scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Buffer delay change in the presence of power and ground noise

01 Jun 2003-Vol. 11, Iss: 3, pp 461-473
TL;DR: An application shows that repeater chains using buffers instead of inherently faster inverters tend to have superior supply-level-induced jitter characteristics, consistent with short-channel MOSFET behavior, including carrier velocity saturation effects.
Abstract: Variations of power and ground levels affect very large scale integration circuit performance. Trends in device technology and in packaging have necessitated a revision in conventional delay models. In particular, simple scalable models are needed to predict delays in the presence of uncorrelated power and ground noise. In this paper, we analyze the effect of such noise-on-signal propagation through a buffer and present simple, closed-form formulas to estimate the corresponding change of delay. The model captures both positive (slowdown) and negative (speedup) delay changes. It is consistent with short-channel MOSFET behavior, including carrier velocity saturation effects. An application shows that repeater chains using buffers instead of inherently faster inverters tend to have superior supply-level-induced jitter characteristics. The expressions can be used in any existing circuit performance optimization design flow or can be combined into any delay calculations as a correction factor.

Summary (4 min read)

1. INTRODUCTION

  • This paper describes a new model for the change in buffer delay caused by both power and ground supply level variations and level variations between stages in sequences of repeaters.
  • The net effect of this is de-correlation of the power and ground voltage variations which in turn make delay variations much more complex.
  • These expressions are intended for inclusion in timing analysis tools and statistical delay estimators as corrections to nominal delay models that account for other effects.
  • Furthermore, they are shown to be largely independent of the buffer load circuit structure, increasing their applicability.
  • The rest of the paper is organized as follows: sections 2-3 define P/G noise, buffer delay nomenclature, and illustrate the P/G-noise-induced buffer delay change.

2.1 Variation of Vdd and Vss

  • The authors use Vdd, Vss, Vi , Vin (adjusted ), Vo , Vout (adjusted ), etc. to represent values related to the ideal power and ground levels, and they use Vdd’, Vss’, Vi’, Vin’, Vo’, Vout’, etc. to represent the corresponding values in the presence of power and ground noise.
  • ∆Vdd and ∆Vss denote the variation of power supply and ground, respectively.
  • In modern bump-bonded and low-inductance package styles, the package distributes power over the whole area of the chip .
  • This noise is increased by the inclusion (within a bump block) of long wire repeater buffers which are often added in a post-placement timing optimization step.
  • On the other hand, switching of the C and E buffers has a non-symmetric effect on and , because they drive loads which are outside of block 1, causing different switching currents to flow through the power and ground ports of block 1.

To allow reference to previous work, it is necessary to formally define the model for buffer delay, as the measurement levels are subject to noise.

  • Without loss of generality, the authors define a buffer’s ideal delay as the time interval between its input and output voltage reaching 50% of the power supply level.
  • In figure 3(a), the authors assume a bump-bond packaging technique, in which every bump has a small power/ground network, relatively independent of the others.
  • Therefore the low and high voltage levels of the input transition ( ) are independent of the change of power supply and ground level for the buffer.
  • The corresponding buffer delays are and .
  • Therefore, , With a rising transition at the input, the changes of delay and output transition time are defined as follows:.

3.1 Differential mode noise

  • The authors define the differential mode noise (DMN) ∆Vdif as: where and The authors have observed through HSpice simulations that the voltage difference (Vdif) between the power supply and ground level affects the inverter delay (tpHL/tpLH), which is the inverter’s ability to propagate signals.
  • Similarly, the change of the above difference (∆Vdif) affects the change of delay (∆tpHL/∆tpLH).
  • The differential mode noise ∆Vdif may become positive or negative, depending on the directions and relative amplitudes of ∆Vdd and ∆Vss.
  • The voltage difference (Vdif) between power supply and ground level determines how fast the buffer charges/discharges its capacitive load, so it affects the delay and the transition time of the buffer output.

3.2 Common mode noise

  • The authors define the common mode noise (CMN) ∆Vcom as: CMN modifies the effective switching threshold of the gate.
  • The threshold shift changes the gate delay as illustrated in figure 4. Figure 4(a) shows a rising transition arriving at the buffer.
  • The points N’ and P’ can be mapped to shifted thresholds in figure 4(a) with ideal power supply and ground level: Obviously, it takes more time for the NFET transistor to be turned on and PFET to be turned off when both power supply and ground level increase.
  • Therefore, the authors make the following observation: Observation 2: For an input with a rising transition, the dependency between the common mode noise (CMN) and the buffer delay change can be expressed by: (6) where kcr is a positive constant determined by the device and technology parameters, input transition time, and the gate load.
  • Therefore, common mode noise has a different effect on rising and falling transitions.

3.3 Loading effects

  • Both differential mode noise (DMN) and common mode noise (CMN) change buffer delays.
  • Since the delay change can be of either sign, the noise sources need to be modeled together.
  • Figure 5 shows alternative load configurations and the corresponding simulated delay change (both rising and falling transition) in 0.18µm technology.

In figure 5 (a), the wire load of the inverter is a distributed

  • RC tree network, including vias, extracted from the layout of a real circuit.
  • As shown in figure 5, for a rising transition, with the increase of ∆Vcom from -0.4 to 0.4 volt, delay change increases.
  • Observation 3 indicates that an appropriate simplified wire load model can be used for delay modeling of the buffer itself.
  • This noise contributes to the local buffer/inverter delay in a complex way which can either increase or decrease the signal delay.

