High-Speed System and Analog Input Output Design
High-Speed System and Analog Input Output Design
Tran
High-Speed
System and
Analog Input/
Output Design
Second Edition
High-Speed System and Analog Input/Output
Design
Thanh T. Tran
High-Speed System
and Analog Input/Output
Design
Second Edition
Thanh T. Tran
Rice University
Houston, TX, USA
© The Editor(s) (if applicable) and The Author(s), under exclusive license to
Springer Nature Switzerland AG 2010, 2023
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of
illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by
similar or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
To my wife, Nga
Preface
This book is the second edition of the High-Speed DSP and Analog System Design.
It covers the high-speed system and analog input/output design techniques and
highlights common pitfalls causing noise and electromagnetic interference problems
engineers have been facing for many years. The material in this book originated from
my high-speed DSP system design guide (Texas Instruments SPRU 889), my system
design courses at Rice University, and my experience in designing computers and
DSP systems for more than 30 years. The book provides hands-on, practical advice
for working engineers and electrical engineering students, including:
• Tips on cost-efficient design and system simulation that minimize late-stage
redesign costs and product shipment delays.
• Fifteen easily-accessible chapters in 271 pages.
• Emphasis on good high-speed and analog design practices that minimize both
component and system noise and ensure system design success, including indus-
try compliance checks.
• Guidelines to be used throughout the design process to reduce noise and radiation
and to avoid common pitfalls while improving quality and reliability.
• Hand-on design examples focusing on audio, video, analog filters, DDR memory,
USB 3.1, and power supplies.
The inclusion of analog systems and related issues cannot be found in other high-
speed design books.
This book is intended for practicing engineers and electrical engineering students
and is organized as follows:
• Chapter 1: Highlights challenges in designing video, audio, computer, and
communication systems.
• Chapter 2: Covers system design methodology, including pre-layout and post-
layout design and simulations.
• Chapter 3: Reviews fundamentals of Alternate Current (AC) and Direct Current
(DC).
vii
viii Preface
• Chapter 4: Covers analog active and passive filter design including operational
amplifier design with single-rail and dual-rail power supplies.
• Chapter 5: Presents an overview of data converter, sampling techniques and
quantization noise.
• Chapter 6: Covers transmission line theories and effects. Demonstrates different
signal termination schemes by performing signal integrity simulations and lab
measurements.
• Chapter 7: Covers transmission line effects in frequency domain. Reviews
Scattering Parameters or S-Parameters and shows how the models are being
used in high-speed digital systems.
• Chapter 8: Shows the effects of crosstalk and methods to reduce interference.
Highlights the importance of current return paths.
• Chapter 9: Provides memory sub-system design considerations, including DDR
overview, signal integrity, and design example. And how to use the tool to do
DDR compliance tests.
• Chapter 10: Provides a USB 3.1 design example using S-Parameter models and
simulation tool to run compliance checks.
• Chapter 11: Covers design considerations of analog phase-locked loop (APLL)
and digital phase-locked loop (DPLL) and how to isolate noise from affecting
APLL and DPLL jitter.
• Chapter 12: Provides an overview of switching and linear power supplies and
highlights the importance of having proper power sequencing schemes and power
supply decoupling.
• Chapter 13: Covers the analytical and general power supply decoupling tech-
niques, AC and DC resistance, and input/output filtering techniques.
• Chapter 14: Covers printed circuit board (PCB) stackup and signal routing
considerations.
• Chapter 15: Describes sources of electromagnetic interference (EMI) and how to
mitigate them.
ix
Contents
xi
xii Contents
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
About the Author
xv
xvi About the Author
As system performance levels and clock frequencies continue to rise at a rapid rate,
managing noise, radiation, and power consumption becomes an increasingly impor-
tant issue. At high frequencies, the traces on a PCB carrying signals act as trans-
mission lines and antennas that can generate signal reflections and radiations that
cause distortion and create challenges in achieving timing requirements, industry
interconnect protocol compliance, and electromagnetic compatibility (EMC) com-
pliance. These can often make it difficult to meet Federal Communication Commis-
sion (FCC) Class A and Class B [1] requirements. Heat sinks and venting that may
be required to address the thermal challenges of high-performance designs can
further exacerbate EMC problems. Many systems today have embedded wireless
local area network (WLAN) and Bluetooth, which will create further difficulties as
intentional radiators are designed into the system.
With these difficulties, it is necessary to rethink the traditional high-speed system
design process. In the traditional approach, engineers focus on the functional and
performance aspects of the design. Noise and radiation are considered only toward
the later stages of the design process if prototype testing reveals problems. But today,
noise problems are becoming increasingly common and more than 70% of new designs
fail first-time EMC testing. As a result, it is essential to begin addressing these issues
from the very beginning of the design process. By investing a small amount of time in
the use of low-noise and low-radiation design methods, and simulations at the begin-
ning of the development cycle, this will generate a much more cost-efficient design by
minimizing late-stage redesign costs and delays in the product ship date.
Typical systems such as the ones shown in Figs. 1.1 and 1.2 consist of many external
to CPU or SoC devices such as Chipset, audio CODEC, video, LCD display,
wireless communication (Bluetooth, GPS, UWB, and IEEE 802.11), Ethernet
DDR
RGB888/
YUV422
HDMI
HD Camera VIDEO
Receive IN RGB888 HD
VIDEO
OUT
Panel
1000
GMII
BaseT SPI
Ethernet Touch-
PHY Screen
1000 HDMI
Ethernet BaseT
GMII
DSP/SoC HD
PHY Display
EMIF
NOR or NAND USB Port1
FLASH USB Port2
Mic Array McASPs
AUDIO COMM SDIO
Audio
IN/OUT IN/OUT WLAN
CODECs
UART
GPS
PCI
WLAN/ PCI SPI
Serial
UWB
ROM
CORE
POWER
Power SUPPLY
PLL OSC
IO/DDR
controller, USB, power supply, oscillators, storage, memory, and other supporting
circuitries. Each of these components can either be a noise generator or be affected
by interferences generated by neighboring components. Therefore, applying good
high-speed design practices and simulating high-speed ports are necessary to min-
imize both component and system-related noise and to ensure interconnect design
compatibility. Multi-gigahertz serial interconnects are increasingly becoming more
popular in high-speed systems for connecting one device to another as shown in
Fig. 1.1. Designing these multi-gigahertz interconnects requires digital engineers to
learn and apply techniques commonly being done by RF and microwave engineers,
for example, scattering parameters or s-parameters.
The coupling between a noise source and noise victim causes electrical noise.
Figure 1.3 shows a typical noise path. The noise source is typically a fast-switching
signal, and the noise victim is the component carrying the signal. The noise victim’s
performance will be impacted by the noise. Coupling takes place through the
parasitic capacitances and mutual inductances of the adjacent signals and circuits.
Electromagnetic coupling occurs when the signal traces become effective antennas,
which radiate and generate interferences to the adjacent circuitries.
There are many mechanisms by which noise can be generated in an electronic
system. External and internal CPU/DSP clock circuits generally have the highest
toggle rates and the primary source of high-frequency noise. Improperly terminated
signal lines may generate reflections and signal distortions. Also, improper signal
routing, grounding, and power supply decoupling may generate significant ground
noise, crosstalk, and oscillations.
Noise can also be generated within semiconductors [2] themselves:
• Thermal noise: Also known as Johnson noise is present in all resistors and is
caused by the random thermal motion of electrons. Thermal noise can be mini-
mized in audio and video designs by keeping resistance as low as possible to
improve the signal-to-noise ratio.
• Shot noise: Shot noise is caused by charges moving randomly across the gate in
diodes and transistors. This noise is inversely proportional to the DC current
flowing through the diode or transistor, so the higher DC operating current
increases the signal-to-noise ratio. Shot noise can become an important factor
when the DSP system includes many analog discrete devices on the signal paths,
for example, discrete video and audio amplifiers.
• Flicker noise: Also known as 1/f noise is present in all active devices. It is caused
by traps where charge barriers are captured and released randomly, causing
4 1 Challenges in High-Speed Systems Design
Audio systems represent one of the greatest challenges for high-speed DSP design.
Relatively small levels of noise often have a noticeable impact on the performance of
the finished product. In audio capture and playback, audio performance depends on
the quality of the audio CODEC being used, the power supply noise, the audio
circuit board layout, and the amount of crosstalk between the neighboring circuitries.
Also, the sampling clock must be stable to prevent unwanted sounds such as pops
and clicks during playback and capture. Figure 1.4 shows a typical signal chain of
the DSP audio design. Most DSPs include a Multi-Channel Buffered Serial Port or
McBSP [3] for interfacing with external audio CODECs. Although this is a propri-
etary interface, it is configurable to work with the industry standard I2S audio
CODECs.
1.3 Challenges in Video System Design 5
MIC IN AUDIO
ADC DSP DAC SPEAKERS
CD/DVD IN AMP
All the blocks shown in Fig. 1.4 from the ADC to the Amp stage are very
sensitive to noise so any interference coupled to any of the blocks will propagate
and generate unwanted audible sounds. Common audio design problems include:
• Noise coupled to the microphone input: Mic input typically has a very high gain
(+20 dB) so a small amount of noise can generate audible sounds.
• Not having an anti-aliasing filter at the audio inputs.
• Excessive distortion due to gain stage and amplitude mismatch.
• Excessive jitter on audio clocks, bit clock, and master clock.
• Lack of good decoupling and noise isolation techniques.
• Not using a linear regulator with high power supply rejection to isolate noise from
the audio CODEC.
• Not having good decoupling capacitors on the reference voltage used for ADC
and DAC converters.
• Switching power supply noise coupled to the audio circuits.
• High impedance audio traces are adjacent to noisy switching circuits and no
shortest current return path is provided in the printed circuit board (PCB) layout
to minimize the current return loop between the DSP and the CODEC.
• Not having isolated analog and digital grounds.
In summary, having good audio performance requires proper design of the ADCs,
DACs, DSP interfaces, clocks, input/output filters, power supplies, and the output
amplifier circuits. The performance of all these circuits not only depends on how
well the circuits are being designed but also on how the grounds and power being
isolated and the PCB traces being routed.
DDR4 Mem
HD IN ADC DAC
HDTV
SD IN (HD & SD) (HD & SD)
HDMI IN
HDMI or HDMI or
HD VIDEO
DVI or DP DVI or DP MONITOR
DVI or SoC or DSP
Receive Transmit
DP IN
TUNER
media interface (HDMI), digital video interface (DVI), and DisplayPort (DP), are
also highly sensitive to system noise as noise causes jitter which increases the bit
error rate (BER). As in any electronic system, it is not possible to eliminate the noise
totally but applying good design techniques will reduce the risk of it having a
negative impact on performance.
In any HD video system, there are many wide high-speed buses switching at a
rate of 66 MHz or higher, and these buses generate broadband noise and harmonics
that cause radiations in the Gigahertz range. This type of interference is difficult to
control because there are so many of these busses on the board and it is not practical
to terminate every signal trace being routed from one point to another. The good
news is that there are good design practices to follow in order to minimize
interference.
Like video and audio systems, communication is another important DSP application
that is highly sensitive to noise and radiation. One of many challenges here is
creating systems with multiple powerful and highly integrated DSPs that deliver
high performance with very low bit error rate and interference. In these systems,
interference not only generates EMI problems but also jams other communication
channels and causes false channel detection. These issues can be minimized by
applying proper board design techniques, shielding, RF, and mixed analog/digital
signals isolation. In some cases, a spread spectrum clock generator may be required
to further reduce the interference and to improve the signal-to-noise ratio. Although
spread spectrum clock reduces the peak level radiation, the harmonics of this clock
are spreading over a wider bandwidth, and this can cause inter-channel interference
so engineers must be careful when using this type of clock generator circuit.
Table 1.1 shows high-speed buses generating harmonics that interfering with embed-
ded Wireless Local Area Networks (WLAN). One example of the communication
Antenna
2.4GHz
or
5GHz
RADIO
Antenna
WLAN
(IEEE
802.11a/ BLUETOOTH
b/g/n DDR4 MEM
PCI
EXPRESS HD SoC or DSP DISPLAY
systems is shown in Fig. 1.6 where both Bluetooth and IEEE 802.11 are being
implemented on the same motherboard and residing on the same 2.4 GHz RF
spectrum. The most difficult tasks are how to prevent the two systems from inter-
fering with each other and how to prevent radiations from the high-speed busses
(PCI Express, DDR4, and display) interfering with the embedded antennas. By
applying the rules outlined in this book, engineers will improve and increase the
probability of design success.
1.6 Summary
This chapter highlights many challenges engineers face today. The good news is that
many of the issues can be mitigated by applying good design practices and using
state-of-the-art tools to simulate the circuits as described in this book.
The chapter arrangement of the book follows the input to output signal flow as
shown in Fig. 1.8 where a number in the circle denotes the chapter number of the
book, and the orange paths are high-speed digital signal paths while black paths are
analog input/output signals. On the soft copy of the book, clicking on the circle
advances to the chapter labeled in the circle.
10 1 Challenges in High-Speed Systems Design
References
In this step, it is critical to define the optimized floorplan of the layout. Separating
analog and digital sections is a must, including using high-speed and low-speed
current loop methods to determine the best placement of the components.
To prepare for the PCB layout, here is a list of recommended tasks.
• Component placement—separating analog and digital components and placing
the components in such a way that minimizes the routing lengths.
• Determining high-speed signal paths, especially signals running at 1 Gbps or
higher. A good rule is to keep 2 vias per trace maximum, and the via must be
A routing document should be created for layout engineer to follow. This document
includes special routing guides generated by system designers.
At the completion of the PCB layout, a post-layout analysis starts by extracting PCB
parasitics to use in the simulations to determine if the layout causes any changes in
2.4 Spice Simulation of Critical Circuits 13
the circuit or system design performance. Tuning the circuit after layout typically is
required to meet the design targets because PCB parasitics are almost always
affecting high-speed traces.
A tool such as Mentor Graphics HyperLynx can extract parasitics accurately and
generate an output in spice format or s-parameter models to use in the simulator. A
detailed explanation of s-parameters will be covered in the later chapter of this book.
Example 2.1 PCB Extraction The layout in Fig. 2.2 only contains four 50-Ω
transmission lines with and without vias. The file is imported to HyperLynx, and
HyperLynx generates spice and s-parameter outputs as shown below (experimental
layout).
The picture in Fig. 2.3 shows TLine2 being selected for extraction and the spice
file of the transmission line and the via. This spice file can be incorporated into a
system for simulating circuits with parasitics.
The picture in Fig. 2.4 shows an s-parameter model created by HyperLynx, and
this model can be included in the system to simulate the design with parasitics.
s-Parameter is commonly being used as this is an industry standard, Touchstone
format, which is supported by many spice-based tools. Details of s-parameter model
will be covered in the later chapter.
The performance of critical circuits like RF filters and high-speed SERDES channels
is often affected by PCB layout, so it is particularly important for designers to extract
parasitics and include them in the system simulation. One example is to use HSPICE
to do critical timing analysis or to plot frequency response of an RF filter.
Figure 2.5 shows a proposed methodology for designing an RF filter.
Figure 2.6 is a circuit synthesized by free online software, Marki Microwave
[4]. The circuit is a fifth-order bandpass filter with cutoff frequencies at 700 MHz
and 1.1 GHz.
The picture below shows a TINA (TI Analog Circuit Simulator) simulation of the
bandpass filter. This confirms the corner frequencies of the filter without any PCB
parasitics (Fig. 2.7).
Now, add PCB parasitics (extracted s-parameter model) to the bandpass filter and
do an HSPICE simulation to check the filter’s performance. The simulation results
with parasitics as shown in the picture below show that the corner frequencies of the
filter shift from 700 to 520 MHz on the low end and from 1.1 GHz to 750 MHz on
2.4 Spice Simulation of Critical Circuits 15
the high end. This demonstrates that the PCB parasitics can significantly change the
performance of the filter. In this case, tuning is required to compensate for the PCB
layout and to achieve the filter specifications (Fig. 2.8).
The Power Integrity and Signal Integrity tasks will be covered in detail in the later
chapters.
References 17
2.5 Summary
It is important for designers to use the right tool for the task. For example,
HyperLynx is a useful tool for PCB extractions and traditional digital signal integrity
simulations, while HSPICE is an excellent tool for critical timing analysis, including
simulating s-parameters. Another excellent tool for extracting layout is ANSYS
SiWave.
In summary, designers should develop their own methodology using the tools
available for them, and here is a list of tasks to do.
Perform circuit or logic simulations to validate functionalities of the design before
doing the actual PCB layout.
Complete floorplanning, including analog and digital components placement and
separation.
Do pre-layout simulations to determine critical paths, maximum number of vias
allowed, and routing guidelines.
For post-layout simulations, be sure to check critical circuits to make sure to
compensate for parasitics if necessary.
References
1
Zc2 ¼ , ð3:1Þ
2πfC2
R3
Amplitude ¼ Vac1, ð3:2Þ
R3 þ R4
assuming C2 is very large and the impedance of C2 is very small at all frequencies as
compared to R3 and R4. For AC signals, V1 5 V source acts as an AC ground.
Example 3.1 Using TI TINA Analog Simulator [1]
For C2 ¼ 100 μF, R3 ¼ 1 kΩ, and R4 ¼ 1 kΩ, V1 ¼ 5 VDC and Vac1 ¼ 2 V peak-to-
peak at 1 kHz,
1 1
Zc2 ¼ ¼ ¼ 1:6 Ω,
2πfC 2π ð1000Þð100e 6Þ
R3 1K
Vout1 ¼ Vac1 ¼ ð2VppÞ ¼ 1Vpp:
R3 þ R4 1K þ 1K
Waveforms of spice simulations of circuit in Fig. 3.1 are shown in Fig. 3.2. As
shown in the calculations, the output voltage is 1Vpp with a DC bias voltage of 5 V.
