterion optimization for the design of
electronic circuits. IEEE Transactions on Circuits and Systems CAS, 28(3):169–179,
March 1981.
[75] V. Litovski and M. Zwolinski. VLSI Circuit Simulation and Optimization. Chapman
Hall, 1997.
[76] Hongzhou Liu, Amit Singhee, Rob A. Rutenbar, and L. Richard Carley. Remebrance
of circuits past: Macromodeling by data mining in large analog design spaces. In
ACM/IEEE Design Automation Conference (DAC), pages 437–442, 2002.
[77] Arun N. Lokanathan and Jay B. Brockman. A methodology for concurrent process-
circuit optimization. IEEE Transactions on Computer-Aided Design of Circuits and
Systems, 18(7):889–902, July 1999.
[79] D. Luenberger. Optimization By Vector Space Methods. John Wiley, New York, 1969.
[81] Pradip Mandal and V. Visvanathan. CMOS Op-Amp sizing using a geometric pro-
gramming formulation. IEEE Transactions on Computer-Aided Design of Circuits and
Systems, 20(1):22–38, January 2001.
[82] H. Alan Mantooth and Mike F. Fiegenbaum. Modeling with an Analog Hardware
Description Language. Kluwer Academic Publishers, November 1994.
[83] P. Maulik, L. R. Carley, and R. Rutenbar. Integer programming based topology selection
of cell-level analog circuits. IEEE Transactions on Computer-Aided Design of Circuits
and Systems, 14(4):401ff, April 1995.
References 189
[84] Petra Michel, Ulrich Lauther, and Peter Duzy. The Synthesis Approach to Digital System
Design. Kluwer Academic Publishers, Boston, 1992.
[85] Gordon E. Moore. Cramming more components onto integrated circuits. Electronics,
38(8), April 1965.
[86] Daniel Mueller, Guido Stehr, Helmut Graeb, and Ulf Schlichtmann. Deterministic
approaches to analog performance space exploration (PSE). In ACM/IEEE Design
Automation Conference (DAC), June 2005.
[88] MunEDA. WiCkeD – Design for Manufacturability and Yield. www.muneda.com, 2001.
[90] Dongkyung Nam and Cheol Hoon Park. Multiobjective simulated annealing: A compar-
ative study to evolutionary algorithms. International Journal of Fuzzy Systems, pages
87–97, June 2000.
[91] Jorge Nocedal and Stephen J. Wright. Numerical Optimization. Springer, 1999.
[93] E. Ochotta, T. Mukherjee, R.A. Rutenbar, and L.R. Carley. Practical Synthesis of High-
Performance Analog Circuits. Kluwer Academic Publishers, 1998.
[94] Emil S. Ochotta, Rob A. Rutenbar, and L. Richard Carley. Synthesis of high-performance
analog circuits in ASTRX/OBLX. IEEE Transactions on Computer-Aided Design of
Circuits and Systems, 15(3):273–294, March 1996.
[96] Rodney Phelps, Michael Krasnicki, Rob A. Rutenbar, L. Richard Carley, and James R.
Hellums. ANACONDA: Simulation-based synthesis of analog circuits via stochastic
pattern search. IEEE Transactions on Computer-Aided Design of Circuits and Systems,
19(6):703–717, June 2000.
[98] Ming Qu and M. A. Styblinski. Parameter extraction for statistical ic modeling based
on recursive inverse approximation. IEEE Transactions on Computer-Aided Design of
Circuits and Systems, 16(11):1250–1259, 1997.
[99] Joao Ramos, Kenneth Francken, Georges G. E. Gielen, and Michiel S. J. Steyaert. An
efficient, fully parasitic-aware power amplifier design optimization tool. IEEE Trans-
actions on Circuits and Systems I: Fundamental Theory and Applications, 2005.
190 REFERENCES
[100] Carl R. C. De Ranter, Geert Van der Plas, Michiel S. J. Steyaert, Georges G. E. Gielen,
and Willy M. C. Sansen. CYCLONE: Automated design and layout of RF LC-oscillators.
IEEE Transactions on Computer-Aided Design of Circuits and Systems, 21(10):1161–
1170, October 2002.
[101] A. Ruehli(Editor). Circuit Analysis, Simulation and Design. Advances in CAD for VLSI.
North-Holland, 1986.
[102] Youssef G. Saab and Vasant B. Rao. Combinatorial optimization by stochastic evolution.
IEEE Transactions on Computer-Aided Design of Circuits and Systems, 10(4):525–535,
April 1991.
[103] T. Sakurai, B. Lin, and A. Newton. Fast simulated diffusion: an optimization algorithm
for multiminimum problems and its application to MOSFET model parameter extraction.
IEEE Transactions on Computer-Aided Design of Circuits and Systems, 11:228–233,
1992.
[104] Sachin S. Sapatnekar, Vasant B. Rao, Pravin M. Vaidya, and Sung-Mo Kang. An exact
solution to the transistor sizing problem for CMOS circuits using convex optimization.