4.1 Inverter model

  • In deep submicron circuits, carrier velocity saturation effects predominate.
  • From Figure 6, the authors have the following nodal equation: where (For the technologies reported in the results, short circuit errors amounted to less than 5% for incremental delay and 20% for incremental transition time, over all simulated cases.).
  • For a rising transition at the inverter input, short circuit current from the NFET is negligible.
  • The authors assume that for a given transistor with a given loading capacitance, the above four values remain unchanged with a small disturbance of power supply and ground levels.

4 and 5.

  • Equations (11) to (15) demonstrate that the incremental change of buffer delay is linear with respect to the power and ground variations, also known as Theorem 1.
  • The coefficient k1 quantifies the effect of the common mode noise while k2 characterizes the effect of the differential mode noise on buffer delay.
  • This theorem shows why the observations in sections 3.1 through 3.3 hold.
  • Under certain circumstances, the input transition time can be changed by the P/G noise.

4.4.1 The trend with technological scaling

  • In classic wire-bond packaging, the power and ground noise were closely correlated for the whole chip.
  • Hence it was appropriate to measure the delay at the 50% point of the changed power supply and ground level.
  • The authors considered as the default value for the high-to-low delay change.
  • This matches the simulation results shown in figure 5 and observation 4.

4.4.2 Power v.s. ground variation

  • As the authors mentioned in section 2.1, power and ground variations are related, but not necessarily correlated.
  • Usually IR-drop is discussed when assuming and .
  • When the changes of power and ground are in the same direction and have similar amplitudes, the authors have .
  • According to equation (11), the buffer delay change will be dominated by the effect induced by common mode noise for .

4.4.3 Definitions of delay

  • The expressions for delay and slope depend on how the authors define and .
  • Comparing these four different delay definitions, the authors can see that the same P/ G noise has different effects on different ∆delay.

4.4.4 Change of waveform

  • Delay expressions will be different when the input/output waveforms differ from the ones shown in figure 2-3.
  • Through HSPICE simulations, the authors have observed that the cases shown in figure 2-3 are typical for deep submicron circuits.
  • When to5 and to5’ fall in region 2 or region 4, the expression for becomes a little more complicated, but is still closed form.

5. MODEL VALIDATION

  • The authors validated their model in both 0.25µm and 0.18µm technologies.
  • As mentioned in section 4.1, the alpha-power law MOSFET model relies on four parameters: α, Vtn, ID0 (drain current at VGS = VDS = Vdd), VDO (drain saturation voltage at VGS = Vdd).
  • The approximations made in the incremental delay change model work better for fast slew rates since the short circuit current is reduced, and feed-forward capacitive coupling is modeled in the nominal delay.
  • The model is not as accurate in estimating the change of transition time (∆toT).
  • Comparison of the technologies shows the trend of increasing sensitivity to supply-level noise with scaling.

6. APPLICATIONS

  • An important feature of the model is its relative lack of dependence on the circuit loading structure.
  • Iterations may result, because the updated delay will further affect the power and ground level.
  • Since their formulas are very simple, an iterative process may be affordable.
  • This section will show some applications of their modeling technique.

6.2 Different buffer chains

  • The linear relationship between the P/G noise and delay change can be used to analyze the delivered jitter for a chain of single-inverter-buffers and a chain of double-inverterbuffers, as shown in figure 7.
  • And the authors assume the parameters for both buffer designs are comparable.
  • This has been experimentally demonstrated by their simulation results in figure 5, and theoretically proved by their equation (11) which indicates .
  • When the amplitude of the common mode noise is at least as large as that of the differential mode noise, the delay change of both buffer designs will be dominated by common mode noise.
  • Delivered jitter (total delay change of the buffer chain) for the double-inverter buffers is smaller than that of a singleinverter buffer chain.

7. CONCLUSION

  • Maintaining signal integrity in deep submicron circuits is a difficult problem.
  • Variations of power and ground levels play an important role because this type of noise significantly degrades circuit performance.
  • Deep submicron circuits have decreased power supply level Vdd and decreased velocity saturation index α, leading to increased sensitivity of delay to P/G noise.
  • Thus, despite the reduction of noise from lower-inductance packaging, the relative magnitude of the delay changes is still a serious potential problem.
  • An application in clock buffer chain design shows that repeater chains, using buffers instead of inherently faster inverters, tend to have superior level - induced delay characteristics.

Did you find this useful? Give us your feedback

Figures (9)