Rules to Remember
• DC sources are AC ground. Connect DC sources to ground when doing AC
analysis.
• AC signals are signals of interest, while DC voltages are generally for biasing.
• Always keep DC sources as stable as possible, power supply decoupling capac-
itors are needed.
• AC-coupled signals have no DC component unless they are biased again on the
other side of the AC-coupling capacitor.
The trend in the industry is moving toward serial busses, incorporating
AC-coupled methods to isolate DC voltages. This way DC voltage of one circuit
Vac1
V1 5V
does not affect another circuit. Some of the AC-coupled methods used in many
industry standards are HDMI, DisplayPort, PCIe, Analog Audio Inputs/Outputs.
These busses enable users to connect devices together and start operating with
almost no user interactions required. In some cases, a system may require users to
enter security information before enabling access.
These ease-of-use interconnections are possible because of intelligent software
being able to identify devices and of solid hardware design that allows circuits to
interact. Software identification methods are beyond the scope of this book. For
hardware, the key enabler is to isolate DC voltages from one system to another by
using AC-coupling technique discussed in the previous section.
In Fig. 3.3, the AC-coupling capacitors in series with the transmit and receive
channels prevent DC voltages of one system affecting the other, and only allow AC
signals to propagate. Without the AC-coupling capacitors, the transmitted signals are
22 3 AC Versus DC
Figure 3.6 shows why having an AC-coupling capacitor placed in the middle of
the two inverters does not function properly. The RC circuit in time domain digital
logic acts as a differentiator which differentiates the input waveform and generates
positive and negative spikes. The amplitude and width of spikes shown in Fig. 3.6
depend on the rise and fall times of the digital signal. This causes a logic failure
because the output of the second inverter is not equal to the input signal when the
input signal is inverted twice.
– Mute the amplifier output by using a circuit that is not tied to power supply
ramping up and down. For example, incorporating a similar circuit in [2].
– Implement a bridge-tied load as shown in Fig. 3.8. In this case, no
AC-coupling capacitor is needed because both inverting and non-inverting
outputs of amplifier are equal, no DC current flowing through the speaker to
generate noise spikes.
• Using a wrong AC-coupling capacitor value is one of the common pitfalls. To
figure the minimum value of the AC-coupling capacitor, it is necessary to find the
input impedance of the circuit being driven by the previous stage and the
minimum operating frequency.
In Fig. 3.8 example, to find the capacitor C1 value for the AMP 1 input
impedance of 20 kΩ and the minimum operating frequency of 20 Hz, here is
the calculation.
1 1
C1 ¼ ¼ ¼ 0:4 μF,
2πfZin 2π ð20Þð20e þ 3Þ
where f is the 3 dB corner of the operating frequency and Zin in the input
impedance of AMP 1. More filter design details will be covered in Chap. 4.
• For DC-coupled circuit, such as the bridge-tied load in Fig. 3.8, the importance
here is to control the two DC ramp rates at the outputs of the inverting and
non-inverting amplifiers to be synchronous so that there is no DC current flowing
through the speaker during powering up and down, and normal operation.
• For digital designs, DC-coupled method is almost always being implemented. If
the power supplies of the two cascading circuits are different, then the solution is
to add a voltage translator to regenerate compatible voltage levels for the next
References 25
stage as shown in Fig. 3.9. Voltage Translators are widely available, and one
low-cost example of designing a low-speed Voltage Translator is using one
N-Channel MOS transistor, T3, as in Fig. 3.10 where the input, Vin, is switching
from 0 to 5 V, and the output, Vout, is going from 0 to 10 V.
References
This chapter presents passive and active filter topologies and design techniques,
including practical design examples and system simulations. In DSP systems, there
are analog filters required for signal conditioning and limiting the bandwidth before
sampling. To design these filters, designers need to be knowledgeable about oper-
ational amplifiers, DC biasing circuits, AC-coupling techniques, and traditional
passive components like inductors, capacitors, and resistors.
Table 4.1 shows the characteristic of passive and active filters and highlights the
advantages and disadvantages of each type.
Vin R1 Vout
C1 Rload
In summary, if gain is not required and source and load impedances are known,
then it is better to go with passive filters. Now, if impedance isolation and gain are
required, then active filters would be better.
The first-order passive filter can easily be realized by one resistor and one capacitor
as shown in Fig. 4.1.
Assuming Rload is much higher than R1, the 3 dB corner frequency for the filter
in Fig. 4.1 is
4.1 Anti-Aliasing Filters 29
1
f 3dB ¼ : ð4:1Þ
2πR1 C1
Example 4.1 An audio ADC has a bandwidth of 20 Hz–20 kHz and requires a first-
order anti-aliasing filter at its input. Design and simulate [2] this filter.
Answer: Let f3dB ¼ 20 kHz,
1
20 kHz ¼ :
2πR1 C 1
Now, let C1 ¼ 0.001 μF and solve for R1. R1 ¼ 8KΩ. As shown in Fig. 4.2, the
3 dB corner frequency is at 20 kHz.
The second-order lowpass circuit requires one inductor and one capacitor. In gen-
eral, the order of the filter circuit is equal to the number of capacitors and inductors in
the circuit. As mentioned in the previous section, the passive filter depends on the
source and load impedances, so let us assume that the source impedance, RS, is much
higher than the load impedance, RL, as shown in Fig. 4.3.
The frequency response of the second-order filter circuit has amplitude peaking at
the 3 dB corner and the amount of this peaking depends on the ratio of RL and RS.
30 4 Analog Filter Design
Vin Rs L1 Vout
C1 RL
C1 1n RL 7k
This peaking typically does not affect the circuit performance as long as the noise at
the corner frequency is very low, so it is crucial for designers to verify that the
frequency response and signal-to-noise over the passband are as expected. It is
common to fine-tune RL and RS to get the frequency response needed for the
application.
For RL RS, the 3 dB corner frequency for the circuit in Fig. 4.3 is
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
R þ RS
f 3dB ¼ pLffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi: ð4:2Þ
2π L1 C1 RL
Example 4.2 An audio ADC has a bandwidth of 20 Hz–20 kHz and requires a
second-order anti-aliasing filter at its input. Design and simulate [2] this filter.
Answer: Let f3dB ¼ 20 kHz, RS ¼ 70 Ω and RL ¼ 7K Ω.
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
7K þ 70
20 kHz ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi :
2π ð7KÞL1 C 1
For practical purposes, active filter designs in this section are focused only on first-
and second-order Butterworth, as these are the most popular topologies being
implemented in audio, video, and communication systems today. To be able to
design active filters, designers need to understand operational amplifier or op amp
and how to bias the op amp to get the maximum dynamic range.
An op amp circuit has three terminals, inverting input and non-inverting input and
output, as shown in Fig. 4.6 and has the following electrical characteristics:
• Very high impedance: For an ideal op amp, it is assumed infinite input imped-
ance. The input currents are very small and are negligible.
32 4 Analog Filter Design
+
+
VP, Non Inverting Input
V D ¼ V P V N, ð4:3Þ
V O ¼ aV D , ð4:4Þ
V O ¼ aðV P V N Þ: ð4:5Þ
As shown in Eq. (4.5), the open loop op amp works like a comparator where the
output voltage is the difference of the input multiplied by the open loop gain.
Op amp typically being implemented in a linear circuit such as filters works in a
closed loop system where the output has a feedback to the input to control the signal
or AC gain. In this case, the inverting and non-inverting inputs are equal; this is also
known as virtual ground. The gain equations are derived as follows.
Non-Inverting Amplifier
I2 ¼ IN þ I1, ð4:6Þ
where IN is the negative input current and is equal to zero since op amp has infinite
input impedance (Fig. 4.7).
4.1 Anti-Aliasing Filters 33
R1
- VO
I1 VN
+
+
VP
U1 LM318
+
VIN
VO VN
I2 ¼ ,
R2
VN
I1 ¼ ,
R1
VO VN VN
¼ : ð4:7Þ
R2 R1
V O R2
Gain ¼ ¼ þ 1: ð4:8Þ
V IN R1
Inverting Amplifier
I2 þ I1 ¼ IN, ð4:9Þ
where IN is the negative input current and is equal to zero since op amp has infinite
input impedance (Fig. 4.8).
VO VN
I2 ¼ ,
R2
V IN V N
I1 ¼ ,
R1
34 4 Analog Filter Design
R2
I2
R1
- VO
I1 VN
+
+
+
VIN
U1 LM318
V O V N V IN V N
þ ¼ 0: ð4:10Þ
R2 R1
VO R
Gain ¼ ¼ 2: ð4:11Þ
V IN R1
Op amp can either be powered by a dual rail supply (VDD) or a single rail supply
(+VDD). For the dual rail supply, the signal is centered around zero volts and swings
between the positive and negative rails as shown in Fig. 4.9. In this case, the op amp
needs to be biased at zero volts as this allows the maximum symmetrical swing.
The rule is to always bias the positive terminal of the op amp as shown in
Figs. 4.10 and 4.12 and the bias voltage must be set at the level where the output
has the maximum swing as shown in Fig. 4.11.
For the op amp with a single rail power supply, the bias voltage must be half of
the power supply to guarantee maximum symmetrical swing as demonstrated in
Fig. 4.11. The circuit shown in Fig. 4.12 has a voltage divider formed by two equal
resistors, R3 and R4, to generate a bias voltage at half of the power supply rail,
+VDD. For an ideal op amp, the positive and negative input voltages are equal. But
this is not the case in the real world where there is always some small offset voltage
between the two inputs. This offset voltage is in the range of microvolts and can be
an issue for small signal detection and processing applications. In general, this offset
4.1 Anti-Aliasing Filters 35
+VDD
-VDD
R2
-VDD
VIN R1
- VO
VN
+
+ U1 LM318
+VDD
voltage is not a concern for video, audio, and communication designs, but it is good
to keep it as low as possible. To minimize the offset voltage, set the parallel
combination of R3 and R4 equal to R2.
R3 R4
R2 ¼ R3 ==R4 ¼ ,
R3 þ R4
and
R23
R3 ¼ R4 so R2 ¼ :
2R3
Therefore, R3 ¼ R4 ¼ 2R2 : ð4:12Þ
This bias voltage can also be generated by a voltage regulator. The advantage of
the regulator is that it rejects the power supply noise. But the disadvantages are
adding more cost and making the design and layout more complicated.
36 4 Analog Filter Design
+VDD
+VDD/2
0V
R2
VIN R1
- VO
VN
+
+ U1 LM318
R3 +VDD
+VDD/2
R4
Again, only the positive terminal of the op amp needs to be biased at half of the
power supply. Due to virtual ground rule, the negative DC voltage is at +VDD/2 and
this also sets the output at +VDD/2. Now the whole circuit is DC balanced, which
enables the signal to swing symmetrically around +VDD/2.
Another important rule to remember is that if a point in the circuit is connected to
a DC voltage, the connection point becomes an AC or signal ground. So for a resistor
divider circuit in Fig. 4.12, it is good to add a bypass capacitor C1 in parallel with R4
to provide a good AC ground as in Fig. 4.13. This capacitor does not affect the signal
4.1 Anti-Aliasing Filters 37
R2
VIN R1
- VO
VN
+
+ U1 LM318
R3 +VDD
+VDD/2
R4
C1
path at all, since the capacitor is on the positive input of the op amp which only has a
DC bias voltage.
For AC signals, C1 bypasses R4 and provides a very low impedance path to
ground. In this case, an RC filter is formed by R3 and C1 and the corner frequency is
1
f 3dB ¼ : ð4:13Þ
2πR3 C1
R1 R2
R3 ¼ R1 ==R2 ¼ , ð4:14Þ
R1 þ R2
where R1 and R2 determine the AC gain of the circuit as shown in Eq. (4.8).
Similarly, Fig. 4.15 shows a biasing circuit for the single rail supply non-inverting
amplifier. The resistors R3 and R4 bias the positive input at half of the supply voltage.
To minimize the offset voltage, set parallel combination of R3 and R4 equal to R2,
since DC current does not flow through R1 because of the DC blocking capacitor C3.
38 4 Analog Filter Design
R2
-VDD
R1
- VO
VN
VIN
+
+ U1 LM318
VP
R3
+VDD
R2
C3 R1
R4
- VO
VN
VIN
+
+ U1 LM318
VP
Rload
R3
VDD
R3 ==R4 ¼ R2 ,
R3 R4
¼ R2 : Since R3 ¼ R4 ,
R3 þ R4
R3 ¼ R4 ¼ 2R2 , ð4:15Þ
The next step is to isolate the DC bias voltages of the op amp to ensure that the input
and output loadings do not change the DC voltages of the circuit. This is done by
adding a DC blocking capacitor in series with the input and the output as shown in
Fig. 4.16. The DC blocking or AC-coupling capacitor affects the frequency response
of the circuit, so designers must perform the analysis to select the capacitor value
such that the corner frequency is outside of the band of interest.
From the input VIN looking into the circuit, the C1 capacitor and resistors R3 and
R4 in Fig. 4.17 form a high-pass filter and the corner frequency is
1
f 3dB ¼ ,
2π ðR2 ==R3 ÞC 1
R2 þ R3
f 3dB ¼ : ð4:16Þ
2π ðR2 R3 ÞC 1
R2
C3 R1
R4
- C2 VO
C1 VN
VIN
+
+ U1 LM318
VP
Rload
R3
VDD
f-3dB FREQUENCY
40 4 Analog Filter Design
The frequency response of the high-pass filter is shown in Fig. 4.17.The role of
this capacitor C3 is to block DC current from flowing through the resistor R1 to keep
all the DC voltages around the op amp equal. For AC signal, C3 and R1 also form a
high-pass filter and its 3 dB corner is
1
f 3dB ¼ : ð4:17Þ
2πR1 C3
For the output AC-coupling capacitor C2, looking out from the op amp output, the
capacitor C2 and the resistor Rload form a high-pass filter and its 3 dB corner
frequency is
1
f 3dB ¼ : ð4:18Þ
2πRload C 2
Again, C2 must be selected to ensure that the corner frequency is lower than the
lowest frequency of the band of interest.
Design Example 4.3 Design a non-inverting audio amplifier having the following
specifications:
• Gain ¼ 2
• Input 3 dB corner frequency ¼ 20 Hz
• Output 3 dB corner frequency ¼ 20 Hz
• From the input to output, there is a 6 dB attenuation at 20 Hz. This is due to
cascading 3 dB input and 3 dB output stages
• Input impedance is 10K Ω
• Output load impedance is 20K Ω
• AC-coupled input and output
• +12 V single rail power supply
Simulate [2] the circuit to verify the results.
Solution:
Use the circuit diagram in Fig. 4.16 and calculate all the component values.
V O R2
Gain ¼ ¼ þ 1 ¼ 2,
V IN R2
So, R2 ¼ R1.
Let R2 ¼ R1 ¼ 20K Ω, reasonable value for audio design.
From Eq. (4.15),
ð40KÞð40KÞ
¼ 20K Ω:
40K þ 40K
R2 þ R3
20 Hz ¼ ,
2π ðR2 R3 ÞC 1
40K þ 40K
C1 ¼ ¼ 0:4 μF:
2π ð20Þð40KÞð40KÞ
1
20 Hz ¼ ,
2πRload C2
1
C2 ¼ ¼ 0:4 μF:
2π ð20Þð20KÞ
1
20 Hz ¼ ,
2πR1 C3
1
C3 ¼ ¼ 0:4 μF:
2π ð20Þð20KÞ
R2 20k
C3 470n R1 20k
R4 40k
- C2 470n VO
C1 470n VN
VIN
+
+ U1 LM318
VP
Rload 20k
R3 40k
+
VG1
VDD 12
VO R2
¼ , ð4:19Þ
V IN R1 þ Z 2
1
where Z2 is the impedance of C2 and is equal to 2πfC and f is frequency.
2
Substitute Z2 into Eq. (4.19). The magnitude of the gain, ignoring the negative
sign as the sign only indicates the output is inverted, is
VO 2πfC 2 R2
¼ : ð4:20Þ
V IN 2πfC 2 R1 þ 1
4.1 Anti-Aliasing Filters 43
R2
VIN C2 R1
- C3 VO
VN
+
+ U1 LM318
R5
R3 +VDD
+VDD/2
R4
C1
1
f corner ¼ , ð4:21Þ
2πR1 C 2
where C2 and R1 are the input impedance looking into the circuit. Substitute the
corner frequency into Eq. (4.20) and the magnitude of the gain at this frequency
becomes
VO R
¼ 2 : ð4:22Þ
V IN 2R1
So, the AC-coupling capacitor at the input of the inverting amplifier forms a high-
pass filter where the corner frequency shown in Eq. (4.21) is at 6 dB as demon-
strated in Fig. 4.21. As frequency gets higher and higher, the gain in Eq. (4.20) is
dominated by the resistors R2 and R1, which lead to the same response as the
inverting amplifier without the AC-coupling capacitor. Again, designers need to
select C2 so that the corner frequency is lower than the lowest frequency of the
signal.
C3 in Fig. 4.20 is calculated the same way as in the non-inverting case.
44 4 Analog Filter Design
-6dB
fcorner FREQUENCY
Design Example 4.4 Design an inverting audio amplifier having the following
specifications:
• Gain ¼ 2
• Input 3 dB corner frequency ¼ 20 Hz
• Output 3 dB corner frequency ¼ 20 Hz
• From the input to output, there is a 9 dB attenuation at 20 Hz. This is due to
cascading 6 dB input and 3 dB output stages
• Input impedance is greater than 10K Ω
• Output load impedance is 20K Ω
• AC-coupled input and output
• +12 V single rail power supply
Simulate [2] the circuit to verify the results.
Solution:
Use the circuit diagram in Fig. 4.20 and calculate all the component values.