IEEE Transactions on Computer-Aided Design of Circuits and Systems, 12(11):1621–
1634, November 1993.
[105] M. Sharma and N. Arora. OPTIMA: A nonlinear model parameter extraction program
with statistical confidence region algorithms. IEEE Transactions on Computer-Aided
Design of Circuits and Systems, 12:982–987, 1993.
[106] C.-J. Richard Shi and Xiang-Dong Tan. Canonical symbolic analysis of large analog
circuits with determinant decision diagrams. IEEE Transactions on Computer-Aided
Design of Circuits and Systems, 19(1):1–18, January 2000.
[107] C.-J. Richard Shi and Xiang-Dong Tan. Compact representation and efficient generation
of s-expanded symbolic network functions for computer-aided analog circuit design.
IEEE Transactions on Computer-Aided Design of Circuits and Systems, 20(7):813, July
2001.
[108] Guoyong Shi, Bo Hu, and C.-J. Richard Shi. On symbolic model order reduction. IEEE
Transactions on Computer-Aided Design of Circuits and Systems, 2006.
[109] C. Spanos and S. Director. Parameter extraction for statistical IC process characteriza-
tion. IEEE Transactions on Computer-Aided Design of Circuits and Systems, 5:66–78,
1986.
[110] Thanwa Sripramong and Christofer Toumazou. The invention of cmos amplifiers using
genetic programming and current-flow analysis. IEEE Transactions on Computer-Aided
Design of Circuits and Systems, 2002.
[111] H.H. Szu and R.L. Hartley. Nonconvex optimization by fast simulated annealing. Pro-
ceedings of the IEEE, 75:1538–1540, 1987.
[112] Sheldon X.-D. Tan. A general hierarchical circuit modeling and simulation algorithm.
IEEE Transactions on Computer-Aided Design of Circuits and Systems, 2005.
References 191
[113] Sheldon X.-D. Tan and C.-J. Richard Shi. Efficient approximation of symbolic expres-
sions for analog behavioral modeling and analysis. IEEE Transactions on Computer-
Aided Design of Circuits and Systems, 2004.
[114] Xiangdong Tan and C.-J. Richard Shi. Hierarchical symbolic analysis of large analog
circuits with determinant decision diagrams. In IEEE International Symposium on
Circuits and Systems (ISCAS), page VI/318, 1998.
[115] Hua Tang, Hui Zhang, and Alex Doboli. Refinement-based synthesis of continuous-time
analog filters through successive domain pruning, plateau search, and adaptive sampling.
IEEE Transactions on Computer-Aided Design of Circuits and Systems, 2006.
[116] Antonio Torralba, Jorge Chavez, and Leopoldo G. Franquelo. FASY: A fuzzy-logic based
tool for analog synthesis. IEEE Transactions on Computer-Aided Design of Circuits and
Systems, 15(7):705–715, July 1996.
[117] Wim M. G. van Bokhoven and Domine M. W. Leenaerts. Explicit formulas for the
solutions of piecewise linear networks. IEEE Transactions on Circuits and Systems I:
Fundamental Theory and Applications, 46(9):1110ff., September 1999.
[118] Geert Van der Plas, Geert Debyser, Francky Leyn, Koen Lampaert, Jan Vandenbussche,
Georges Gielen, Willy Sansen, Petar Veselinovic, and Domine Leenaerts. AMGIE–A
synthesis environment for CMOS analog integrated circuits. IEEE Transactions on
Computer-Aided Design of Circuits and Systems, 20(9):1037–1058, September 2001.
[121] Jacob K. White and Alberto Sangiovanni-Vincentelli. Relaxation Techniques for the
Simulation of VLSI Circuits. Kluwer Academic Publishers, 1987.
[122] Claudia Wieser. Schaltkreisanalyse mit Worst-Case Abstaenden. PhD thesis, Technische
Universitaet Muenchen, 1994.
[123] J. Wojciechowski and J. Vlach. Ellipsoidal method for design centering and yield estima-
tion. IEEE Transactions on Computer-Aided Design of Circuits and Systems, 12:1570–
1579, 1993.
[124] X. Xiangming and R. Spence. Trade-off prediction and circuit performance optimization
using a second-order model. International Journal of Circuit Theory and Applications,
20:299–307, 1992.
[125] D. Young, J. Teplik, H. Weed, N. Tracht, and A. Alvarez. Application of statistical design
and response surface methods to computer-aided VLSI device design II: desirability
functions and Taguchi methods. IEEE Transactions on Computer-Aided Design of
Circuits and Systems, 10:103–115, 1991.
192 REFERENCES
[126] T. Yu, S. Kang, I. Hajj, and T. Trick. Statistical performance modeling and parametric
yield estimation of MOS VLSI. IEEE Transactions on Computer-Aided Design of
Circuits and Systems, 6:1013–1022, 1987.
[127] T. Yu, S. Kang, J. Sacks, and W. Welch. Parametric yield optimization of CMOS analogue
circuits by quadratic statistical circuit performance models. International Journal of
Circuit Theory and Applications, 19:579–592, 1991.