Content maybe subject to copyright    Report

1
ABSTRACT
Variations of power and ground levels affect VLSI circuit
performance. Trends in device technology and in packaging
have necessitated a revision in conventional delay models. In
particular, simple scalable models are needed to predict
delays in the presence of uncorrelated power and ground
noise. In this paper, we analyze the effect of such noise on
signal propagation through a buffer and present simple,
closed-form formulas to estimate the corresponding change
of delay. The model captures both positive (slowdown) and
negative (speedup) delay changes. It is consistent with short-
channel MOSFET behavior, including carrier velocity satu-
ration effects. An application shows that repeater chains
using buffers instead of inherently faster inverters tend to
have superior supply-level-induced jitter characteristics. The
expressions can be used in any existing circuit performance
optimization design flow or can be combined into any delay
calculations as a correction factor.
Categories & Subject Descriptors: [Computer-
Aided Engineering]: Computer-aided design (CAD).
General Terms: Algorithms.
Keywords: Power and ground noise, differential mode
noise, common mode noise, incremental delay change.
1. INTRODUCTION
This paper describes a new model for the change in buffer
delay caused by both power and ground supply level
variations and level variations between stages in sequences
of repeaters. These delay changes are a large component of
the total timing jitter for a signal where the jitter accounts for
all noise sources such as substrate noise and coupling noise
as well as power level noise. There is a substantial amount of
previous work in this area, notably papers: [4][6][13][14]
[19]. However, for several reasons described below, we
believe that the problem bears re-examination and a renewed
effort to create a fast, simple model suitable for mass
implementation in a modern design flow.
Growth in design sizes and scaling of interconnections have
led to the requirement for insertion of very large numbers of
buffer/repeaters in recent designs [1][7][9][15]. Among
them, [9] proposes an optimum multistage buffer design to
drive long uniform lines. [7] and [15] consider simultaneous
buffer insertion and wire sizing for timing optimization. [1]
presents comprehensive buffer insertion techniques for noise
and delay optimization. Because of the buffer/repeaters’
preponderance in number, use in heavily loaded nets, and use
in clock and timing circuits, buffer delays account for a large
percentage of all critical timing nets in a design. In some of
these applications, total timing uncertainty (not just worst-
case delay) is important. Disturbance of the buffer delay will
affect the type, number and position of buffers that optimize
the interconnect delay. Therefore, it is essential to take the
change of delay into consideration in order to adjust the
solution for timing optimization. At the same time, scaling
of power supply levels and improving transconductance of
devices have increased the sensitivity of buffers to supply-
level-induced delays. Finally, increases in chip-level design
scales and modern packaging strategies such as bump
bonding have localized supply variations so that buffers in
one set of supply levels are driving buffers in another zone
with different supply levels. Since power loading is logic
switching-dependent and supply sources are localized,
power and ground levels need not be inversely correlated as
is typical in a wire bonded die.
Under such conditions, power-level-induced delay changes
may either increase or decrease the effective delay of a
buffer, and successive stages may or may not accumulate
incremental delays. One must consider both power and
ground levels at the signal source and at the current buffer to
derive an equivalent delay change. This value can be
substantially smaller than that predicted by superposing
ground-bounce and power level changes [3][5][10][20][21].
The superposition approximation works well only if the
variations in the power distribution network are mirrored in
the ground network. However, due to changes in packaging
technology and the number of pins that can feasibly be
devoted to power and ground connections, no single parasitic
dominates the noise on the power and ground nets. Yet
another trend in technology is the relative reduction of gate
parasitics compared to those from the interconnection
network. The net effect of this is de-correlation of the power
and ground voltage variations which in turn make delay
variations much more complex. In particular, the worst-case
Buffer Delay Change in the Presence of Power and Ground
Noise
Lauren Hui Chen*, Malgorzata Marek-Sadowska**, Forrest Brewer**
*Synopsys Inc., Mountain View, CA, USA
** Electrical and Computer Engineering Department, University of California, Santa Barbara, CA 93106, USA
*laurenc@synopsys.com, **{mms, forrest}@ece.ucsb.edu

2
delay caused by power noise may not occur simultaneously
with the worst-case delay due to ground noise, and the
superposition may cause a substantial overestimation.
Second, the delay effects of common mode voltage shifts
will be shown to be larger in scale than equivalent
differential mode changes. (Differential mode voltage shifts
are the most commonly studied model). Lastly, changes in
power distribution and clocking strategies, and the potential
for future changes, create the need for a timing model which
is independent of common assumptions about power level
noise sources. We do assume that large scale power level
changes result from an ensemble effect of many devices at a
variety of differing slew rates. However, in a practical
design there must always be some amount of local
decoupling capacitance (both parasitic and added). This
capacitance limits the magnitude of the highest speed noise
excursions. (This cannot be done in later stages of the power
network design because of inductance in both the physical
network and the packaged capacitors). We therefore assume
that the remaining large magnitude noise excursions occur
at a somewhat slower time scale than the typical switching
transitions in buffers meeting common design requirements.
In the following, we analyze the effect of P/G (power/
ground) noise on buffer delay, and present linear, closed-
form formulas for the corresponding incremental changes in
delay based on a short-channel transistor model. The
expressions simultaneously model both the power supply
and ground levels, resulting in positive (slowdown) or
negative (speedup) delay changes. These expressions are
intended for inclusion in timing analysis tools and statistical
delay estimators as corrections to nominal delay models that
account for other effects. The expressions make few
assumptions about the specific shape of the P/G
noise waveform. Furthermore, they are shown to be largely
independent of the buffer load circuit structure, increasing
their applicability.
The rest of the paper is organized as follows: sections 2-3
define P/G noise, buffer delay nomenclature, and illustrate
the P/G-noise-induced buffer delay change. Section 4
presents our new model. Section 5 demonstrates the
accuracy and fidelity of the model. Applications of the
model and concluding remarks are presented in sections 6-
7. Detailed derivations are given in Appendices A and B.
2. BUFFER DELAY CONVENTIONS
A buffer is a chain of tapered inverters. Here, we consider
buffers consisting of one inverter or two inverters.
2.1 Variation of V
dd
and V
ss
We use V
dd
, V
ss
, V
i
(input), V
in
(adjusted input), V
o
(output),
V
out
(adjusted output), etc. to represent values related to the
ideal power and ground levels, and we use V
dd
, V
ss
, V
i
,
V
in
, V
o
, V
out
, etc. to represent the corresponding values in
the presence of power and ground noise. V
dd
and V
ss
denote the variation of power supply and ground,
respectively.
(power noise) (1)
( ) (ground noise) (2)
In small-scale wire-bond packaging styles, a dominant
supply level noise source is bond wire inductance in the
package. Neglecting I/O current drives, the power and
ground noise of the chip due to simultaneous switching
typically follows an inverse pattern, and V
dd
is often
symmetric to V
ss.
However, in modern bump-bonded and
low-inductance package styles, the package distributes
power over the whole area of the chip (figure 1). Every
bump connects to an underlying local power/ground
network. To save chip metallization area and improve
density, global power distribution metal is reduced in
preference to thicker package distribution layers. Increasing
design scales causes an increase in long wire loading, and in
more wires connecting between different power domains.
Logic-level-dependent currents flow between such blocks,
causing asymmetric power and ground noise within a block.
This noise is increased by the inclusion (within a bump
block) of long wire repeater buffers which are often added
in a post-placement timing optimization step.
Figure 1 shows a simple circuit which uses bump-bond
packaging. Each block is defined by the subcircuit supplied
by a pair of bumps (V
dd
/V
ss
). Suppose that the cells A to E
have transitions. Switching of the buffers A, B and D has a
symmetric effect on the power and ground noise ( and
), because they drive loads (consisting largely of
parasitic interconnect capacitance) within the same block.
On the other hand, switching of the C and E buffers has a
non-symmetric effect on and , because they
drive loads which are outside of block 1, causing different
switching currents to flow through the power and ground
ports of block 1. Since wires leaving a block are likely to be
physically long, these currents are proportionally large. In
general, there is little reason for the currents flowing out of
the block via loads to cancel.
2.2 Incremental buffer delay change
To allow reference to previous work, it is necessary to
formally define the model for buffer delay, as the
measurement levels are subject to noise.
Without loss of generality, we define a buffer’s ideal delay
as the time interval between its input and output voltage
reaching 50% of the power supply level. Figure 2 illustrates
the definition for an inverter delay given ideal power supply
and ground levels. t
pHL
is the high-to-low delay when the
input of the inverter has a rising transition. The input and
output transition times are t
r
and t
oT
, respectively. Other
time values are: t
i5
, t
o5
, t
o1
and t
o9,
which are times when the
V
dd
V
dd
V
dd
=
V
ss
V
ss
V
ss
V
ss
==
V
ss
0=
V
dd
1( )
V
ss
1( )
V
dd
1( )
V
ss
1( )