V O R2
Gain ¼ ¼ ¼ 2 ðneglect minus sign as it only indicates the phaseÞ,
V IN R1
So, R2 ¼ 2R1.
Let R1 ¼ 20K so R2 ¼ 2(20K) ¼ 40K.
From Eq. (4.12),
The input impedance is equal to R1 which is 20K since the positive and negative
inputs of the op amp are equal, AC ground.
From Eq. (4.21), the input corner frequency (6 dB) is
1
20 Hz ¼ ,
2πR1 C2
1
C2 ¼ ¼ 0:2 μF:
2π ð20Þð40KÞ
1
20 Hz ¼ ,
2πR5 C3
1
C3 ¼ ¼ 0:4 μF:
2π ð20Þð20K Þ
1
f 3dB ¼ ¼ 20 Hz,
2πR3 C 1
1
C1 ¼ ¼ 0:1 μF:
2π ð20Þð80KÞ
The final circuit is shown in Fig. 4.22 and simulation in Fig. 4.23.
The simulation in Fig. 4.23 verified that there is a 9 dB attenuation at 20 Hz; the
overall circuit has a gain of 2 or +6 dB and, at 20 Hz, the gain is +6 dB 9 dB
attenuation. Therefore, the graph in Fig. 4.23 shows roughly a 3 dB gain at 20 Hz.
R2 40k
VG1
R5 20k
R3 80k
V1 12
C1 100n
R4 80k
Let us assume that the active filters are running on a single rail power supply. This is
more common in todays electronics as it is more expensive to implement a design
powered by a dual rail power supply.
There are two topologies for the first-order lowpass filter design, inverting
lowpass and non-inverting lowpass. For the non-inverting topology, the gain must
be greater than 1 as shown in the previous section, and, for the inverting topology,
the output is 180 phase shifted from the input.
To create an inverting lowpass filter, simply take the circuit in Fig. 4.22 and add a
capacitor in parallel with the resistor R2. The new circuit shown in Fig. 4.24 has the
upper 3 dB corner of
1
f 3dB ¼ : ð4:23Þ
2πR2 C4
All other design parameters and methodologies are the same as those shown in the
previous op amp design sections.
Design Example 4.5 Design a first-order inverting audio filter having the same
specifications as in Design Example 4.4 but adding an upper frequency limitation at
20 kHz. Simulate the circuit to verify the results.
4.1 Anti-Aliasing Filters 47
C4
R2
VIN C2 R1
- C3 VO
VN
+
+ U1 LM318
+
VG1
R5
R3
V1 12
C1
R4
Solution:
Use the circuit in Fig. 4.24. From Eq. (4.23), the upper 3 dB frequency is
1
20 kHz ¼ :
2πR2 C 4
1
C4 ¼ ¼ 200 pF:
2π ð40KÞð20 kHzÞ
The final circuit and simulation are shown in Figs. 4.25 and 4.26, respectively. It is
verified that the circuit frequency response has an upper frequency limitation at
20 kHz. This is the first-order filter circuit with the 3 dB frequency at 20 kHz and
the slope decaying at 20 dB per decade.
As mentioned earlier, the output of the inverting lowpass filter has a 180 phase
shift as compared to the input. Figure 4.27 shows the simulated gain of 2 and phase
relationship of the input and output waveforms.
Now to realize the non-inverting first-order lowpass filter circuit, take the
non-inverting amplifier circuit and add a capacitor in parallel to R2 to limit the op
amp bandwidth and an RC filter at the positive input of the op amp as shown in
Fig. 4.28. Op amp typically has a high gain bandwidth product and can go unstable if
the bandwidth is not limited to the operating frequency range. Doing stability
48 4 Analog Filter Design
C4 200p
R2 40k
VG1
R5 20k
R3 80k
V1 12
C1 100n
R4 80k
C4
R2
C3 R1
R4
- C2 VO
C1 R5 VN
VIN
+
+ U1 LM318
VP
Rload
R3
+
VG1 C5
VDD 12
analysis is beyond the scope of this book but designers can learn more in [3]. This
new circuit is shown in Fig. 4.28 and the 3 dB corner is dominated by the resistor
R5 and the capacitor C5,
1
f 3dB ¼ : ð4:24Þ
2πR5 C5
With C4 connected in parallel with resistor R2, the op amp bandwidth is limited to
the 3 dB corner set by this RC filter. The corner frequency is
50 4 Analog Filter Design
1
f 3dB ¼ : ð4:25Þ
2πR2 C4
Also, in Fig. 4.28, the filter resistor R5 needs to be one-tenth of the parallel
combination of the resistors R3 and R4. This is to minimize the effects of the voltage
divider formed by R5 and the parallel combination of R3 and R4.
Design Example 4.6 Design a first-order non-inverting audio filter having the same
specifications as in Design Example 4.3 but adding an upper frequency limitation at
20 kHz. Simulate the circuit to verify the results.
Solution:
Use the circuit in Fig. 4.28. From Eq. (4.24), the upper 3 dB frequency is
1
20 kHz ¼
2πR2 C 5 :
1 1 R3 R4
R5 ¼ ðR ==R4 Þ ¼ ¼ 2K,
10 3 10 ðR3 þ R4 Þ
1
C5 ¼ ¼ 0:00398 μF:
2π ð20KÞð2 kHzÞ
Now, since the maximum upper frequency is 20 kHz, let us limit the op amp to
40 kHz which is two times the signal bandwidth. This provides plenty of bandwidth
margins.
From Eq. (4.25),
1
40 kHz ¼ ,
2πR2 C4
1
C4 ¼ ¼ 200 pF:
2π ð20KÞð40 kHzÞ
The final circuit and simulation are shown in Figs. 4.29 and 4.30, respectively. It is
verified that the circuit frequency response has an upper frequency limitation at
20 kHz. This is the first-order non-inverting filter circuit with the 3 dB frequency of
20 kHz and a slope of 20 dB/decade.
Figure 4.31 shows the simulated gain of 2 and phase relationship of the input and
output waveforms. The simulation results are correct and correlated with the calcu-
lations very well.
4.1 Anti-Aliasing Filters 51
C4 200p
R2 20k
C3 470n R1 20k
R4 40k
- C2 470n VO
C1 470n R5 2k VN
VIN +
+ U1 LM318
VP
Rload 20k
R3 40k
+
VG1 C5 3.98n
VDD 12
The best and the most popular topology for a second-order lowpass design is Sallen-
Key circuit [4]. The three different Sallen-Key circuits are unity gain, gain of 2, and
gain of higher than 2. The higher the gain the more unstable the circuit becomes, so it
is best to keep the overall gain of 2 or less. If more gain is required for the design,
then use another op amp as a gain stage to boost the overall gain. This guarantees
good stability.
52 4 Analog Filter Design
C1
2Q
- VO
VIN R1 R2
+
+ U1 LM318
1 1
+
VIN C2 1/2Q
Figure 4.32 shows a second-order lowpass filter with a gain of 1 [4]. The resistor and
capacitor values are normalized to 1 and Q where Q determines the gain peaking or
GP at the corner frequency as shown in Fig. 4.33. The gain peaking or GP equation is
2Q2
GP ¼ 20 log 10 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi , for Q > 0:707: ð4:26Þ
4Q2 1
AMPLITUDE GP
40dB/Decade
Slope
fcorner FREQUENCY
C1
R4
1
1
1
- VO
VIN R1 R2 R3
+
+ U1 LM318
1 Q
+
VIN C2 1/Q
where Km is the new resistance and Rold is the normalized resistance shown in
Fig. 4.32.
1
C new ¼ C , ð4:28Þ
K f K m old
where Kf is the corner frequency in rad/s, Kf ¼ 2πf; f is the corner frequency in Hertz.
Since R3 and R4 are equal in Fig. 4.34 [4], the overall circuit gain is 1 plus the ratio of
resistor R4 over resistor R3 or 2 as shown in the non-inverting amplifier section. All
the values in the circuit are calculated by the same methods in the Sallen-Key Circuit
with Gain ¼ 1 section.
54 4 Analog Filter Design
C1
R4
1
2-1/Q
1
- VO
VIN R1 R2 R3
+
+ U1 LM318
1 1
+
VIN C2 1
R4 2 1=Q 1
1þ ¼1þ ¼3 : ð4:29Þ
R3 1 Q
All the values in the circuit are calculated by the same methods demonstrated in
the Sallen-Key Circuit with Gain ¼ 1 section.
Design Example 4.7 Design a non-inverting second-order Sallen-Key lowpass
filter having the following specifications:
• Gain ¼ 2
• Corner frequency ¼ 20 kHz
• Gain peaking ¼ 5 dB
• Output load impedance ¼ 20K Ω
• +12 V single rail power supply
• Use the same input and output coupling capacitors as in Design Example 4.3.
Simulate [2] the circuit to verify the results.
Solution:
For Gain ¼ 2, use the circuit in Fig. 4.34.
For corner frequency ¼ 20 kHz,
Let Km ¼ 2 104.
For gain peaking ¼ 4 dB,
4.1 Anti-Aliasing Filters 55
2Q2
4 dB ¼ 20 log 10 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ,
4Q2 1
The final circuit is shown in Fig. 4.36 where C3, C5, and C4 came from the Design
Example 4.3. R6 and R7 form a voltage divider to bias the op amp at half of the power
supply voltage. The parallel combination of the resistors R6 and R7 is selected to be
10 times larger than the total resistance of R1 and R2. This guarantees that the bias
resistors will not affect the overall gain.
The simulations in Figs. 4.36 and 4.37 show the following results:
• Gain peaking ¼ 4 dB
• Corner frequency ¼ 20 kHz
C1 398p
R4 20k
C5 470n
- C4 470n VO
VIN C3 470n R1 20k R2 30k R3 20k
+
+ U1 LM318
C2 265p R5 20k
+
VIN
R6 1M
V1 12
R7 1M
4.2 Summary
As demonstrated throughout this chapter, filter designs are complicated and require
doing thorough system analysis using a circuit simulator such as [2]. A filter
topology is selected based on the following criteria:
• Is the gain greater than 1? If yes, then an active filter is needed. If no, then either
an active or a passive filter can be selected.
• How much attenuation is needed to prevent aliasing for the ADC input and to
reject the sampling noise for the DAC output? This determines first, second, or
higher order filter topology.
• Is maintaining the input and output phase important? If yes, then only a
non-inverting filter can be used.
• For active filters, is the circuit running on a single or dual rail power supply?
Refer to the DC and AC Couple and Biasing Op Amp sections, and bias the
circuit to allow for a maximum symmetrical swing.
• For passive filters, is inductor shielding required? Inductors tend to radiate high-
frequency noise, so shielding may be required to contain the radiation. In
summary, DSP systems require analog filters to limit the signal bandwidth before
sampling, to reconstruct the analog signal from the DAC output, and to eliminate
analog noise modulated on the signal. These filters are critical, and proper design
techniques outlined in this chapter should be followed to achieve the performance
goals.
References
Figure 5.1 shows a typical DSP system where the input and output are analog and
data converters plus processing elements reside in the middle of the signal chain. The
theories and applications of the input Gain Stage, Anti-Aliasing Filter (ADC input),
and Reconstruction Filter (DAC output) are covered in Chap. 4. This chapter focuses
on the ADC and DAC blocks of the DSP Signal Chain in Fig. 5.1.
In general, a DSP system captures an analog input, amplifies the signal, band-
limits the signal for sampling, converts it to digital, processes the data in digital
domain, converts it back to analog, and filters the sampling noise to reconstruct the
analog signal.
The design goal is to maintain or improve the signal quality as it propagates
through all the blocks shown in Fig. 5.1. The question is why is it necessary to
convert the analog signal to digital and process it in digital domain? It is because:
• There are no linear and nonlinear distortions.
• Data compression is possible in digital domain. This is key for video, audio, and
communication as these systems have limited transmission bandwidth.
• It is easy to upgrade the system by replacing the software and or DSP algorithms.
MEM
DIGITAL
WORD IDEAL LINE
101
100
011
1 LSB
010
001
000
0 1 2 3 4 5 ANALOG
VALUE
An ADC converts the analog input to the digital output word by sampling and
comparing the analog level to the digital word. This sampling point is to decide what
digital code is equivalent to this analog value. For example, in Fig. 5.2, sampling
5.2 Analog-to-Digital Converter (ADC) 61
Table 5.1 ADC input and Digital code Analog range (V) Quantization error
output
000 0–0.5 1/2 LSB or 0.5 V
001 0.5–1.5 1 LSB or 1 V
010 1.5–2.5 1 LSB or 1 V
011 2.5–3.5 1 LSB or 1 V
100 3.5–4.5 1 LSB or 1 V
101 4.5–5.5 1 LSB or 1 V
110 5.5–6.5 1 LSB or 1 V
111 6.5–7.5 1 LSB or 1 V
ANALOG ANALOG-TO-
ANTI-ALIASING DIGITAL
INPUT DIGITAL
FILTER OUTPUT
CONVERTER
Sampling
Clock, fS
point 0 occurs at 1 V analog level translating to 001 digital output word, sampling
point 1 occurs at 2 V analog level translating to 010 digital output word, and so on.
In Fig. 5.2, 1 LSB is defined as one Least Significant Bit and at every sampling
point; the analog level can vary from 1/2 LSB from the center. This is known as a
quantization error. The LSB voltage, VLSB, is equal to
V ref
V LSB ¼ , ð5:1Þ
2N
where Vx is 12 V LSB V X 12 V LSB and b1 is the most significant bit and bN is the
least significant bit.
Figure 5.3 shows a practical ADC block diagram where the analog input must be
band-limited before being converted to digital word. This is because the Nyquist
sampling theory defined that the sampling clock must be at least two times the analog
bandwidth to prevent aliasing. Aliasing is the image of the analog signal folded back
into the frequency of interest; aliasing degrades the ADC performance. Therefore, an
anti-aliasing filter must be placed at the input of the ADC.
62 5 Data Converter Overview
AMPLITUDE
fA fs -fA fs fs +fA 2fs -fA 2fs 2fs +fA 3fs -fA 3fs 3fs+fA
FREQUENCY
fA fS FREQUENCY
Aliasing for fS < 2fA
5.2.1 Sampling
An ADC utilizes the sampling clock, fs, to sample the analog input and represents the
level in a digital word as shown in the previous section. Sampling is equivalent to
amplitude modulating the signal, fA, into a carrier equal to the sampling frequency
and generates a frequency spectrum shown in Fig. 5.4.
Nyquist Theorem states that the sampling frequency fs must be equal to or greater
than two times the signal bandwidth fA or
f s 2f A : ð5:3Þ
If the sampling frequency is less than two times the signal bandwidth, then the
sampled signal aliases back into the signal bandwidth causing the dynamic range
degradation. Figure 5.5 shows an aliasing region when this is the case. So, to
guarantee compliant to Nyquist Theorem, the ADC must have an anti-aliasing filter
at the input to limit the signal bandwidth and a sampling frequency greater than two
times the signal bandwidth. One important rule-of-thumb for selecting an anti-
aliasing filter is that higher sampling frequency ADC requires a lower order anti-
aliasing filter. This is because higher sampling frequency pushes the noise spectrum
further away from the analog spectrum and enables the lower order filter to prevent
5.2 Analog-to-Digital Converter (ADC) 63
AMPLITUDE
Lower Order Filter to Limit the
Analog Bandwidth for High fS
fA fS
No Aliasing for fS > 2fA FREQUENCY
the noise image from aliasing back into the band of interest. Figures 5.5 and 5.6
demonstrate this phenomenon.
Todays technologies enable very high sampling frequency ADC, so it is not
necessary to have higher than a second-order anti-aliasing filter at the input of
the ADC.
AMPLITUDE
ANALOG
SIGNAL
100
011
010
001
000
TIME
0 1 2 3 4 5
1/q
-p/2 +p/2
X
5.2 Analog-to-Digital Converter (ADC) 65
2 31=2
ð
q=2
6 7
V Q ðrmsÞ ¼ 4 x2 f Q ðxÞdx5
q=2
2 31=2 2 31=2
ð
q=2 ð
q=2
61 7 61 7
¼4 x2 dx5 ¼4 x2 dx5
q q ð5:4Þ
q=2 q=2
2 31=2
ð
q=2 h i 1=2
61 7 1 1 3 þq=2
¼4 x2 dx5 ¼ x
q q 3 q=2
q=2
q
¼ pffiffiffiffiffi :
12
2 31=2
ðT
1
V IN ðrmsÞ ¼ 4 V 2 ðt Þdt 5 ,
T
0
where V(t) ¼ Acos(2πfct) and A is the zero-to-peak voltage as shown in Fig. 5.9.
Therefore,
2 312
ðT
1
V IN ðrmsÞ ¼ 4 A2 cos 2 ð2πf c t Þdt 5 :
T
0
66 5 Data Converter Overview
VREF
0V
t
Since
1 þ cos ð4πf c t Þ
cos 2 ð2πf c t Þ ¼ ,
2
2 31=2
ðT 2 1=2
A2 1 þ cos ð4πf c t Þ 5 A T A
V IN ðrmsÞ ¼ 4 dt ¼ ¼ pffiffiffi : ð5:5Þ
T 2 T 2 2
0
V
V INðrmsÞ ¼ prefffiffiffi , ð5:6Þ
2 2
or
V ref ¼ V peak‐to‐peak ¼ 2:828V INðrmsÞ :
The signal-to-noise or SNR is defined as the log of the ratio of the input RMS
voltage over the quantization noise.