3
input or output voltage reaches 50%, 10%, and 90% of V
dd
,
respectively.
(3)
(4)
Figure 3 (a) and (b) illustrate the delay and slope when there
is a P/G noise. In figure 3(a), we assume a bump-bond
packaging technique, in which every bump has a small
power/ground network, relatively independent of the others.
Therefore the low and high voltage levels of the input
transition ( ) are independent of the change of
power supply and ground level for the buffer. The
corresponding buffer delays are and . In figure
3(b), conventional wire-bond packaging technique is
assumed. For such a technique, the dominant noise on the
power supply and ground level is often due to the wiring
inductance in the package, so the whole chip’s power and
ground noises are synchronized. Therefore, the voltage level
of the input transition ( ) will be the same as that
of the buffer. The corresponding buffer delays are and
. Because of power supply and ground level changes,
we are interested in delay measured at different voltage
levels: the 50% point between the ideal V
dd
and V
ss
(
and ), and the 50% point between the disturbed V
dd
and V
ss
( and ). Hence we have four types of
disturbed buffer delays, corresponding to four types of
buffer delay changes.
The special output voltage points are defined as follows:
The disturbed high-to-low delay and slope are given by:
where , indicating four different definitions of
the disturbed buffer delay and output slope illustrated in
figure 3 (a) & (b).
In figure 3(a) we have and . In figure
3(b) we have and . Therefore,
,
With a rising transition at the input, the changes of delay
and output transition time are defined as follows:
The results shown in figure 5 are according to the first type
of delay definition with j = 1.
3. P/G NOISE DELAY EFFECTS
The changes of power supply and ground level affect signal
propagation through an inverter in several different ways,
which will be discussed in this section.
C
A
B
D E
BLOCK 1
BLOCK 2
V
dd
’(2)
V
ss
’(2)
V
dd
’(1)
V
ss
’(1)
V
dd
Pin
Parasitics
V
ss
Pin
Parasitics
Local Redistribution
Parasitics
Local Redistribution
Parasitics
V
dd
V
ss
V
dd
Package
Redistribution
Layer / Plane
V
ss
Package
Redistribution
Layer / Plane
Figure 1. Power distribution in bump-bond packaging
t
pHL
t
o5
t
i5
=
t
oT
t
o1
t
o9
0.8
--------------------=
t
pHL
1( )
t
pHL
2( )
t
pHL
3( )
t
pHL
4( )
t
pHL
1( )
t
pHL
3( )
t
pHL
2( )
t
pHL
4( )
V
dd
0.9V
dd
0.5V
dd
V
D0
V
tn
0.1V
dd
V
tn
V
dd
---------
t
r
t
o9
t
r
t
o5
t
D0
t
o1
time
V
i
(input)
V
o
(output)
V
ss
voltage
1 2 3 4
region
NMOS CUT SAT SAT LIN
V
TR
t
i5
t
pHL
0.8t
oT
Figure 2. Ideal inverter transition and nomenclature
V
o1
V
ss
0.1 V
dd
V
ss
( )+=
V
o5
V
ss
0.5 V
dd
V
ss
( )+=
V
o9
V
ss
0.9 V
dd
V
ss
( )+=
t
pHL
j( )
t
o5
j( )
t
i5
j( )
=
t
oT
j( )
t
o1
j( )
t
o9
j( )
0.8
------------------------=
j 1 2 3 4, , ,=
t
o1
1( )
t
o1
2( )
=
t
o9
1( )
t
o9
2( )
=
t
o1
3( )
t
o1
4( )
=
t
o9
3( )
t
o9
4( )
=
t
oT
1( )
t
oT
2( )
=
t
oT
3( )
t
oT
4( )
=
t
pHL
j( )
t
pHL
j( )
t
pHL
=
t
oT
j( )
t
oT
j( )
t
oT
=

4
3.1 Differential mode noise
We define the differential mode noise (DMN) V
dif
as:
where
and
We have observed through HSpice simulations that the
voltage difference (V
dif
) between the power supply and
ground level affects the inverter delay (t
pHL
/t
pLH
), which is
the inverter’s ability to propagate signals. Similarly, the
change of the above difference (V
dif
) affects the change of
delay (t
pHL
/t
pLH
).
The differential mode noise V
dif
may become positive or
negative, depending on the directions and relative
amplitudes of V
dd
and V
ss
. The voltage difference (V
dif
)
between power supply and ground level determines how fast
the buffer charges/discharges its capacitive load, so it affects
the delay and the transition time of the buffer output. The
larger the difference ( ), the faster the
output charging/discharging and the smaller the buffer delay
( ), which is stated in observation 1.
Observation1: The buffer delay change is linearly dependent
on the differential mode noise (DMN), as will be shown in
section 4:
(5)
where k
d
is a positive constant dependent on the device and
technology parameters, the input transition times, and the
gate load. The expression of k
d
can be found from equation
(A5) in Appendix B. Similar effects hold for both high-to-
low delay t
pHL
and low-to-high delay t
pLH
.
3.2 Common mode noise
We define the common mode noise (CMN) V
com
as:
CMN modifies the effective switching threshold of the gate.
The threshold shift changes the gate delay as illustrated in
figure 4. Figure 4(a) shows a rising transition arriving at the
buffer. Figure 4(b) illustrates the gate threshold shift and the
corresponding delay change caused by the power and
ground variations.
In figure 4(b), N and P are the original points when the
NFET switches from cutoff to saturation region and PFET
switches from saturation to cutoff, respectively. The
V
dd
0.5V
dd
time
(input) (output)
V
ss
voltage
t
pHL
0.8t
oT
,
(2)
V
dd
t
pHL
,
(1)
t
i5
,
(1)
t
i5
,
(2)
t
o5
,
(1)
t
o1
,
t
o5
,
(2)
t
o9
,
,
V
o9
,
V
o5
,
V
o1
,
V
ss
,
V
o
,
V
i
,
V
ss
V
dd
V
dd
V
ss
V
ss
V
dd
Figure 3. Notation for disturbed delay and slope
V
dd
0.5V
dd
time
(input) (output)
V
ss
voltage
t
pHL
,
(4)
V
dd
t
pHL
,
(3)
t
i5
,
(3)
t
i5
,
(4)
t
o5
,
(3)
t
o1
,
t
o5
,
(4)
t
o9
,
,
V
o9
,
V
o5
,
V
o1
,
V
ss
,
V
o
,
V
i
,
0.8t
oT
V
ss
V
dd
V
ss
V
dd
V
dd
V
ss
(a) (b)
V
dif
V
dif
V
dif
V
dd
V
ss
= =
V
di f
V
dd
V
ss
=
V
dif
V
dd
V
ss
=
V
di f
V
dif
> V
dif
0>
t
pHL
t
pHL
< t
pHL
0<
t
pHL
DMN
k
d
V
dif
k
d
= V
dd
V
ss
( )=
V
com
V
dd
V
ss
+=
V
dd
V
i
V
GS
(P)
V
GS
(N)
V
ss
V
GS
P( )
V
i
V
dd
=
V
GS
N( )
V
i
V
ss
=
V
dd
V
dd
,
V
ss
V
ss
,
V
ss
V
dd
V
tn
V
tn
|V
tp
|
|V
tp
|
V
tn
(shift)
|V
tp
|
(shift)
t
n
,
t
p
,
t
n
t
p
(a)
(b)
N’
P’
P
N
Figure 4. Threshold shift of an inverter
wire
load
V
o