V IN ðrmsÞ
SNR ¼ 20 log 10 : ð5:7Þ
V Q ðrmsÞ
V ref
V LSB ¼ ,
2N
pffiffiffi
3
SNR ¼ 20 log 10 pffiffiffi 2N
2 ð5:8Þ
¼ 6:02N þ 1:76 dB:
Equation (5.8) indicates that the performance of an ADC depends on the number
of bits used to quantize the analog signal. It is roughly 6 dB/bit. This equation was
derived assuming that the only error in the system is quantization error. In the real
world, other factors such as power supply noise and clock jitter generate additional
errors that degrade the signal-to-noise significantly. So, another equation to measure
the overall performance of the ADC is
SNR 1:76
ENOB, Effective Number of Bits ¼ : ð5:9Þ
6:02
For example, if a 16-bit ADC has an SNR specification of 86 dB, the Effective
Number of Bits is
86 1:76
ENOB ¼ ¼ 14:
6:02
What this means is that the 16-bit ADC only performs at a 14-bit level due to
quantization noise and other system-related noise, such as power supplies, clocks,
and others degrading its performance.
As silicon technology advances to enable faster switching and higher data rate, the
digital waveforms also must change to be compatible with low voltage, low power
devices. The switching logic levels for IO (inputs and outputs) used to be 3.3 V, and
now these logic levels are as low as 1.2 V. Reducing the switching voltage signif-
icantly lowers the power consumption as defined in Eq. (5.1).
where C is the capacitive load, V is the switching voltage, and f is the switching
frequency. For digital logic, the input impedance of a gate is very high, so the timing
is mostly affected by the input gate capacitance. From Eq. (5.1), the most effective
way to reduce the power consumption of a system is to lower the switching voltage.
68 5 Data Converter Overview
For example, lowering the switching voltage by half reduces the dynamic power by
four. Many battery-powered portable systems apply voltage and frequency scaling
techniques to extend the battery life by lowering both the voltage and frequency
when the system is not active.
Traditionally, Fig. 5.11, digital signal transmission scheme is a single ended
scheme in which the timings are based on input high voltage (Vih), input low voltage
(Vil), rise time, fall time, and duty cycle (Pw/T ). For high-speed systems, the trend is
going to differential signal scheme as shown in Fig. 5.12 which has a positive going
signal and a negative going signal transmitted simultaneously.
The advantages of differential signaling over single ended signaling are:
• High or low state does not require the signal to switch from Vil to Vih. The signal
only has to go above or below the common voltage as shown in Fig. 5.11. This
significantly lowers the dynamic power as described.
• Higher common mode noise cancellation. The two signals within one differential
pair must be routed the exact same way, next to each other, one time or two times
5.2 Analog-to-Digital Converter (ADC) 69
the width of the trace. Since both signals are the same, the noise coupled to one is
similar to noise coupled to the other signal. This noise on both positive and
negative signals is referred as common mode noise. This common mode noise is
generally rejected by the receiver. The receiver is typically being implemented
using a differential to single ended amplifier, and its output is
V out ¼ Gain V inp V inn , ð5:11Þ
where Gain is the amplifier’s gain, Vinp is the positive input signal, and Vinn is
the negative input signal.
Let Vn be noise voltage coupled to both positive and negative signals in one
differential pair. In this case,
Equation (5.14) shows that the differential receiver cancels common mode
noise coupled to the transmitted signals. This is a major advantage for high-speed
signaling.
Another important concept in high-speed design is understanding the eye dia-
gram. Eye diagram is a composite signal which has important timing parameters
embedded in the diagram. For example, Fig. 5.13 shows an eye diagram of the
differential to single ended receiver output. Figure 5.14 eye diagram is being used to
show data signal distortion, noise, and signal attenuation due to transmission line
effects and intersymbol interference (ISI). ISI is the effect of adjacent pulses spilling
over the neighboring pulses.
A DAC converts the digital input codes to the analog output values and its transfer
function is shown in Fig. 5.16 [1]. Equation (5.15) shows the relationship between
the digital word, Bin, and the analog output signal, Vout.
V out ¼ V ref b1 21 þ b2 22 þ b3 23 þ . . . þ bN 2N ¼ V ref Bin , ð5:15Þ
where
Similar to ADC, for the 3-bit DAC and Vref of 8 V, the VLSB is equal to 1 V.
Table 5.2 shows an example of a 3-bit DAC taking a digital word and converting it
to an equivalent analog level assuming that the sampling error is 0.5 LSB.
Figure 5.17 shows a practical DAC block diagram where the digital input is being
converted to analog and the Reconstruction Filter eliminates the sampling noise
modulated on the analog waveform. The design of this filter depends on which DAC
is being used and how much noise suppression is necessary to achieve a certain
signal-to-noise specification. The filter topologies and design methodologies are
covered in Chap. 4.
72 5 Data Converter Overview
ANALOG
VALUE IDEAL LINE
1 LSB
0
DIGITAL
000 001 010 011 100 101
CODE
DIGITAL-TO-
RECONSTRUCTION ANALOG
ANALOG
FILTER OUTPUT
CONVERTER
Sampling
Clock, fS
The filter requirements for the DAC are analog bandwidth, fA, samples per second
input, fs, and stop band attenuation. As shown in Fig. 5.4, the sampled data has an
image closest to the band of interest at
5.4 Practical Data Converter Design Considerations 73
Table 5.2 DAC input and Digital input code Analog output (V) Error (V)
output
000 0 0.5
001 1 0.5
010 2 0.5
011 3 0.5
100 4 0.5
101 5 0.5
110 6 0.5
111 7 0.5
ANALOG ANALOG
ADC DSP DAC
INPUT OUTPUT
Sampling Sampling
Clock, fS Clock, fS
Image ¼ f s f A :
For example, a video signal has a bandwidth of 6 MHz, and the DAC input is
27 MSPS (Mega Samples Per Second). The image of this video signal is
A typical high-level block diagram shown in Fig. 5.18 consists of input ADC [4],
DSP, and DAC [5]. Here are the important parameters to consider when selecting
ADCs and DACs:
• Resolution and signal-to-noise (SNR)
• Input and output voltage range
• Sampling frequency
• Differential nonlinearity
• Integral nonlinearity
74 5 Data Converter Overview
ADC or DAC resolution determines the number of digital bits to represent an analog
signal. Higher resolution always yields higher SNR. Here is an SNR specification of
ADC [4].
For a 10-bit ADC, SNR ¼ 55 dB for fA ¼ 10 MHz and fs ¼ 110 MSPS.
Let us calculate the Effective Number of Bits or ENOB using Eq. (5.9).
ENOB indicates that the performance of this 10-bit ADC is equivalent to the
performance of an ideal 9-bit ADC. Achieving the theoretical resolution is challeng-
ing, so getting 9-bit performance out of a 10-bit ADC is considered a very
good ADC.
Sampling frequency determines the frequency spacing between the analog signal
and its images as shown in Fig. 5.4. So, a higher sampling frequency is preferred.
This allows lower order input and output filters. The disadvantage of higher sam-
pling frequency is that it tends to radiate more effectively as more traces on a PCB
become effective antennas at higher frequency. Refer to Chap. 15 for more details on
EMI.
In the ADC [4] data sheet, the SNR specification shows a sampling frequency of
110 MHz for a 10 MHz signal. The minimum sampling frequency is
The oversampling ratio of 5.5 indicates that the ADC samples the input signal at a
rate 5.5 times higher than the minimum sampling frequency required by the Nyquist
Theorem. In general, oversampling ADC or DAC has higher performance than the
non-oversampling one.
ADC and DAC datasheets specify a maximum input voltage and a minimum output
voltage, respectively. To maximize the performance of the system, it is
recommended to amplify or attenuate the signal to get the maximum symmetrical
5.4 Practical Data Converter Design Considerations 75
4V ANALOG Gain =
INPUT 0.5
R2
swing. For example, an ADC has two inputs, and each input has a maximum input
specification of 2 V peak-to-peak. The two outputs from the previous stages measure
1 and 4 V peak-to-peak. To balance the ADC inputs, one input path must include a
gain circuit and the other has an attenuator as shown in Fig. 5.19.
The gain stages of 2 and 0.5 can be implemented using op amps as shown in
Chap. 4. Another option to design a gain of 0.5 is using a resistor divider to divide
the input voltage by half. Here is an example.
For the voltage divider shown in Fig. 5.20, the output voltage is
R2
Vo ¼ V : ð5:16Þ
R1 þ R2 IN
The Differential Nonlinearity (DNL) error [6] occurs on both ADC and DAC and is
defined as the difference between the step width and one LSB. From Fig. 5.21 [6],
the ADC DNL errors occur at 1 and 4 V analog levels. Ideally, the 001 code is
centered at 1 V with 1/2 LSB tolerance and the 100 code is centered at 4 V with
1/2 LSB, but due to DNL errors, the 001 code is at 1 V with 1/2 LSB tolerance
and the 100 code is at 4 V with 1/2 LSB to +1 LSB tolerance.
76 5 Data Converter Overview
DIGITAL
WORD
OUTPUT
101
100
DNL = +1/2LSB
011
010
DNL = -1/2LSB
001
1 LSB
000
0 1 2 3 4 5 ANALOG
INPUT
Figure 5.22 [6] shows a DAC DNL error where the input 010 code has an analog
range larger than 1 LSB voltage. In this case, the error is +1/4 LSB.
The Integral Nonlinearly (INL) is defined as the maximum deviation of the actual
transfer function from an ideal straight line. Like DNL, both ADC and DAC have
INL errors. For ADC, the amount of error is measured at the transition from one
digital code to another and compared it to the ideal transition point as shown in
Fig. 5.23 [6]. For example, the actual transition from 001 code to 010 code happens
at the 1 V point instead of 1.5 V point. This early transition produces a 1/2 LSB
INL error.
5.5 Summary 77
ANALOG
OUTPUT
1 LSB
2
DNL=+1/4LSB
1
0
000 001 010 011 100 101 DIGITAL
INPUT
WORD
Figure 5.24 [6] shows an INL error caused by the DAC. In this case, the
maximum deviation from the ideal curve happens at the 011 input digital code and
the error is equal to 1/2 LSB.
5.5 Summary
DIGITAL
WORD
OUTPUT
101
Actual
Transition
100
INL = -1/4LSB
011
Ideal
Transition
010
001
000
0 1 2 3 4 5 ANALOG
INPUT
• DNL and INL errors occurred in ADC and DAC cannot be eliminated using
system design techniques. These are inherent in the data converter itself. Choose
the converters with the lowest DNL and INL specifications if possible.
• Overall, the design goal for the data converter system is to match the performance
specified in the converter datasheet. For example, if the converter has an 80 dB
SNR specification, the system performance target should be 80 dB SNR. This is
the best performance that can be achieved because all other components (anti-
aliasing filters, amplifiers, power supplies, and reconstruction filters) around the
converter tend to generate additional noise and errors in the system.
References 79
ANALOG
OUTPUT
INL=+1/2LSB (maximum
deviation from ideal)
3
0
000 001 010 011 100 101 DIGITAL
INPUT
WORD
References
1. D. Johns, K. Martin, Analog Integrated Circuit Design (John Wiley & Sons, New York, 1997)
2. Texas Instruments Inc., Interface Circuits for TIA/EIA-644 (LVDS) Design Notes, SLLA038B
(2002)
3. Mentor Graphics (HyperLynx VX2.7) HyperLynx Signal Integrity Simulation Software. http://
www.mentor.com/products/pcb-system-design/circuit-simulation/HyperLynx-signal-integrity/
4. Texas Instruments Inc., TVP7002 Triple 8-/10-Bit 165-/110-MSPS Video and Graphics Digitizer
With Horizontal PLL, SLES205A (2007)
5. Texas Instruments Inc., THS8200 All-Format Oversampled Component Video/PC Graphics D/A
System with Three 11-Bit DACs, CGMS Data Insertion, SLES032B (2009)
6. Texas Instruments Inc., Understanding Data Converters, Mixed-Signal Products, SLAA013
(1995)
Chapter 6
Transmission Line (TL) Effects
Transmission line (TL) effects are one of the most common causes of noise problems
in high-speed DSP systems. When do traces become TLs and how do TLs affect the
system performance? A rule-of-thumb is that traces become TLs when the signals on
those traces have a rise time (Tr) less than twice the propagation delay (Tp). For
example, if a delay from the source to the load is 2 nS, then any of the signals with a
rise time less than 4 nS becomes a TL. In this case, termination is required to
guarantee minimum overshoots and undershoots caused by reflections. Excessive
TL reflections can cause electromagnetic interference and random logic or DSP false
triggering. As a result of these effects, the design may fail to get the FCC certification
or to fully function under all operating conditions such as at high temperatures or
over-voltage conditions.
There are two types of transmission lines, lossless and lossy. The ideal lossless
transmission line has zero resistance while a lossy TL has some small series
resistance that distorts and attenuates the propagating signals. In practice, all TLs
are lossy. Modeling of lossy TLs is a difficult challenge that is beyond the scope of
this book. Since the focus of this book is only on practical problem-solving methods,
it assumes a lossless TL to keep things simple. This is a reasonable assumption
because in DSP systems where the operating frequency is less than 1 GHz the losses
on printed circuit board traces are negligible compared to losses in the entire signal
chain, from analog to digital and back to analog.
L1 L2 L3 L4
C1 C2 C3 C4
1
V p ¼ pffiffiffiffiffiffi , ð6:1Þ
LC
rffiffiffiffiffiffiffi
L
Zo ¼ , ð6:2Þ
C
where L is inductance per unit length and C is capacitance per unit length.
Another important property of the TL is the propagation delay, Td. Equation (6.3)
for Td is
1 pffiffiffiffiffiffi
Td ¼ ¼ LC: ð6:3Þ
Vp
The source and load TL reflections depend on how well the output impedance and
the load impedance, respectively, are matched with the characteristic impedance.
The load and source reflection coefficients, Eqs. (6.4) and (6.5), are
ZS ZO
ΓS ¼ source reflections ¼ , ð6:4Þ
ZS þ ZO
ZL ZO
ΓL ¼ load reflections ¼ , ð6:5Þ
ZL þ ZO
where ZS and ZL are the source impedance and load impedance, respectively.
The following example shows the characteristics of a TL with no load and with a
3 V signal source driving the line.
Example 6.1 Calculate the voltage at the open-ended load of the transmission line
below (Fig. 6.2).
ZO 50
V initial ¼ V clk ¼3 ¼ 2 V,
ZS þ ZO 25 þ 50
Z S Z O 25 50
ρS ¼ ¼ ¼ 0:333,
Z S þ Z O 25 þ 50
ZL ZO
ρL ¼ ¼ 1:
ZL þ ZO
6.1 Transmission Line Theory 83
Zs = 25 Td = 1.8nS, Zo = 50 ohms
Open
Vclk = 3.0V Transmission Line
Ended
1.8nS 2.0V
3.6nS 4.0V
5.4nS 3.3V
7.2nS 2.6V
9.0nS 2.8V
3.1V
In Fig. 6.3, the overshoot voltage can be calculated using a lattice diagram [1] as
follows (Fig. 6.4).
At T1 ¼ 1.8 nS: VL ¼ Vi + ρLVi ¼ 2 + 2 ¼ 4.0 V,
At T2 ¼ 3.6 nS: VS ¼ 4.0 V + ρS(ρLVi) ¼ 4.0 V 0.67 V ¼ 3.33 V,
At T3 ¼ 5.4 nS: VL ¼ 3.33 V + ρL(ρS(ρLVi)) ¼ 3.33 0.67 ¼ 2.66 V,
At T4 ¼ 7.2 nS: VS ¼ 2.66 V + ρS[ρL(ρS(ρLVi)] ¼ 2.66 + 0.22 ¼ 2.88 V,
At T5 ¼ 9.0 nS: VL ¼ 2.88 + ρL{ρS[ρL(ρS(ρLVi)]} ¼ 3.1 V.
As shown in Example 6.1, the reflections with a 3 V source caused the signal to
overshoot as high as 4 V at the load as explained below:
• The initial voltage level at the load at time T1 depends on the load impedance,
which is infinite for an open load, and the characteristic impedance of the TL.
• The voltage level at time T2, when the reflected signal arrives at the source,
depends on the source impedance and the characteristic impedance of the TL.
84 6 Transmission Line (TL) Effects
AMPLITUDE IN
VOLTS (V)
4V
3V
2V
1V
0V
T1 T2 T3 T4 T5 TIME
(1.8) (3.6) (5.4) (7.2) (9.0) (in nS)
• The voltage level at time T3, when the reflected signal arrives at the load again,
depends on the reflected voltage at T2 plus the reflected voltage at time T3.
• This process continues until a steady state is reached. In this example, the steady
state occurs at T5, which is 9 nS from T1, as shown in Fig. 6.3.
Figure 6.5 shows the waveforms at the load for both terminated and unterminated
circuits. As shown in the previous example, the terminated TL has a zero-reflection
coefficient and therefore no ringing occurs on the waveform as seen on the top graph
of Fig. 6.5. The problem is that in high-speed digital design, adding a 50-Ω resistor
to ground at the load is not practical because this requires the buffer to drive too
much current per line. In this case, the current would be 3.3 V/50 ¼ 66 mA. A
technique known as parallel termination can be used to overcome this problem. It
consists of adding a small capacitor in series with the resistor at the load to block
DC. The RC combination should be much less than the rise and fall times of the
signal propagating on the trace.
Figure 6.6 shows a parallel termination technique. This method can be used in the
application where one output drives multiple loads as long as the traces to the loads
called L2 are a lot shorter than the main trace L1.
To use the parallel termination technique, it is necessary to calculate the maxi-
mum allowable value for L2 according to Eq. (6.6) below assuming the main trace L1
and the rise time Tr are known.
Tr
L2 , max ¼ L110 : ð6:6Þ
6.2 Parallel Termination Simulations 85
Load 2
Load 3
Load 1 L2
Zs = 25 L1, Td = 1.8nS, Zo = 50 ohms
Load 4
Load 5
R
Transmission Line
C
Parallel termination techniques become useful when designers must use a single
clock output to drive multiple loads to minimize the clock skew between the loads.
In this case, having a series resistor at the source limits the drive current to the loads
and may cause timing violations by increasing rise times and fall times. This
simulation example includes one 6 in. trace (L1) and two 2 in. stubs. The DSP [2]
drives the main L1 trace and one memory device connected to each end of the 2 in.
trace. It is reasonable to neglect the effects of the stubs provided they are short and
meet the criteria shown in Fig. 6.7. In this case, only one parallel termination (68 Ω
and 10 pF) is required at the split of the main trace to the loads. Referring to the
simulation result in Fig. 6.8, the waveforms at the loads look good and meet all the
timing requirements for the memory devices. As expected for the “no series”
termination case, the waveform at the source does not look good but this does not
affect the system integrity at the load.
Figure 6.9 shows an example of one clock output driving two loads connected
using a daisy-chain topology. The distance from the source to the first load (first
86 6 Transmission Line (TL) Effects
SDRAM Load # 1
6" Trace
2" Trace
SDRAM Load # 2
Parallel
Termination 2" Trace
OSCILLOSCOPE
Design file: 5912CLK.TLN Designer: TI
BoardSim/LineSim, HyperLynx
Comment: NOTE: The signals recorded at the terminating loads is identical therefore only the magenta signal is shown.
4500.0 Probe 1:U(A0)
Probe 2:U(D0)
4000.0 Probe 3:U(D1)
3500.0
3000.0
V
o
l 2500.0
t
a 2000.0
g
e
- 1500.0
m
V
- 1000.0
500.0
0.000
-500.0
0.000 5.000 10.000 15.000 20.000 25.000 30.000 35.000 40.000 45.000 50.000
Time (ns)
Added RC
2 nd SDRAM Parallel
1 st SDRAM Termination
DM642 EVM
10 Ohms
Series
Term.
CLOCK
OUT FROM DSP
SDRAM) is the same as the distance from the first load to the second load (second
SDRAM). In this case, the reflections coming from the second load distort the clock
signal at the first load. The best way to minimize this distortion is by adding a parallel
termination at the second load to reduce the impedance mismatch and therefore
reduce the reflections as shown in Figs. 6.10 and 6.11. This system still requires a
CLOCK
1st Mem
2nd Mem
CLOCK
1st Mem
2nd Mem
series termination at the source to control the edge rate of the whole signal trace. This
resistor needs to be small so that the source and sink currents are large enough to
drive two loads. In this example, the series termination resistor is 10 Ω.
In general, high-speed DSP systems consist of many CMOS devices where the input
impedance is very high, typically in MΩ and the input capacitance is relatively
small, less than 20 pF. In this case, with no load termination, the TL looks like a
transmission line with a capacitive load, rather than an open circuit. The capacitive
load helps reduce the rise time and allows the designers to use only a series
termination at the source. This approach is becoming quite common in high-speed
systems.
In Fig.. 6.12, the voltage at the load is slowly charged up to the maximum
amplitude of the clock signal. Initially, the load looks like a short circuit. Once the
capacitor is fully charged, the load becomes an open circuit. The source resistor ZS
controls the rise and fall times. Higher source resistance yields slower rise time. The
load voltage at any instant of time, t, greater than the propagation delay time, can be
calculated using the following equation:
V L ¼ V clk 1 eðtT d Þ=τ , ð6:7Þ
where t is some instant of time greater than the propagation delay and τ ¼ CLZo,
where CL and Zo are the load capacitor and characteristic impedance, respectively.
One of the well-known techniques to analyze the PC board is using a signal integrity
software [3] to simulate the lines. Figure 6.13 shows a setup used for the simulations.
Zs = 25 Td = 1.8nS, Zo = 50 ohms
Transmission Line C
Vclk = 3.0V
OSCILLOSCOPE
Design file: 507201C.HYP Designer: TI
BoardSim/LineSim, HyperLynx
7.000 Probe 5:U2.E6
6.000
5.000
4.000
V ol t a g e -V -
3.000
2.000
1.000
0.000
-1.000
-2.000
-3.000
0.000 40.00 80.00 120.00 160.00 200.00
Time (ns)
As discussed earlier, most high-speed system designs use this technique since it is
possible to optimize the load waveforms simply by adjusting the series termination
resistors. This technique also helps reduce the dynamic power dissipation, since the
initial drive current is limited to the maximum source voltage divided by the
characteristic impedance. Figure 6.17 shows the setup used for the simulation of
the audio clock driven by an audio CODEC external to the DSP.
Figure 6.18 shows an audio clock that transmits by U17 and receives by U3. The
design has a 20-Ω series termination resistor but no parallel termination at the load.
This demonstrates the concept discussed earlier.
OSCILLOSCOPE
Design file: 507201C.HYP Designer: TI
BoardSim/LineSim, HyperLynx
7.000 Probe 5:U3.G13
6.000
5.000
4.000
V ol t a g e -V -
3.000
2.000
1.000
0.000
-1.000
-2.000
-3.000
0.000 40.00 80.00 120.00 160.00 200.00
Time (ns)
In summary, the simulation results correlate very well with the actual lab measure-
ments. Designers need to understand the TL characteristics and terminate traces to
minimize reflections that may cause random circuit failures, excessive noise injected
into the power, ground planes, and electromagnetic radiation.
One final comment about the TL is that the previous examples were based on a
model where a signal trace is on top of a ground plane known as a microstrip model.
Other techniques, such as a ground grid, are also commonly used. Example 6.2
demonstrates the effects of the ground grid. In this configuration, the designers need
to understand the current flows and their effect on the characteristic impedance.
Example 6.2 Figure 6.21 shows an example of using a ground grid, instead of
ground plane, for the PC board. As shown in this figure, the current path is not
immediately under the signal trace, so there is a large current return loop that yields
higher inductance and lower capacitance per unit length. In this case, the character-
istic impedance is higher than if a continuous ground plane was used.
Figure 6.22 also shows another example of using a ground grid where the signal is
being routed diagonally. As shown in this figure, the current return has to travel on a
zig-zag pattern back to the source and creates a large current return loop that yields
higher inductance and lower capacitance per unit length. In this case, the character-
istic impedance is higher than using a continuous ground plane and higher than the
case where the signal is routed in parallel with the ground grid as shown in Fig. 6.21.
So, if ground grid is required in a design, the best approach is to route the high-
speed signals right on top of the grids and parallel to the grid to ensure the smallest
current return loops. This lowers the characteristic impedance to the level equivalent
to the impedance of the continuous ground plane. This is difficult to accomplish
since a complex board has many high-speed traces. Therefore, continuous ground
plane is still the best method to keep characteristic impedance and EMI low.
Ground
grid Current
Return
Input
Signal
Signal trace is routed between the two ground paths of the grid
Input
Signal
Signal trace is routed diagonally
References
1. S. Hall et al., High Speed Digital System Design (John Wiley & Sons, New York, 2000)
2. Texas Instruments Inc., OMAP5912 Applications Processor Data Manual (2003). http://focus.ti.
com/lit/ds/symlink/omap5912.pdf
3. Mentor Graphics, Hyperlynx Signal Integrity Simulation Software (2004). http://www.mentor.
com/products/pcb-system-design/circuit-simulation/hyperlynx-signal-integrity/
Chapter 7
Transmission Line (TL) Effects
in Frequency Domain
Transmission line (TL) effects are one of the most common causes of noise and
design problems in high-speed systems. Traditionally, in digital design, the
switching frequency and the rise or fall time of the signal could generate excessive
overshoots or undershoots which can cause timing violations as described in
Chap. 6. As the operating frequency gets higher, >100 MHz, the digital waveform
must change to accommodate for fast switching, lower supply voltage, and lower
rise and fall times. The trend makes it more difficult for designers to control noise
and radiation while maintaining good signal and power integrities.
One of the most effective ways to reduce noise, EMI, and PCB routing area is
reducing the number of inputs and outputs to and from an IC. The common method
for device interconnects is converting parallel busses to serial busses in a system. But
to be logically compatible with parallel busses, the serial busses must run at a much
higher speed, depending on how many parallel data bits are being serialized. For
example, an 8-bit parallel bus running at 100 MHz can be directly converted to a
single-bit serial bus running at 800 MHz, the number of parallel bits multiplied by
the switching speed. Converting parallel-to-serial and serial-to-parallel method is
known as SERDES, Serializer, and Deserializer.
To push the switching speed to multi-gigahertz range, up to 32 Gbps, the industry
has adapted modeling and simulating techniques commonly being used in micro-
wave circuits, known as scattering parameter or s-parameter. This chapter covers the
fundamentals of s-parameter and how it is being applied in a high-speed digital
system.
With data rate running up to multi-gigahertz speed, the traditional method of digital
design is no longer valid. Digital designers must learn and adapt RF/microwave
design techniques in modeling systems or components using scattering parameters
Output Voltage, j
Sij ¼ , ð7:1Þ
Input Voltage, i
Output Voltage Amplitude
MagðsÞ, Magnitude ¼ , ð7:2Þ
Input Voltage Amplitude
SdB ¼ 20 log 10 ðMagðsÞÞ, ð7:3Þ
PhaseðsÞ ¼ PhaseðoutputÞ PhaseðinputÞ: ð7:4Þ
The most popular network used for analyzing differential pairs is a four-port
network shown in Fig. 7.2. In the four-port network, the naming convention of the
ports is the same as the two-port network in which S12 is a transmission coefficient
when a voltage applied to Port 2 is transferred to Port 1.
In SERDES protocol, before transmitting a single-ended signal from one device
or system to another, the single-ended signal is first converted to one different pair,
which has a positive component and a negative (180 out of phase) component of the
original signal. Then, the differential pair is transmitted on a PCB to its destination.
The advantages of differential signaling are good noise immunity, low radiation, and
low dynamic power consumption as described in Chap. 5.
In Fig. 7.2 four-port network, the insertion losses (IL) are S12, S21, S34, and S43.
And the return losses (RL) are S11, S22, S33, and S44. In addition to insertion and
return losses, the four-port network has crosstalk terms, S13, S31, S14, S41, S32,
S23, S42, and S24. The crosstalk term is defined as the voltage applied at one port is
coupled to another port. NEXT is a near-end crosstalk where a voltage applied to one
port couples to the nearest port, and FEXT is a far-end crosstalk where a voltage
applied to one port couples to the farthest port, not on the same channel as the
transmitted port. For example, S14 is a far-end crosstalk when a voltage applied to
Port 4 (far end port) couples to Port 1; Port 1 is not on the same transmission path as
Port 4 (Fig. 7.3).
The four-port network s-parameters consist of different modes, known as mixed-
mode s-parameters. The modes include how the network responds to differential and
common mode stimulus signals. The table in Fig. 7.4 demonstrates different modes
of the four-port mix-mode s-parameters.
The end-to-end HyperLynx LineSim [3] circuit shown in Fig. 7.10 consists of a four-
port network model extracted from the layout in Fig. 7.9 and the Pericom PI3EQX1001
Redriver model (PI3EQX1001_FG1_EQ1_TTT_4PORT_S_PARAMETER_REVA).
Simulation results in Fig. 7.11 show the insertion loss curves with and without
Retimer. With the Retimer, there is a gain of +9 dB at 10 GHz as compared to a loss
of 6 dB at the same frequency without the Retimer.
standards only allow for two vias on a high-speed trace. Refer to Chap. 14 for
more details about PCB layout,
• If vias are required for routing, make sure to add ground via next to the signal to
allow for high-speed current return. Also, verify if the stubs on via cause any
signal integrity issues or not. Methods like microvias or laser back-drilling are
commonly being used to remove via stubs. Vias on high-speed traces must be
modeled and included in the end-to-end simulations. Also, via placement is
critical, so make sure to review the PCB guidelines required for the intended
protocol. In general, place the via as close to the source or destination as possible.
References
In any electronic system, it is neither practical nor is it necessary to eliminate all the
noise, as noise is not a problem until it interferes with the surrounding circuitries or
radiates electromagnetic energy that exceeds FCC limits and or degrades the system
performance. When noise interferes with other circuits it is called crosstalk.
Crosstalk can be transmitted through electromagnetic radiation or electrically cou-
pling, such as when unwanted signals propagate on the power and ground planes or
couple onto the adjacent circuits. One of the most challenging problems designers
are facing in today’s electronic systems is to determine the source of crosstalk,
especially in the case where crosstalk randomly causes system failures. Because
components are so tightly packed into a very small printed circuit board (PCB). This
chapter outlines the crosstalk mechanisms and design methodologies to minimize
the effects of crosstalk.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 103
T. T. Tran, High-Speed System and Analog Input/Output Design,
https://doi.org/10.1007/978-3-031-04954-5_8
104 8 Effects of Crosstalk
SIGNAL
TRACE ON
PCB
SOURCE CURRENT
VIDEO VIDEO
AMP RETURN CURRENT AMP
CURRENT RETURNS
BELOW THE SIGNAL
TRACE (SMALLEST
AREA)
SIGNAL
TRACE ON
PCB
SOURCE CURRENT
AUDIO AUDIO
AMP RETURN CURRENT AMP
CURRENT RETURNS
ON THE SHORTEST
PATH
current
density curve
signal trace
H
Ground
signal current, on the other hand, returns on the path of least inductance, normally
underneath the signal trace. Knowing the current return paths is important for
designers to optimize the system design to reduce crosstalk.
The current return density and the amount of crosstalk can be estimated as shown
in Figs. 8.3 and 8.4. Based on the equations shown in the figures, the spacing
between the traces and the distance that they run in parallel determines the amount
of crosstalk. Obviously, moving the traces further from each other will reduce the
crosstalk.
8.1 Current Return Paths 105
H
Ground
I 1
ID ¼ ð8:1Þ
πH 1 þ D2
,
H
K
Crosstalk ¼ 2 , ð8:2Þ
1 þ HD
where D is the distance between the traces, H is the height of the signal to the
reference plane, and K is the coupling constant less than 1.
There are two types of crosstalk, forward and backward. Forward crosstalk, also
known as capacitively coupled crosstalk. This occurs when the current flows in the
same direction as the source. With backward crosstalk, which is also called induc-
tively coupled crosstalk, the coupling current flows in the opposite direction of the
source.
The following simulations [2] demonstrate the concept of reducing forward and
backward crosstalk by spacing the aggressor and victim traces. The model simulates
two parallel 5 mils wide, 12-in.-long traces. The source of the trace is connected to a
DSP and the load to DDR memory. As shown in Fig. 8.5, D0 line is an aggressor and
D1 line is a victim.
Figure 8.6 shows simulation results. On the victim trace, the first negative-going
pulse, which has a 200 mV peak, is the forward crosstalk. The positive-going pulse
of 240 mV is the backward crosstalk. The backward pulse width is about two times
the coupling region. In this case, the coupling region is 3.54 nS and the simulation
also shows a 4 nS backward crosstalk pulse.
The crosstalk Eq. (8.2) is
K
2 :
1 þ HD
106 8 Effects of Crosstalk
OSCILLOSCOPE
Design file: UNNAMED0.TLN Designer: TI
BoardSim/LineSim, HyperLynx
3000.0 Probe 3:U(B0).DQ0
Probe 4:U(B1).DQ1
2500.0
2000.0
1500.0
1000.0
Voltage-mV-
500.0
0.000
-500.0
-1000.0
-1500.0
-2000.0
0.000 4.000 8.000 12.000 16.000 20.000
Time (ns)
K 1
Max: Crosstalk ¼ D2 ¼ 5 2 ¼ 0:8 V:
1þ H 1þ 10
As expected, the simulations in Fig. 8.6 showed that the peak-to-peak crosstalk is
440 mV, which is much less than the maximum crosstalk estimated.
Now, let us test the condition where the two traces are placed further away from
each other, by making D ¼ 15 mils. The maximum estimated crosstalk is now 0.3 V
while the simulation in Fig. 8.7 shows a forward crosstalk of 100 mV and a
backward crosstalk of 90 mV. The peak-to-peak crosstalk is about 190 mV, which
is again much less than the calculated maximum of 300 mV. This simulation
demonstrates how the rule-of-thumb provided earlier overestimates crosstalk.
In summary, accurately calculating and simulating the crosstalk of a system is not
possible due to many complex capacitive and inductive coupling paths that are
involved. The examples show how difficult it is to estimate and simulate crosstalk
OSCILLOSCOPE
Design file: UNNAMED0.TLN Designer: TI
BoardSim/LineSim, HyperLynx
3000.0 Probe 3:U(B0).DQ0
Probe 4:U(B1).DQ1
2500.0
2000.0
1500.0
1000.0
Voltage -mV-
500.0
0.000
-500.0
-1000.0
-1500.0
-2000.0
0.000 2.000 4.000 6.000 8.000 10.000 12.000 14.000 16.000 18.000 20.000
Time (ns)
and the effects of spacing on the adjacent signal. The following points need to be
considered before finalizing the design:
• When the PCB is designed, minimize the height, H, between the high-speed
signal routing layer and the ground plane. Lower H yields lower crosstalk.
• Maximize the spacing, D, between the signals. Higher D yields lower crosstalk.
• For board layout, analyze the critical signals and minimize the coupling regions.
• Slow the edge rates if possible because this reduces crosstalk.
Crosstalk can also be caused by high-speed signals that are routed on traces that form
effective antennas. The first step in determining whether a trace is acting as an
antenna is to calculate the wavelength of the signal using the following equation:
C
λ¼ , ð8:3Þ
f
where C is the speed of light or 3 108 m/s and f is the frequency in Hz.
The equation shows that a 100 MHz clock signal has a wavelength of 3 m or
9.84 ft. A good rule for minimizing radiation is making sure that the trace length is
not longer than the wavelength divided by 20. So, in the case of the 100 MHz clock
signal, the signal length should be kept below 0.15 m or 0.492 ft. Keeping the traces
below 0.5 ft is easy, but the square wave clock signal consists of multiple harmonics
and each of the harmonics can radiate even when the traces are very short. Here is an
example.