5
corresponding switching times are indicated t
n
and t
p
. For
noise of limited amplitude, the transistor thresholds (V
tn
and
V
tp
) do not change significantly, namely,
This causes a shift of the NFET and PFET switching points
from N and P to N’ and P’. The corresponding switching
time shifts to t
n
and t
p
, respectively.
The points N’ and P’ can be mapped to shifted thresholds in
figure 4(a) with ideal power supply and ground level:
We observe in figure 4(b) that:
Obviously, it takes more time for the NFET transistor to be
turned on and PFET to be turned off when both power
supply and ground level increase.
On the other hand, when power supply and ground level
decrease, we have
Hence, it takes less time for the NFET transistor to be
turned on and PFET to be turned off for decreased power
and ground level.
Therefore, we make the following observation:
Observation 2: For an input with a rising transition, the
dependency between the common mode noise (CMN) and
the buffer delay change can be expressed by:
(6)
where k
cr
is a positive constant determined by the device
and technology parameters, input transition time, and the
gate load. The expression of k
cr
can be found from equation
(A5) in Appendix B.
Similarly, for an input with a falling transition, the
dependency between the common mode noise (CMN) and
the buffer delay change can be expressed by:
(7)
where k
cf
ia a positive constant determined by the device
and technology parameters, input transition time, and the
gate load.
For a rising transition, the effective switching threshold of
an inverter is set by the current balance of the two active
transistors. For positive common mode noise, this switching
threshold is higher, and therefore it is reached later by the
rising transition. Thus the delay is increased even though the
voltage across the inverter, and hence the current drive,
remains constant.
For a falling transition, the effective switching threshold of
the gate rises, so that the threshold voltage level is reached
earlier, and the effective gate delay decreases. This effect
occurs even if the differential mode noise is zero, as it is the
switching level - not the current drive - that is altered by the
common mode noise.
We note an analogy here: a rising input transition with
positive common mode noise can be thought of as climbing
a rising mountain, which takes more time than climbing a
mountain of initially the same, yet stationary, dimensions. A
falling input transition with the same positive common
mode noise is analogous to going downhill when the bottom
of the mountain rises, which takes less time than descending
a corresponding fixed dimensions mountain.
Therefore, common mode noise has a different effect on
rising and falling transitions.
3.3 Loading effects
Both differential mode noise (DMN) and common mode
noise (CMN) change buffer delays. Since the delay change
can be of either sign, the noise sources need to be modeled
together. Figure 5 shows alternative load configurations and
the corresponding simulated delay change (both rising and
falling transition) in 0.18µm technology. Note: delay = 0
when V
dd
= V
ss
= 0.
In figure 5 (a), the wire load of the inverter is a distributed
RC tree network, including vias, extracted from the layout
of a real circuit. In figure 5 (b), the wire load is simplified to
an RC π-model. In figure 5 (c), the wire load is further
simplified to a single resistor plus a single capacitor. In
figure 5 (d), an effective loading capacitor is used to replace
the inverter’s output load. These simplified wire-load
models in figures 5(b)-(d) can be obtained using the
methods described in [16]. The delay is measured when its
input (V
i
) and output (V
o
/ V
o
/ V
o
/ V
o
”’) voltage reach
50% of the ideal power supply voltage, respectively. The
range for the power and ground noise is from -20% (-0.36
volt) to 20% (0.36 volt) of the power supply voltage, which
is set to 1.8 volt for the selected technology. The range for
the change of delay is from -30ps to 30ps.
It is interesting to note that each of the four wire load
models displays a linear relationship between the change of
power/ground level and the change of inverter delay.
Furthermore, the linearity improves when the change of
power and ground level is smaller than 20%. In practical
designs, the tolerable range for the power and ground levels
V
GS
N( )
CUT
V
i
V
ss
V
tn
0>= =
V
GS
P( )
CU T
V
i
V
dd
V
tp
0<= =
V
GS
N( )
CU T
V
i
V
ss
V
tn
sh ift( )
= =
V
GS
P( )
CUT
V
i
V
dd
V
tp
sh ift( )
= =
V
ss
0> V
tn
sh ift( )
V
tn
> t
n
t
n
> increasing delay
V
dd
0> V
tp
shift( )
V
tp
> t
p
t
p
> increasing delay
V
ss
0< V
tn
sh ift( )
V
tn
< t
n
t
n
< decreasing delay
V
dd
0< V
tp
shift( )
V
tp
< t
p
t
p
< decreasing delay
t
pHL
CM N
k
cr
V
co m
k
cr
= V
dd
V
ss
+( )=
t
pL H
CMN
k
cf
V
com
k
cf
= V
dd
V
ss
+( )=

Citations
More filters
Proceedings ArticleDOI
02 Apr 2005
TL;DR: The critical dimensions beyond which optical interconnect becomes advantageous over electrical interconnect are shown to be approximately one tenth of the chip edge length at the 22 nm technology node.
Abstract: Interconnect has become a primary bottleneck in integrated circuit design As CMOS technology is scaled, it will become increasingly difficult for conventional copper interconnect to satisfy the design requirements of delay, power, bandwidth, and noise On-chip optical interconnect has been considered as a potential substitute for electrical interconnect in the past two decades In this paper, predictions of the performance of CMOS compatible optical devices are made based on current state-of-art optical technologies Electrical and optical interconnects are compared for various design criteria based on these predictions The critical dimensions beyond which optical interconnect becomes advantageous over electrical interconnect are shown to be approximately one tenth of the chip edge length at the 22 nm technology node

188 citations


Cites background or methods from "Buffer delay change in the presence..."

  • ...The effect of the common mode noise, however, needs to be considered explicitly [ 15 ], tcom ¼� atr 2ð1 þ aÞVdd_ideal ðDVdd þ DVssÞ, (4)...

    [...]

  • ...The delay uncertainty caused by power/ground noise consists of two components, one component tdif is due to the differential mode noise (i.e., DVdd � DVss), which affects the delay by changing the effective driver resistance; the other component tcom is due to the common mode noise (i.e., DVdd þ DVss), which affects the delay by changing the effective switching threshold [ 15 ]....

    [...]

  • ...Since signal delay is directly related to the power and ground voltage levels, a specific model [ 15 ] is used to analyze the effect of power/ground noise on the delay....

    [...]

  • ...In this paper, the 50% delay is based on the effective power/ground voltage levels (rather than the ideal power/ground), which corresponds to the second delay definition among the four types of buffer delays described in [ 15 ]....

    [...]

Journal ArticleDOI
10 Dec 2002
TL;DR: In this article, the impact of power-supply noise on the performance of high-frequency microprocessors is analyzed. But the authors focus on the average supply voltage during switching.
Abstract: This paper analyzes the impact of power-supply noise on the performance of high-frequency microprocessors. First, delay models that take this noise into account are proposed for device-dominated and interconnect-dominated timing paths. For typical circuits, it is shown that the peak of the noise is largely irrelevant and that the average supply voltage during switching is more important. It is then argued that global differential noise can potentially have a greater timing impact than common-mode noise. Finally, realistic values for the model parameters are measured on a 2.53-GHz Pentium4 microprocessor using a 130-nm technology. These values imply that the power-supply noise present on the system board reduces clock frequency by 6.7%. The model suggests that the frequency penalty associated with this power-supply noise will steadily increase and reach 7.6% for the 90-nm technology generation.