Example 8.1 Let f ¼ 500 MHz, the fifth harmonic of the 100 MHz clock,
C 3 108
λ¼ ¼ ¼ 0:6 m:
f 500 106
C 3 108
λ¼ ¼ ¼ 0:03 m or 3 cm which is 1:18 in:
20f 20 500 106
What this means is that depending on the rise and fall times of the 100 MHz
signal, the energy of the fifth harmonic can radiate and interfere with the adjacent
circuits when this signal trace is longer than 1.18 in. The energy of the harmonics
depends on the rise and fall times of the signal as shown in Fig. 8.8. In this figure, it
is assumed that the clock waveform has a 50% duty cycle, and rise and fall times are
8.2 Crosstalk Caused by Radiation 109
Tr Pw
T
1
20dB/decade
Amplitude π Pw 1
π Tr
40dB/decade
Higher Tr reduces
amplitude of harmonics
1/T (3) (5)
Frequency (harmonic)
equal. With these assumptions, only odd harmonics of the clock are present. The
amplitude of the harmonics starts decaying at the first pole frequency, f1, at a rate of
20 dB/decade and then increases to 40 dB/decade at the second pole frequency f2.
The equations for f1 and f2 are
1
f1 ¼ , ð8:4Þ
Pw
1
f2 ¼ , ð8:5Þ
πT r
AUDIO
CODEC MPEG-2
CLOCKS
ENCODE
VIDEO
DECODE&
CPU MPEG-2
ENCODE
DECODE
MODEM
TO
SPEAKERS MEMORY
AUDIO
CODEC MPEG-2
ENCODE
RF IN VIDEO
TUNER
DECODE CPU
MPEG-2
TO DECODE
VIDEO
DISPLAY
ENCODE HARD
DRIVE
• The Video Decoder receives the baseband video signal from the Tuner and
digitizes the signal to prepare for digital signal compression. This video data
rate for an analog TV channel with 640 480 30 frames/s resolution is around
147 Mbits/s (data rate ¼ 640 480 16 bits/pixel 30 frames/s ¼ 147 Mbits/s).
The CPU is responsible for running a high-level operating system and for
managing all the video and audio data. The digital video data are captured by the
CPU and are stored in external memories such as DDR.
The CPU transfers the digital video data stored in the memories to the MPEG-2
Encoder [4]. The Encoder compresses the data from 147 to 2 Mbits/s bit rate and
sends the compressed bit rate back to the external memories. The compressed data
are then sent to the hard drive for storage.
• The CPU then reads the compressed data from the hard drive and sends the data to
the MPEG-2 Decoder [4]. The Decoder decompresses the data from 2 Mbits/s bit
rate back to 147 Mbits/s bit rate and sends it to the Video Encoder. The Encoder
converts the digital decoded data to analog signals and displays it on a TV screen.
• The Modem in the time-shifting system is there for communicating with the
service provider server to request TV guides and software updates.
8.2 Crosstalk Caused by Radiation 111
• The Audio CODEC is responsible for digitizing the analog audio signals received
from the Tuner or an external audio device. The digital audio data follows the
same path as the digital video signal in which the data are being compressed,
stored, and playback the same way. This CODEC also receives the digital audio
data from the CPU and converts it to analog signals for playing out to the
speakers.
In this time-shifting system design, the crosstalk can occur anywhere within a
system even though the system was fully FCC certified, so it is difficult to find the
root causes of the problem. For example, any clock can generate harmonics that
radiate to the antenna input and interfere with the TV channel. To illustrate this
effect, Fig. 8.11 shows a video screen with horizontal lines generated by the third
harmonic (55.2 MHz) of the 18.4 MHz clock radiating to the antenna input.
55.2 MHz harmonic happens to be within Channel 2 (54–60 MHz) of the NTSC
spectrum [5] and causes interferences that cannot be rejected by the Tuner because
the Tuner cannot distinguish between the in-band noise and the actual TV signal.
In this case, the best way to get rid of the interference is to reduce the energy
radiated by the third harmonic of the 18.4 MHz clock. Figure 8.8 shows that
increasing the rise time of the signal, assuming that this is not causing any setup
and hold time violations, attenuates the harmonic amplitude and reduces the radia-
tion. The two ways to reduce the rise time are lowering the clock buffer slew rate if
possible or adding a series termination resistor at the output of the clock buffer as
shown in Fig. 8.12.
It is very difficult to calculate the value of the resistor R, so the best method is to
vary the resistance until the noise has longer appeared on the display. At this point,
INTERFERENCE
(JAGGED
LINES ON THE
SCREEN)
R AUDIO
CLOCK
BUFFER CODEC
112 8 Effects of Crosstalk
NO
INTERFERENCE
measure the rise and fall times of the 18.4 MHz clock and verify that reducing the
rise and fall times does not cause any timing violations. Figure 8.13 shows a design
example of having a 75-Ω series termination resistor to reduce the rise time and get
rid of the interference as shown in Fig. 8.11.
8.3 Summary
As highlighted in Sects. 8.1 and 8.2, the current return paths and the signal rise time
play the key role in generating crosstalk that interferes with the adjacent circuits and
causes random system failures and or system performance degradation.
• Slow down the rise and fall times if possible. Increasing the rise time reduces the
power spectral density of the harmonics as shown in Fig. 8.8.
• Keep high-speed signals as short as possible and make sure that the wavelengths
of the third and fifth harmonics of the signal are much less than the wavelength
divided by 20.
• Always add a series termination at the source of the clock signal as demonstrated
in Fig. 8.12. This helps reduce the transmission line effects and provides a way to
reduce the rise time if necessary.
• Always route the high-speed signals away from any critical high impedance
traces. High impedance traces are the input traces to the video and audio ampli-
fiers. Also, space the traces at least one width apart from each other to reduce the
coupling; for example, a 5 mils trace should have a gap of at least 5 mils to
another trace.
• The best method to minimize crosstalk is to place the clock generator component
in the middle of the system as shown in Fig. 8.9. This ensures minimum routing
clock traces to all other sections.
Using a spread spectrum clock generator is another way to reduce the peak
radiated power but be careful with the jitter generated by these clock buffer devices.
For example, in one spread spectrum clock buffer [6], the amplitude of the seventh
References 113
harmonic can be attenuated by 13 dB by setting the spread spectrum at 2%. The
issue with this setting is that the 100 MHz clock output jitter will increase by 529 ps.
If this increase in jitter is still within the allowable limits of the DSP clock input, then
this solution is acceptable.
References
1. H. Johnson, M. Graham, High-Speed Digital System Design (Prentice Hall, Englewood Cliffs,
1993)
2. Mentor Graphics, HyperLynx Signal Integrity Simulation Software (2004). http://www.mentor.
com/products/pcb-system-design/circuit-simulation/hyperlynx-signal-integrity/
3. TIVO DVR. http://www.tivo.com/dvr-products/tivo-hd-dvr/index.html
4. International Organization for Standardization, Information Technology—Generic Coding of
Moving Pictures and Associated Audio Information: Video. ISO/IEC 13818-2:2000 (2000)
5. Standard NTSC Channels & Frequencies. http://radiotechnicalservices.com/tvchannels.pdf
6. Texas Instruments Inc., Spread Spectrum Clocking Using the CDCS502/503 Application Report,
SCAA103 (2009). http://focus.ti.com/lit/an/scaa103/scaa103.pdf
Chapter 9
Memory Sub-system Design Considerations
The most critical bus in a DSP system today is the memory bus where a large amount
of ultra-high-speed data is being transferred from the DSP to the physical memory
devices and vice versa. The data on this bus are switching very fast. The rise and
fall times of the data, memory clocks, and control signals are approaching
sub-nanosecond range. These fast transients generate noise, radiation, power supply
droops, signal integrity, and memory timing issues. This chapter covers memory
sub-system design techniques to minimize the effects of high-speed data
propagating.
It is assumed in this chapter that the memory is DDR memory, since DDR design
presents many challenges as DDR transmits and receives data at both edges of the
memory clock. These include a noise-sensitive analog circuit called Delay Locked
Loop or DLL. In this case, the internal and external noise can cause excessive DLL
jitter which leads to memory failures.
The three types of DDR memories shown in Table 9.1 are DDR1, DDR2, DDR3,
DDR4, and LPDDR4 or Low Power DDR4. The transfer rate in DDR is generally
defined as MT/s or Mega Transfers per Second. So DDR4-3200 is DDR4 data
transfer rate of 3200 MT/s. In this case, the clock rate is 1600 MHz as each edge
of the clock initiates one transfer.
Even though DDR has evolved from DDR1 to DDR4, soon to be DDR5, the
fundamentals of DDR Write and Read Cycles remain essentially unchanged.
Figure 9.1 shows the basic DSP and DDR interface, and the signal definitions are
in Table 9.2.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 115
T. T. Tran, High-Speed System and Analog Input/Output Design,
https://doi.org/10.1007/978-3-031-04954-5_9
116 9 Memory Sub-system Design Considerations
DSP DDR
SDRAM
CLK, CLK# CLK, CLK#
DQS DQS
DQM DQM
WE#, CAS#, WE#, CAS#,
RAS# RAS#
ADDRESS ADDRESS
VDD/2
VREF VREF
VOLTAGE
VDD
DIVIDER
Figure 9.2 shows a memory Write Cycle where D0 is the least significant data bit and
DQS is the Data Strobe. The data transfers happen at both edges of DQS and DQS is
typically a 90 phase-shift from the first data burst. Figure 9.3 shows an oscilloscope
capture of the actual memory bus.
9.1 DDR Memory Overview 117
DQS
1 0 1 0 1 0 1 0
D0
DQS
1 0 1 0 1 0 1 0
D0
Figure 9.4 shows a memory Read Cycle where D0 is the least significant data bit and
DQS is the Data Strobe. The data transfers happen at both edges of DQS and DQS is
synchronized with D0. In the Read Cycle, memory device outputs DQS and drives
the data bus synchronously.
As covered in Chaps. 2 and 3, transmission line effects and crosstalk are results of
bad signal integrity design. Since memory timing is so critical, the excessive over-
shoots, undershoots, and glitches on the signal can false trigger and cause memory
read and write failures. The worst part is that the memory controller such as DDR
controller relies on a noise-sensitive analog circuit such as DLL to synchronize and
delay the strobes to read the incoming data from external memory devices.
Simulated Figs. 9.5, 9.6, and 9.7 [1] demonstrate good, bad, and ugly signal
integrity designs, respectively. The bad case in Fig. 9.6 has a glitch right at the
switching threshold which may cause false clocking; the system may see this glitch
as a high and low transition just like a clock input and respond to it.
For the ugly case in Fig. 9.7, the overshoots and undershoots are so excessive that
the peak of the overshoot crosses over the minimum input high voltage, Vih, and
causes false clocking. Also, these overshoots and undershoots generate a lot of noise
and radiation. The rule-of-thumb is to fine-tune the signal until all the overshoots are
much lower than the threshold voltage, Vih.
DDR2 Memory System Design Example 9.1
The following design rules for good memory signal integrity are:
• Apply good decoupling techniques as shown in Chap. 5. It is highly
recommended to have one high-frequency capacitor (0.01–0.22 μF) per DDR
power pin and one bulk capacitor (10 μF) for the DDR region. These decoupling
techniques are also required for the memory devices in the design.
9.2 DDR Memory Signal Integrity 119
Glitches at
the signal
switching
threshold
• Add termination resistors on the data bus and the control signals. Where to place
the resistors on the bidirectional bus depends on which device has higher drive
strength. Place the resistors nearby the device with the higher drive strength. For
example, DDR memory devices typically have strong buffers to allow for non-
120 9 Memory Sub-system Design Considerations
Overshoot
crosses over
the input
high voltage
& causes
false
triggering
embedded designs such as PC. In this case, always put the termination resistors
right by the memory devices. Ideally, add termination resistors at the output of the
device driving the bidirectional bus. See the design examples in Figs. 9.8 [2]
and 9.9.
• Isolate and decouple the DLL power supply and the VREF voltage pins. For the
DLL, follow the rules described in the PLL chapter, Chap. 6. To generate VREF,
use a resistor divider and divide the memory power supply voltage, Vdd, required
for both DSP and memory devices. See the design example in Fig. 9.9 [3].
DDR4 Memory System Design Example 9.2 For DDR4-3200 or higher designs, it
is difficult to manually characterize timings as done in Sect. 9.1, because specialized
equipment such as Logic Analyzer is not readily available to measure pico-second
timings. The industry now relies on automated design tools to analyze DDR memory
buses and to verify timing margins. One example here is HyperLynx DDR Batch
Simulation [1, 4], which runs worst-case analysis for data read/write and command/
address, and provides comprehensive reports with pass/fail tests, including eye
diagrams as in Figs. 9.10 and 9.11.
In Fig. 9.10, the blue block in the middle of the eye diagram is the mask or the
keep-out region in which any signal touching this mask indicates DDR timing
violations. Figure 9.11 shows passes and fails of all the WRITE/READ cycles. For
failed cases, engineers must find the root cause and fix the memory bus design to
pass all WRITE/READ cycles before releasing the PCB to fabrication. The last step
of memory validation is to run extensive memory tests on the actual hardware to
verify timings and guarantee compatibility.
9.2 DDR Memory Signal Integrity 121
References
USB 3.1 is one of the latest industry standards which operates up to 10 Gbps speed.
Designing USB channel requires extensive s-parameter simulations and running
compliance tests. This chapter shows an example of how to use HyperLynx
SERDES Compliance Wizard [1] tool to develop and simulate USB 3.1 channel,
and to check USB compatibility.
The design objective is to develop a USB channel that has a length of 18 in. between
the USB HUB and USB Device and determine if USB Redriver is required to
comply with USB standards. Figure 10.1 shows the block diagram of this design
example.
To begin, use a PCB design software and design two differential pairs, and each
pair has 100 Ω differential impedance. From the layout, extract the differential pairs
and generate an s-parameter model, having 4 ports and 16 in. in length. Then, use
LineSim in HyperLynx to simulate the channel. Since the two channels, TRANS-
MIT (TX) and RECEIVE (RX), are identical, only one channel is being designed to
demonstrate the concept here. In practice, both channels must be modeled to make
sure that the design is fully compliant.
The LineSim circuit in Fig. 10.2 is the USB channel design based on the datapath
defined in the block diagram, Fig. 10.1.
As expected, the simulation results in Fig. 10.3 show the design failed Eye
Height, Blue Mask (minimum open area required for USB) touching the border of
the Eye Diagram. This concludes that a Redriver is required to compensate for the
losses associated with a 16-in. channel.
Now, to add a Redriver to the design, a circuit in Fig. 10.4 was created in
HyperLynx LineSim [1] consists of one TI TUSB1002A USB 3.2 Redriver [2],
USB Transmit, USB Receive, and the PCB Channel. The Channel includes three
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 125
T. T. Tran, High-Speed System and Analog Input/Output Design,
https://doi.org/10.1007/978-3-031-04954-5_10
126 10 USB 3.1 Channel Design
Fig. 10.2 HyperLynx LineSim circuit of USB 3.1 channel with no redrivers or retimers in the
transmission paths. J1, J2 and J3 are s-parameter models of the channel extracted from the PCB
layout.
s-parameter models, and each model is a four-port network 6 in. in length. The total
channel is 18 in.
When the Redriver is inserted into the USB channel to compensate for the
channel loss, the simulation must be done in a closed loop in which USB Transmit
(U1) sends data to Redriver input (U3), and Redriver (U3) sends data output back to
USB Receive (U1). In this configuration, one simulation reports the performance of
the complete USB loop, transmit and receive paths.
The simulation results in Fig. 10.5 now show PASS (USB 3.1 compliance) on
both paths, from U1 to U3 and U3 to U1. The Eye Width and Eye Height were
measured at the Bit Error Rate (BER) of 1 10 12.
Figure 10.6 shows the Eye Diagram of the USB channel from U1 to U3 that
passes the compliance tests with plenty of margin. For example, the USB standard
requires a minimum Eye Height of 70 mV while the design has an Eye Height of
464 mV. And the minimum Eye Width is 0.286 UI while the design has an Eye
Width of 0.774 UI. UI is Unit Interval which is one divided by the bitrate. For
USB 3.1 10 Gbps, the UI is 100 pS.
Again, Fig. 10.7 shows the Eye Diagram of the USB channel from U3 to U1 that
passes the compliance tests with plenty of margin. Eye Width and Eye Height were
measured 0.856 UI and 0.555 UI, respectively.
Fig. 10.9 Serpentines placed near the via with the shorter trace
• Always route the differential pair together and if bends are necessary, use rounded
bends. Also, if serpentines need to be added to match the positive and negative
signals, put the serpentines at the end near the area that causes the mismatch as
shown in Fig. 10.9.
In this design example, only the channel on a PCB was analyzed. In practice, in
addition to the channel, USB design must include all other losses associated with
USB connectors, cables, transmitter, and receiver. The good thing here is that the
simulation models of these components are readily available, so designers can
incorporate the models in the simulations to validate if the complete design passes
the compliance tests.
References
PLL is the heart of practically all electronic components and or modules where
different clock frequencies are required to synchronize the data transmitting and
receiving to and from externals, respectively. The input clock to the PLL is much
lower than the DSP maximum clock frequency. PLL is typically used as a frequency
synthesizer to generate the clock for the DSP core. For example, the input clock to
the 1.2 GHz DSP [1] is 66 MHz.
PLL is an analog circuit that is very sensitive to power supply noise. Noise causes
jitter and excessive jitter causes timing violations which lead to system failures. The
two main PLL architectures are analog PLL (APLL) and digital PLL (DPLL).