174 citations


Cites methods from "Buffer delay change in the presence..."

  • ...The derivation is based on the transistor equations proposed by Sakurai and Newton in [15], [16] and is much more general than the one proposed in [ 17 ]....

    [...]

Journal ArticleDOI
TL;DR: The primary focus of this paper is to discuss the modeling of jitter caused by power supply noise (PSN), named power supply induced jitter (PSIJ).
Abstract: The primary focus of this paper is to discuss the modeling of jitter caused by power supply noise (PSN), named power supply induced jitter (PSIJ). A holistic discussion is presented from the basics of power delivery networks to PSN and eventually to the modeling of PSIJ. The in-depth details and a review of several methodologies available in the literature for the estimation of PSIJ are presented.

45 citations


Cites background or methods from "Buffer delay change in the presence..."

  • ...As mentioned in Section II, the propagation delay of any circuit is affected by the power supply fluctuations [59]....

    [...]

  • ...Differential mode noise, common mode noise, and the loading effects are some of the factors that affect the propagation delay [59]....

    [...]

  • ...The change in buffer delays depending on the polarity of the power and ground noise ( VDD and VSS, respectively) is given as [59], [60]...

    [...]

Journal ArticleDOI
TL;DR: The transfer functions relating supply voltage fluctuations to jitter are analytically derived in closed form expressions for a single-ended buffer from a linear differential equation obtained from asymptotic linear inverter I-V curves.
Abstract: The transfer functions relating supply voltage fluctuations to jitter are analytically derived in closed form expressions for a single-ended buffer. The analytic transfer functions are derived from a linear differential equation obtained from asymptotic linear inverter I-V curves. The transfer functions are validated by comparison with HSPICE simulations. The estimated jitter is compared with the simulated jitter using eye diagrams with single-frequency and multitone supply voltage fluctuations.

44 citations

01 Jan 1997
TL;DR: A comprehensive analysis and estimate of simultaneous switching noise (SSN) including the velocity saturation effects seen in the submicron transistors during the switching of output drivers is presented and SPICE simulation results show that skewing buffer switching with additional inverter stages reduces SSN and increases buffer propagation delay.
Abstract: Complementary metal-oxide-semiconductor (CMOS) output buffers, comprised of a series of tapered inverters, are used to drive large off-chip capacitances. The ratio of the size of transistors between two consecutive stages is the buffer taper factor. With higher frequency of operation and simultaneous switching of the output drivers, the parasitic inductance present at the pin-pad-package interface results in significant switching noise on the power lines. A comprehensive analysis and estimate of simultaneous switching noise (SSN) including the velocity saturation effects seen in the submicron transistors during the switching of output drivers is presented. The effect of SSN on the overall buffer propagation delay and transition time is discussed. The presence of SSN results in an increase in the optimum taper factor between inverter stages for a given capacitive load. Beyond a critical value, the output transition time of a tapered buffer increases with reducing taper factor due to SSN. SSN can be reduced by skewing the switching of output buffers. SPICE simulation results show that skewing buffer switching with additional inverter stages reduces SSN and increases buffer propagation delay.

42 citations

References
More filters
Journal ArticleDOI
TL;DR: In this paper, an alpha-power-law MOS model that includes the carrier velocity saturation effect, which becomes prominent in short-channel MOSFETs, is introduced and closed-form expressions for the delay, short-circuit power, and transition voltage of CMOS inverters are derived.
Abstract: An alpha -power-law MOS model that includes the carrier velocity saturation effect, which becomes prominent in short-channel MOSFETs, is introduced. The model is an extension of Shockley's square-law MOS model in the saturation region. Since the model is simple, it can be used to handle MOSFET circuits analytically and can predict the circuit behavior in the submicrometer region. Using the model, closed-form expressions for the delay, short-circuit power, and transition voltage of CMOS inverters are derived. The delay expression includes input waveform slope effects and parasitic drain/source resistance effects and can be used in simulation and/or optimization CAD tools. It is found that the CMOS inverter delay becomes less sensitive to the input waveform slope and that short-circuit dissipation increases as the carrier velocity saturation effect in short-channel MOSFETs gets more severe. >

1,596 citations

Journal ArticleDOI
TL;DR: An improved timing model for CMOS combinational logic is presented, which yields a better understanding of the switching behavior of the CMOS inverter than the step-response model by considering the slope of the input waveform.
Abstract: An improved timing model for CMOS combinational logic is presented. The model is based on an analytical solution for the CMOS inverter output response to an input ramp. This model yields a better understanding of the switching behavior of the CMOS inverter than the step-response model by considering the slope of the input waveform. Essentially, the propagation delay is shown to be the sum of the step-response delay and an input dependent delay that may account for as much as 50-100 percent of the total delay. The matching between the ramp input and the characteristic input waveforms is shown to be easily performed for excellent agreement in output response and propagation delay. Even though the short-circuit current is neglected, its influence is shown to be small and may be corrected. As an example, the timing model is used to optimize CMOS output buffers for minimum delay. If the intrinsic output load capacitance is included in the model, the optimum tapering factor is shown to be not e but a value in the range 3-5 depending on process parameters and design style. Also, due to the input dependence of the propagation delay, the last inverter stage in the buffer should have a larger tapering factor than the other stages for minimum delay.