Understanding the differences helps to make the design tradeoffs often required
for minimizing noise and jitter caused by external circuitries, such as the power
supply and other noisy switching devices.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 131
T. T. Tran, High-Speed System and Analog Input/Output Design,
https://doi.org/10.1007/978-3-031-04954-5_11
132 11 Phase-Locked Loop (PLL)
/M
M
f OUT = f IN
N
3. Based on the error signal, the CP charges or discharges the current store on the
loop filter, an RC filter is shown in Fig. 11.1. This increases or decreases the VCO
control voltage. For some PLL architectures, increasing the VCO control voltage
increases the clock frequency and decreasing the voltage lowers the clock output
frequency.
4. The phase correction continues until both the feedback signal from the Divide-by-
M counter and the reference clock are synchronized. At this point, the error
voltage should be zero.
5. The output clock frequency is equal to the ratio of the Divide-by-M counter and
the Divide-by-N counter multiplied by the input clock frequency. As a rule-of-
thumb, a higher multiplier ratio yields higher jitter, so keep the M and N ratio as
low as possible when designing with PLLs. The PLL output frequency, fout, for a
given input frequency, fin, is
M
f out ¼ f , ð11:1Þ
N in
Jitter in PLL design is defined as the signal timing displacement from a reference
clock. The three main sources of DSP PLL jitter are jitter generated by the reference
clock itself, power supply noise, and noise coupling from external and internal
circuitries. The following lists important techniques for designers to minimize the
DSP PLL jitter:
• Select a reference clock oscillator with the lowest jitter specification possible.
• Heavily filter the clock circuit to reduce the effect of noise on the output jitter. See
the following section on PLL isolation.
• Use a series termination resistor at the output of the reference clock to control the
edge rate.
• Distribute the clock differentially if possible. Differential signals reject common
mode noise and crosstalk.
• Set the multiplier as low as possible to achieve maximum DSP operating fre-
quency. Keep in mind that a higher multiply ratio yields higher output jitter.
In all cases, jitter can be minimized but cannot be eliminated. The three types of
deterministic jitter [2] important for frequency synthesizers and DSP performance
are long-term jitter, cycle-to-cycle jitter, and period jitter.
See Fig. 11.2 where long-term jitter is defined as a time displacement from the ideal
reference clock input over a large number of transitions. Long-term jitter measures
the deviation of a rising edge over a large number of cycles (N ) after the first rising
edge.
where Max Period is the maximum period equal to 1 divided by the operating
frequency measured at N number of cycles and Min Period is the minimum period
equal to 1 divided by the operating frequency measured at N number of cycles.
The long-term jitter can be measured using automatic jitter measurement equip-
ment [3] or using a high-speed digital sampling scope. Here are the steps to measure
long-term jitter using a scope:
• Use a high-speed 10 GHz sampling oscilloscope.
• Use the input clock to trigger the scope and set the scope in the Infinite
Persistence mode.
• The deviation is measured from the first rising edge to the Nth cycle. The “fuzz”
shown on the scope in Fig. 11.3 is the long-term jitter.
See Fig. 11.4 where cycle-to-cycle is defined as the deviation of the clock period
between two consecutive clock cycles.
In Fig. 11.4, the cycle-to-cycle is measured by subtracting t2 from t1, t3 from t2,
and so on.
See Fig. 11.5 where period jitter is defined as the maximum deviation in the clock’s
transition from its ideal position. These periods are non-successive.
The main differences between the APLL and DPLL are that the DPLL replaces the
analog filter with a digital controller block that filters the phase error in the digital
domain and replaces the VCO with a Digital Controller Oscillator (DCO). The
advantages of the DPLL are:
• The DPLL supports a wide range of input frequencies from 30 kHz to 65 MHz or
higher.
• The DPLL design requires a smaller silicon area to implement and consumes less
power than the APLL.
• The DPLL does not have analog filter components such as capacitors which can
cause leakage current. This leads to lower power consumption.
• The DPLL block is scalable and portable. The same design can be implemented
on different process technology nodes.
• The DPLL design can be optimized for low jitter. But it may not be acceptable for
jitter sensitive designs such as USB, audio, and video clocks.
The disadvantages of the DPLL are:
• It is very sensitive to external and internal power supply noise. Use of a linear
regulator plus a Pi filter to isolate the power supply from the DPLL is
recommended.
• Low power supply rejection ratio.
• In addition to power supply sensitivity, quantization noise and phase detector
dead zone are the major sources of output jitter.
• Requiring a DAC block to control the oscillator. This makes the DPLL more
sensitive to noise.
Figure 11.6 shows a typical DPLL architecture [5] and Table 11.2 describes the
function of each block in the architecture.
/M
M
f OUT = f IN
N
11.4 APLL and DPLL Jitter Characterization 137
Table 11.3 shows a jitter comparison between an analog and a digital PLL that shows
the effects of process variation where Hot is fast, Cold is slow and Baseline is
typical. In this DSP design, the DPLL power supply is isolated by an internal low
dropout regulator (LDO) while the APLL is connected directly to the common
power supply plane. To test the noise sensitivity, 100 mV of noise modulating
from 100 Hz to 1 MHz is injected into the power supply rails. The results showed
that the peak-to-peak period jitter is less than 3% for the DPLL and is less than 2%
for the APLL. With the LDO, the DPLL jitter is less than 4% up to 50 mV of noise
on the power supply.
Designers need to be careful when injecting a signal onto the power supply to do
jitter measurements. The nature of the signal used for simulating a noisy power
supply condition can have a major impact on the PLL jitter. A squarewave signal
with a frequency less than the PLL bandwidth characterizes the worst case PLL jitter.
As far as the amplitude of the noise, the peak-to-peak voltage must be within the
power supply limits. For example, for a 1.6 V 3% Core, the maximum acceptable
peak-to-peak noise is 96 mV (48 mV min and þ48 mV max).
138 11 Phase-Locked Loop (PLL)
As shown in previous sections, both an APLL and a DPLL are sensitive to noise,
especially to noise frequency within the PLL bandwidth. PLL isolation is needed in
order to prevent the high-frequency PLL signal from propagating out of the PLL
section and affecting other circuits. PLL isolation can also attenuate the external
noise propagating to the PLL circuit which causes excessive jitter. In many cases,
external power supply noise causes the PLL to go unstable and the DSP to lock-up
randomly.
The two important filter schemes discussed in this document to isolate the PLL are
low-frequency filtering and high-frequency filtering. For high-frequency filtering, a
Pi or T network filter can be used as shown in Figs. 11.7 and 11.8:
The Pi filter circuit consists of one ferrite bead, L and two capacitors, C1 and C2.
This circuit provides both input and output isolation where noise from the 3.3 V
supply is attenuated by the ferrite bead and the C2 capacitor and noise generated by
the PLL circuit is isolated by the ferrite bead and the C1 capacitor. Refer to Chap. 13
for the filter design and simulation information.
FERRITE BEAD
C1 C2
VDD_3V3 VDD_PLL
L1 L2
C1
A T filter consists of two ferrite beads and one capacitor as shown in Fig. 11.8.
Just like in a Pi filter, 3.3 V supply noise is attenuated by the L1 ferrite bead and the
C1 capacitor and PLL noise is isolated by the L2 ferrite bead and C1 capacitor. Refer
to Chap. 13 for the filter design and simulation information.
Both Pi and T circuits are good for filtering high-frequency noise but they are not
as effective for low-frequency filtering since ferrite beads have almost zero AC
impedance at low frequency. The Pi circuit has an advantage over the T circuit.
Because this topology makes it possible to place the capacitor closer to the PLL
voltage pin that ensures low impedance to ground and also the smallest current loop
area, which reduces noise and EMI.
For low-frequency isolation, there are two common techniques, Pi filter with
large bulk capacitor and linear voltage regulator.
One method for low-frequency filtering is shown in Fig. 11.9, where a resistor
R replaces the ferrite bead and a bulk capacitor C3 (10–33 μF) is added to the circuit.
Low-frequency noise is attenuated by the resistor R and the bulk capacitor C3. The
resistor needs to be selected such that the voltage drop across the resistor is
negligible. The low frequency 3 dB corner for this filter is approximated by
Eq. (11.2). Notice that C1 and C2 are negligible in this case since its value is a lot
lower than the bulk capacitor C3.
1
f 3dB ¼ : ð11:2Þ
2πRC 3
Design Example 11.1 Design a PLL power supply filtering circuit that provides a
20 dB attenuation at 15 kHz. The tolerance for the PLL power supply is 5% and the
maximum current consumption is 10 mA.
Design steps are:
• The Pi filter circuit in Fig. 11.9 is a single pole filter neglecting C1 and C2. For a
single pole filter, the attenuation is 20 dB/decade starting at the 3 dB corner
frequency as shown in Eq. (11.2).
• f20dB ¼ 10 f3dB, slope is 20 dB/dec so the frequency at 20 dB is equal to
10 times the frequency at 3 dB. Therefore, f3dB ¼ (15 kHz)/10 ¼ 1.5 kHz.
VDD_3V3 VDD_PLL
R
+
C1 C2 C3
10uf
1
f 3dB ¼ ¼ 1:5 kHz,
2πRC 3
VDD_3V3 R3 10 VDD_PLL
C2 10n
C1 10n
C3 10u
VDD_5V VDD_PLL
Linear
Reg
+
C1 C2 C3
10uf
low-frequency transients and high-frequency noise from entering the PLL circuit.
The method shown in Fig. 11.12 is more expensive to implement than other methods
described previously. But it is extremely effective in keeping the PLL voltage as
clean as possible to guarantee lowest PLL jitter. Refer to Chap. 4 for design
considerations.
One issue with using a linear regulator is that it does not reject high frequency
very well. As shown in Fig. 11.13, the ripple rejection is approaching 0 dB
(no rejection at all) for noise that is higher than 1 MHz. This high-frequency noise
can cause more jitter in the PLL.
In summary, the best way to isolate PLL is using a combination of Pi filter and
linear regulator. In this case, the Pi filter can be implemented with ferrite bead and
142 11 Phase-Locked Loop (PLL)
VDD_5V L1 VDD_PLL
LINEAR
REGULATOR
+
C3 C4
C2 C1
11.6 Summary
Because of low power consumption and fast response time, most of the PLL designs
integrated into the DSP today are based on digital PLL concepts. As discussed,
DPLL is very sensitive to power supply and input noise, so proper design noise
isolation filters are required to achieve the lowest jitter possible. The best approach is
using a combination of Pi filter and linear regulator as shown in Sect. 11.5.2. This
may not be possible due to PCB space limitations so designers must make design
compromises. If there is not enough room for the regulator circuit, then
implementing a Pi filter using a resistor instead of a ferrite bead is the second-best
References 143
References
Power supply design is perhaps the most challenging aspect of the entire process of
controlling noise and radiation in high-speed system design. This is largely because
of the complexity of the dynamic load switching conditions. These include devices
going into or out of low power modes, excessive in-rush current due to bus
contention and charging decoupling capacitors, large voltage droop due to inade-
quate decoupling and layout, oscillations that overload the linear regulator output,
and high current switching noise generated by switching voltage regulators. A clean
and stable power supply design is required for all devices to guarantee system
stability. This chapter outlines the importance of proper power supply design and
the methods to minimize unwanted noise.
The two types of power supplies commonly being used in high-speed systems are
linear and switching power supplies. The linear power supply has the best low noise
characteristics typically required for analog audio, video, PLL, and data converter
circuits. The disadvantages of this architecture are its power efficiency and dissipa-
tion. As shown in Fig. 12.1, the linear power supply consists of two main stages,
input/output transistor and error amplifier. The input DC voltage here must be higher
than the output voltage and the minimum input voltage varies depending on the
component selected. So, it is important for designers to review the power supply’s
specifications and set input and output voltage levels appropriately.
The circuit in Fig. 12.1 operates as follows:
• The transistor T1 operates in a linear region where the emitter current, Ie, (output
current) is controlled by the base current, Ib, and the gain of the transistor, β.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 145
T. T. Tran, High-Speed System and Analog Input/Output Design,
https://doi.org/10.1007/978-3-031-04954-5_12
146 12 Power Supply Design Considerations
Ic Ie
DC Input T1 Regulated Output
R1
Ib Error Amp
C2
-
R2
Reference Voltage
R3
I e ¼ I c þ I b, ð12:1Þ
I c ¼ I b β: ð12:2Þ
Ie ¼ I bβ þ Ib: ð12:3Þ
• If the output voltage drops due to higher current load, the error amplifier config-
ured as a negative feedback circuit compares the Regulated Output divided by the
resistors R1 and R2 to the Reference Voltage and drives higher base current Ib to
maintain regulation. As shown in Eq. (12.3), the output current Ie is increased
with the increase in the base current Ib.
• If the output voltage increases due to lighter current load, the error amplifier sees
more negative voltage at the input and lowers the base current. This leads to lower
output current and again the system maintains regulation.
Like any other feedback system, if there are changes in the input voltage and the
load current, the system takes some time to stabilize, and this time typically is
specified in the component datasheet under the transient response section. The
major issue with linear regulator is the power dissipation across the transistor T1
for high output current applications. The power dissipation is
PT1 ¼ ð12 5Þ 1 ¼ 7 W:
This power dissipation generates a lot of heat and increases the device operating
temperature to the point where heatsink is required to keep the device temperature
12.1 Power Supply Architectures 147
SWITCHING
+5V DC INPUT POWER
+3.3V I/O
SUPPLY
(3.3V)
SWITCHING
POWER
+1.2V CORE
SUPPLY
(1.2V)
DSP
LINEAR
POWER
+3.3V Analog
SUPPLY
(3.3V)
ADC, DAC,
ANALOG
DIGITAL I/O
VIDEO,
AUDIO
under the maximum allowable limits. As the current requirement increases with
higher performance CPU/DSP and the system becomes smaller and smaller, it is no
longer practical to use the linear regulators to generate all the supply voltages. In this
case, it is best to use switching power supplies for the main power rails and linear
regulators to provide clean low noise supplies to the noise-sensitive circuits such as
analog and mixed analog/digital data converter circuits as shown in Fig. 12.2.
Design Example 12.1 Let us design a low noise high ripple rejection linear regulator
to provide a þ5 V to the audio circuit assuming þ12 V is available on the board.
Design steps:
• Audio circuits are very sensitive to low-frequency noise, so it is best to select a
regulator with a high power supply rejection ratio and with an external adjust pin
to allow for additional decoupling. So, let us use LM317.
• As shown in the LM317 datasheet [1], this device has a Ripple Rejection
specification of 62 dB minimum when using a 10 μF capacitor to decouple the
Adjust pin.
• Figure 12.3 shows the complete LM317 circuit as recommended in the datasheet.
• In Fig. 12.3, C1 prevents high-frequency noise from affecting the LM317 perfor-
mance. D1 and D2 are diodes needed to discharge the currents in C2 and C3 to
avoid these currents discharging into the output of LM317 during powering up
and down of the regulator. These diodes are reverse based on normal operation.
C2 is required to get better Ripple Rejection specification and C3 is a typical
148 12 Power Supply Design Considerations
D1 1N4002
U1 LM317
+12V IN +5V ANALOG
IN OUT
ADJ
D2 1N4002
C1 100n R1 240 C3 1u
C2 10u R2
decoupling capacitor. Keep in mind that C3 does not need to be a large decoupling
capacitor because LM317 is already doing a good job rejecting low-frequency
noise.
• R1 is fixed at 240 Ω and R2 is to be calculated as follows. From the LM317 datasheet,
R
V out ¼ V ref 1 þ 2 þ I adj R2 , ð12:5Þ
R1
For this application, it is acceptable to neglect 50 μA and solve for R2. Therefore,
R
5 ¼ 1:25 1 þ 2 ,
240
R2 ¼ 720 Ω. Put the complete circuit shown in Fig. 12.3 in the circuit simulator
[2] and the results are shown in Fig. 12.4 where the output is regulated at þ5 V
when the input is þ12 V.
As indicated earlier, for high current consumption circuits, it is best to go with
switching power supply architecture. Because it provides much better power effi-
ciency and lower power dissipation as compared to linear power supply. However,
this architecture generates excessive output switching noise that can degrade system
performance and cause EMI failures if designers are not carefully controlling the
switching noise by applying proper design, PCB layout, and isolation techniques.
The two types of switching power supplies are “buck” and “boost” converters.
The buck converter requires the input voltage to be higher than the output voltage; it
12.1 Power Supply Architectures 149
Error Amp
- 4 Pulse Width Modulator (PWM)
+
+
5 6
R5
R6
is also known as a step-down converter. And the Boost converter generates an output
that is higher than the input voltage; this is also known as a step-up converter. For
high-speed systems, as many different voltages are required and these voltages can
easily and economically be derived from the highest input voltage, a buck converter
is preferred. In a buck converter as shown in Fig. 12.5, the three main stages are:
(1) the power transistor stage; (2) the error amplifier stage; and (3) the pulse width
modulator or PWM stage. The converter operates as follows:
• The power transistor T2 operates in a saturation region where the transistor is
being driven fully on or off. When the transistor is on, there is only a small
resistance, Rdson in milli-ohms range, in series with the input and the output LC
filter. This leads to very low power dissipation and high efficiency.
• The emitter output of the T2 transistor is a digital waveform with a variable duty
cycle controlled by the PWM circuit. This emitter output is filtered by the L1 and
C1 and the output of the filter is a regulated DC output with some switching noise
modulated on it.
150 12 Power Supply Design Considerations
• The regulated DC output is fed back to the error amplifier circuit through the
resistor divider R4 and R5. This is a negative feedback loop so if the DC voltage
increases, the error amplifier output will be driven more negative. If the DC
voltage decreases, the error amplifier output will go more positive.