456 citations

Journal ArticleDOI
TL;DR: An extension of the effective capacitance equation is proposed that captures the complete waveform response accurately, with a two-piece gate-output-waveform approximation, for the "effective load capacitance" of a pc interconnect.
Abstract: With finer line widths and faster switching speeds, the resistance of on-chip metal interconnect is having a dominant impact on the timing behavior of logic gates. Specifically, the gates are switching faster and the interconnect delays are getting longer due to scaling. This results in a trend in which the RC interconnect delay is beginning to comprise a larger portion of the overall logic stage delay. This shift in relative delay dominance from the gate to the RC interconnect is increased by resistance shielding. That is, as the gate "resistance" gets smaller and the metal resistance gets larger, the gate no longer "sees" the total net capacitance and the gate delay may be significantly less than expected. This trend complicates the timing analysis of digital circuits, which relies upon simple, empirical gate delay equations for efficiency. In this paper, we develop an analytical expression for the "effective load capacitance" of a pc interconnect. In addition, when there is significant shielding, the response waveforms at the gate output may have a large exponential tail. We show that this waveform tail can strongly influence the delay of the RC interconnect. Therefore, we propose an extension of the effective capacitance equation that captures the complete waveform response accurately, with a two-piece gate-output-waveform approximation. >

347 citations

Journal ArticleDOI
TL;DR: This approach represents the first work on buffer insertion to incorporate signal slew into the delay model while guaranteeing optimality and efficiently computes the complete, optimal power-delay trade-off curve for added design flexibility.
Abstract: We present efficient, optimal algorithms for timing optimization by discrete wire sizing and buffer insertion. Our algorithms are able to minimize a cost function subject to given timing constraints; we focus on minimization of dynamic power dissipation, but the algorithm is also easily adaptable to, for example, area minimization. In addition, the algorithm efficiently computes the complete, optimal power-delay trade-off curve for added design flexibility. An extension of our basic algorithm accommodates a generalized delay model which takes into account the effect of signal slew on buffer delay which can contribute substantially to overall delay. To the best of our knowledge, our approach represents the first work on buffer insertion to incorporate signal slew into the delay model while guaranteeing optimality. The effectiveness of these methods is demonstrated experimentally.

255 citations

Journal ArticleDOI
TL;DR: In this paper, the authors derived analytical delay expressions for CMOS gates in the sub-micrometer region, and derived closed-form delay formulas for both inverters and series-connected MOSFET structures.
Abstract: In order to derive analytical delay expressions for CMOS gates in the submicrometer region, a realistic MOS model which incorporates an nth power law MOS model is developed. Closed-form delay formulas are obtained for CMOS inverters and series-connected MOSFET structures (SCMSs) that include short-channel effects. It is shown that the ratio of the delay of NAND/NOR to the delay of the inverter becomes smaller in the submicrometer region, because the V/sub DS/ and V/sub GS/ of each MOSFET in the SCMS are smaller than those of an inverter MOSFET. The smaller voltages in turn mitigate and relax the severe carrier velocity saturation in miniaturized MOSFETs. The results of the analysis for submicrometer VLSI designs show that if the maximum number of series-connected MOSFETs is considered to be five in 2- mu m designs, then the number can be increased to six or seven in the submicrometer circuit design. In typical cases in VLSI designs, the delay ratio for N-SCMS is much less than N/sup 2/. The delay dependence on input terminal position for SCMS structures is also described. >

243 citations


"Buffer delay change in the presence..." refers background in this paper

  • ...In deep submicron circuits the signal transitions are fast, so we can ignore the short circuit current [17][ 18 ]....

    [...]

Frequently Asked Questions (14)
Q1. What are the contributions in this paper?

In this paper, the authors analyze the effect of such noise on signal propagation through a buffer and present simple, closed-form formulas to estimate the corresponding change of delay. 

Deep submicron circuits have decreased power supply level Vdd and decreased velocity saturation index α, leading to increased sensitivity of delay to P/G noise. 

Switching of the buffers A, B and D has a symmetric effect on the power and ground noise ( and ), because they drive loads (consisting largely of parasitic interconnect capacitance) within the same block. 

Because of the buffer/repeaters’ preponderance in number, use in heavily loaded nets, and use in clock and timing circuits, buffer delays account for a large percentage of all critical timing nets in a design. 

Growth in design sizes and scaling of interconnections have led to the requirement for insertion of very large numbers of buffer/repeaters in recent designs [1][7][9][15]. 

At the same time, scaling of power supply levels and improving transconductance of devices have increased the sensitivity of buffers to supplylevel-induced delays. 

The range for the power and ground noise is from -20% (-0.36 volt) to 20% (0.36 volt) of the power supply voltage, which is set to 1.8 volt for the selected technology. 

On the other hand, switching of the C and E buffers has a non-symmetric effect on and , because they drive loads which are outside of block 1, causing different switching currents to flow through the power and ground ports of block 1. 

Variations of power and ground levels play an important role because this type of noise significantly degrades circuit performance. 

Without loss of generality, the authors define a buffer’s ideal delay as the time interval between its input and output voltage reaching 50% of the power supply level. 

Using the α-power law MOSFET model, the authors derived general formulas to estimate the influence of power and ground noise on delay and slope. 

This noise contributes to the local buffer/inverter delay in a complex way which can either increase or decrease the signal delay. 

despite the reduction of noise from lower-inductance packaging, the relative magnitude of the delay changes is still a serious potential problem. 

For a falling transition, the effective switching threshold of the gate rises, so that the threshold voltage level is reached earlier, and the effective gate delay decreases.