• The output of the error amplifier is an input to the PWM stage. This PWM
compares the error amp voltage to a sawtooth waveform. If the error amp voltage
increases, lower regulated DC output voltage, the PWM generates a higher duty
cycle waveform to drive the power transistor. This leads to an increase in the
regulated output voltage to maintain regulation. If the error amp voltage
decreases, higher regulated DC output voltage, the PWM generates a lower
duty cycle waveform to drive the power transistor. This leads to a decrease in
the regulated output voltage to maintain regulation.
• The regulated output voltage is V out ¼ V dc TTon , where Ton is the high time and
T is the period.
Figure 12.6 demonstrates how the system compensates for higher output voltage.
Higher output voltage leads to lower duty cycle signal and therefore lowers the
output voltage back to the regulated level.
Figure 12.7 shows that the duty cycle of the signal increases when there is a
decrease in the output voltage. This again forces the system back into regulation.
Design Example 12.2 Let us design a buck converter power supply for a DSP
system that has the design specifications shown in Table 12.1 [4].
Refer to [4] to calculate all the component values shown in Fig. 12.8.
5 Voltage at 4
3
12.1 Power Supply Architectures 151
5 Voltage at 4
VCC_12V
C1 C2a C2b
+ R3 R20
100K 1% 100K, 1% 22uF 25V 22uF 25V
9
5
6
7
8
25uF C7
U54A
S1 D1
S2 D2
S3 D3
D4
PAD
100nF 1 16 4 Q1
EN HDRV G
2 15 Si7860DP
FB SW C6 VCC_3.3V
1
2
3
3 14
COMP BOOT L1
4 13 100nF
VDD LDRV 800nH
5 12
UVLO BP
6 11 C9a C9b C9c
RT SS_SEL
9
5
6
7
8
R9 + + +
100K 1% 7 10
PAD
S1 D1
S2 D2
S3 D3
D4
ILIM PGOOD 4
PwrPd
Q2 47uF 47uF 47uF
C13 8 9 G
GND SYNC C12 Si7868DP
R6 100nF R7 TPS40195RGYR
17
1
2
3
100K 7.5K 1% 4.7uF
C4 R2
R11
8.2nF 2.32K 1% 10K, 1%
R10 C14
C5
49.9 1% 1nF
220pF
R5
2.2K 1%
The performance of this buck converter depends on the component values, the
component placements, and the layout. Here are some important points to remember.
• Always follow the manufacturer design guidelines and layout.
• Place the switching power supply circuit at a corner of the PCB away from the
rest of the system components.
• Keep all the switching current loops as small as possible. Refer to the manufac-
turer datasheet to figure out the possible current loops.
• In general, switching power supplies like Buck converter have y high current
switching characteristics and this generates harmonics as high as 10–100 times
the fundamental frequency. Proper shielding, decoupling, and isolating methods
outlined in the manufacturer datasheet and in this book must be taken into
consideration in order to increase the probability of success.
Designing power supply is not just designing the power supply itself but is necessary
to implement a system power to guarantee minimum noise and radiation to achieve
higher performance, higher reliability, and lower cost. Generally, systems with low
noise yield lower cost, because noisy systems tend to fail at a higher rate in
manufacturing. In many cases, system designers had to take an expensive approach
to solve issues by unnecessarily redesigning the mechanical chassis to add better
shielding to reduce noise. This type of costly activity could have been prevented by
applying good low noise design techniques right at the beginning.
It is understandable that redesigning electrical systems for low noise and EMI is
difficult for engineers who are not familiar with the latest high-speed design tech-
niques outlined in this book.
Assuming that the power supply itself was done properly as shown in the
previous sections, the power integrity depends on how far the CPU/DSP is placed
12.2 System Power Supply Architectural Considerations 153
Power Supply
14
+ 1
3
- 2
Load Cap
7
Current return path
Power Supply
14
+ 1
3
- 2
Load Cap
7
away from the power supply module and how well the CPU/DSP is being decoupled.
Figures 12.9 and 12.10 show two circuits, where Fig. 12.10, has a decoupling
capacitor close to the device. Assuming the same power supply trace inductances,
Fig. 12.9 has a larger dynamic current return path leading to a larger power supply
voltage droop and greater electromagnetic radiation. This may cause random system
failures that are difficult to resolve. Refer to Chap. 13 for details of how to implement
proper power supply decoupling capacitors to suppress noise while improving
system reliability and performance.
One of the most challenging tasks for system designers is determining an
acceptable noise level in a system. CPU/DSP data manuals clearly specify operating
conditions but cannot account for the dynamic nature of high-speed systems. This is
because the dynamic switching characteristics depend on how fast the system is
turning on or off and how the loads are changing during transient conditions. The
following are some of the important issues that must be addressed during the power
supply design process:
• Power supply transient response, such as load regulation, line regulation, power
supply ripple, power supply noise rejection, and power sequencing for multiple
rails.
• Power supply decoupling to ensure minimum voltage droop at the pins of the
CPU/DSP.
154 12 Power Supply Design Considerations
• Refer to the device data manual to get the maximum current consumption for the
Core supply. Many of the devices come with a power spreadsheet that can be used
to estimate the current consumption of a particular CPU/DSP operating condition.
• Select a regulator with at least two times the maximum Core current capability.
This provides an adequate margin to handle the dynamic current conditions.
• Be cautious with the current starvation condition. During startup, the surge
current may exceed the maximum limit of the regulator for a short period of
time. The selected regulator should have a soft-start capability to prevent thermal
or over-current shutdown conditions from occurring.
• The final design step for the Core voltage regulator is whether a heatsink is
required.
IO voltage regulator design depends on the external loads in the specific application.
For fast switching signals, the IO currents are supplied by the decoupling capacitors,
not by the regulator itself due to the parasitic inductance associated with the power
supply trace or plane. The dynamic current calculation will be shown in the
decoupling section. The following guidelines provide a conservative method to
design an IO voltage regulator for the DSP itself. It should be noted that this method
applies to the DSP power alone as opposed to the entire system.
• Count the number of outputs from the CPU/DSP. All GPIOs should be consid-
ered as outputs.
• Multiply the number of outputs by the source current specified in the data manual.
• Add the total source current using the maximum IO current consumption spec-
ified in the data manual.
• Then, multiply the result by 2 to provide a 100% margin.
• Due to transmission line effects, IO current may surge during switching but this
condition will be absorbed by the local CPU/DSP decoupling capacitors.
• The final step is to determine if heatsink is required or not.
156 12 Power Supply Design Considerations
Once designers complete the power supply architecture for a particular CPU/
DSP, the next step is to determine if the CPU/DSP requires sequencing. The supply
rails may be sequenced, for example, to ramp up the Core before the IO or vice versa.
Proper sequencing is necessary to avoid internal contention.
• Improper reset during power-up. Reset must be asserted longer than the minimum
reset pulse specified in the data manual.
• Core and IO not coming up within the specified time limits. Typically, DSPs do
not require a power sequence but there is a time limit for one supply rail to be on
while the other is off.
• Improper reset of the JTAG emulation port. For example, TRST needs to be
stable and low. Excessive noise coupled with this signal may cause a startup
problem or bus contention.
• Boot mode configuration pins are not being driven to proper states before
releasing reset. Refer to the device data manual to make sure that the configura-
tion pins have proper pull-ups and pull-downs and these pins have reached a
stable logic level before releasing reset.
Figures 12.11 and 12.12 show two CPU/DSP power supply architectures [5], in
Fig. 12.11, the Core and IO rails are powered up synchronously and in Fig. 12.12 the
Core supply rail is ramped up before the IO. Refer to [6] to obtain more details about
power management architectures.
The number of power supply rails required for CPU/DSP is increasing constantly as
more and more peripherals are being integrated. And managing these rails during
powering up or down of the DSP is a very difficult task. Typically, a high-
performance CPU/DSP consists of at least three power supply rails, þ1.8 V for
DDR2, þ3.3 V for data converters, þ1.2 V for core. The internal logic has many
voltage translations to enable all the blocks communicating to each other. During
power-up, if one power supply rail goes up before another for some period, the
internal logic can get to an unknown state which can cause internal bus contention
and the system to go unstable. Designers must refer to the CPU/DSP datasheet and
design in the power sequence if it is required. The problem is that the power supply
sequence circuits shown in Figs. 12.11 and 12.12 can only guarantee the power-up
sequence of the power supply itself, not the whole system. This is because the
decoupling capacitors used around the CPU/DSP as shown in Fig. 12.10 affect the
time it takes for the power supply to ramp up to the final operating voltage. This time
can be calculated as follows:
dV
I power ¼ Cdecoupling ,
dt
dV
dt ¼ C decoupling , ð12:6Þ
I power
where Cdecoupling is the total decoupling capacitance, Ipower is the current sourced
from the power supply, dV is the change in voltage, and dt is the time it takes to get to
the dV value.
As shown in Eq. (12.6), for a given power supply, the time it takes to reach the
final voltage level depends on the total decoupling capacitance. So, to guarantee a
particular sequence, it is very important for designers to do the following:
• Refer to the DSP datasheet and determine whether power sequencing is
necessary.
• If the selected DSP requires a power-up sequence, then use a topology like the
one shown in Fig. 12.12 to develop the power supply system.
158 12 Power Supply Design Considerations
• Use Eq. (12.6) and calculate the ramp time for each power supply rail and verify
that the power-up sequence was achieved at a system level with all the decoupling
capacitors installed on the board.
• Use a current probe and measure the power-up currents (Core and IO) to make
sure that there are no bus contentions. Keep in mind that all capacitors appear like
a short circuit to ground when they start from a zero-volt state, so the surge
current may be higher than expected during startup. Be sure to provide adequate
margin in the power supply design to avoid false-triggering the over-current
protection circuits. A good rule-of-thumb is adding a 50% margin to the maxi-
mum current consumed in the design.
• If there was excessive current consumption during power-up, designers need to
check the following.
– Is the system reset active? Reset signal should be asserted during this time.
– Is the power-up sequence, correct?
– Are any of the inputs left floating? All the inputs need to be pulled-up or down,
but they cannot be floating. Some of the DSPs have internal pull-ups or downs
integrated but not all of them, so make sure to check the data sheets and enable
the pull-ups and downs appropriately.
– Is the clock output of the DSP started running immediately after the power
supply reaches its operating range? If not, check the emulation reset signal to
make sure that this input is being driven correctly. If not, noise can couple to
this input and randomly put the device in test modes.
Figure 12.13 summarizes the steps to verify the power sequencing and to debug
the system startup over-current conditions.
Power
Sequencing
Verification
YES
Overcurrent?
NO
Correct YES
Power
YES Sequence?
Clock-Out
Present?
NO
YES
Clean
NO Reset?
Complete
Redesign
Power NO
Check Sequence Floating YES
Emulation Input
Pins (JTAG) RC Pins?
Filter
Reset Signal
NO
Recheck
Power Supply
Add Design
External
Pull-ups or Downs
12.4 Summary
As demonstrated in this chapter, selecting the right power supply architectures for
the high-speed system including surrounding analog/digital circuits and doing the
system floorplan design are the two most important critical tasks for designers to get
done first before getting started on the actual implementation. Good power integrity
is key for achieving a low noise and low EMI system design and here is a list of
recommendations to improve the probability of success:
• Develop a detailed system block diagram showing all the power supply require-
ments for all the components (CPU, DSP, ADC, DAC, video, audio, PLLs, DDR,
etc.)
• Apply the techniques described in this chapter and calculate the current require-
ments for all the blocks. It is recommended to add a 50% margin to the overall
current budget as this helps the system to better-handle dynamic situations.
• Highlight the noise sensitive circuits such as ADC, DAC, analog video/audio, and
PLL and isolate these circuits by using high power supply rejection linear
regulators if possible. Avoid powering these circuits with switching regulators.
• Do the floorplan design. Place the switching power supplies far away from analog
and high-speed circuits. The best place for noisy power supplies is at the corner of
the PCB.
• Select the power supply topologies and begin circuit implementation and layout.
Refer to Sect. 12.2.1 for power sequencing.
References
Poor power integrity is one of the most common root causes of system-related
problems. This is because there are too many things that could affect power delivery
to multiple devices on a system. These include DC resistance of PCB traces or
power/ground planes, AC impedance of PCB traces, power supply decoupling
around the DSP, and or other surrounding circuits such as DDR, clocks, and analog-
to-digital and digital-to-analog converters. One of the most challenging tasks for
designers is to determine the best decoupling techniques to achieve low noise and
high performance. In general, component manufacturers provide a conservative
recommendation for power supply decoupling, but in many cases, it is not practical
to follow this recommendation because of PCB space availability, power consump-
tion, EMI, or safety requirements. Also, component manufacturers always provide
development platforms for designers to evaluate and these platforms typically are a
lot larger than the actual design and are not required to be FCC certified, so copying
what was done on the development platform is not a guarantee that the design will be
successful. This chapter will discuss five important topics for designers: (1) DC
resistance of traces; (2) AC impedance; (3) a general rule-of-thumb decoupling
method; (4) an analytic decoupling method; and (5) how to make design tradeoffs
to achieve the best noise performance possible.
Once designers select and design a power supply for the DSP, the next step is to
determine the decoupling capacitors needed to ensure that the power supply droop
under all dynamic operating conditions is lower than the specified limits. For
example, a 5% tolerance rating on a 3.3 V IO supply requires the ripple to be less
than 165 mV. Let us first consider the situation where no decoupling capacitor is
used as shown in Fig. 13.1.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 161
T. T. Tran, High-Speed System and Analog Input/Output Design,
https://doi.org/10.1007/978-3-031-04954-5_13
162 13 Power Integrity
U19A
14
+ 1
3
- 2
Load Cap
7
14081/SO
In Fig. 13.1, the DSP labeled U19A is driving a capacitive load and is switching
at a fast rate. Now, let us assume that the regulator is placed 5 in. away from the DSP
and is routed with a 5-mil trace to the DSP. During fast switching, the power supply
trace becomes an open circuit because of the parasitic inductance associated with the
trace. This generates a large voltage droop at the pin of the DSP which can be
estimated as follows [1].
dI
Droop ¼ L Max , ð13:1Þ
dt
dI 1:52ΔV
Max ¼ C, ð13:2Þ
dt ðT r Þ2
This example demonstrates that for a 5-in. trace, 2 ns signal, 50 pF load, and 3.3 V
IO, the maximum power supply droop is 3.76 V. This level of droop is certain to
cause random system failures. To compensate, decoupling capacitors are placed
close to the DSP to provide the required charge during switching. What is the best
method to filter the noise from the DSP system? Noise characteristics differ so much
from system to system that no one method guarantees low noise and low radiation
for all cases. However, designers can apply best practices outlined here to minimize
the noise and to improve the probability of success. Before going into the decoupling
techniques, it is important to understand the characteristics of the common compo-
nents (capacitors, inductors, and ferrite beads) being used to filter out the power
supply noise.
The key specification of a capacitor used for decoupling is the self-resonant fre-
quency. The capacitor remains capacitive below and starts to appear as an inductor
above this frequency. Here is a series equivalent circuit representing the capacitor
(Fig. 13.2).
The series equivalent circuit for a capacitor has three different components:
equivalent series resistance (ESR), equivalent series inductance (ESL), and the
capacitance itself. The self-resonant frequency happens at the point where the
impedance of the capacitor, C, is equal to the impedance of the inductor, L.
1
Z C , capacitor ¼ ,
ωC
ESR
ESL
CAP-
164 13 Power Integrity
Z L , inductor ¼ ωL,
where L is inductance.
At resonance, ZL is equal to ZC or
1
¼ ωL,
ωC
1
ω2 ¼ ,
LC
1
ω ¼ pffiffiffiffiffiffi ,
LC
where ω ¼ 2πf.
Therefore, the self-resonant frequency is
1
fR ¼ pffiffiffiffiffiffi : ð13:3Þ
2π LC
The inductor also has a self-resonant frequency. The inductor remains inductive
below and starts to appear as a capacitor above this frequency. Here is a series
equivalent circuit of the inductor (Fig. 13.4):
The formula for calculating the resonant frequency of an inductor is the same as
for a capacitor.
166 13 Power Integrity
DCR
IND-
1
fR ¼ pffiffiffiffiffiffi :
2π LC
Figure 13.5 shows the inductor frequency response. For a particular inductor, the
impedance increases with frequency and reaches the highest impedance point at the
resonant frequency, fR. For frequencies above the resonant frequency, the impedance
of the inductor is dominated by the parasitic capacitor, C, and this causes the
impedance to decrease with frequency. It is recommended to operate in the inductive
region of the curve as this region guarantees a close to ideal impedance response of
the inductor. Like capacitors, there are different types of inductors and the two main
ones are air core and magnetic core. Air core is the coil with air or insulating core and
magnetic core is the coil wrapped around magnetic materials such as iron and ferrite.
Inductors are commonly being used in RF and high-power circuits but are rarely
being designed in high-speed DSP systems. Because it is better and lower cost to use
ferrite beads to isolate and filter the noise in DSP systems.
Here are general rules for using inductors to filter noise in a DSP system:
• Inductors are expensive and are sensitive to noise. Depending on the switching
speed of the signals propagating through it, an inductor can also generate and
radiate noise.
13.1 Power Supply Decoupling Techniques 167
Ferrite beads have electrical characteristics that are similar to ideal inductors. The
key difference is that the ferrite bead has no or negligible parasitic capacitance until
the frequency reaches GHz range as shown in Fig. 13.7. So, the ferrite bead behaves
like an inductor over a wide frequency range. As shown in Fig. 13.6, ferrite bead
always has a small DC resistance so review the specifications carefully and select a
component that has the right AC impedance and low IR drop for the design. The
ferrite bead generally performs well at frequencies higher than 30 MHz. It is
commonly used to isolate power supplies and noise-sensitive circuits such clocks,
video, and audio CODECs.
Some manufacturers provide free design tools to help engineers selecting and
simulating the ferrite bead circuits. Figure 13.7 shows an impedance response
provided by one of the ferrite bead design tools [3].
Two important parameters to select a ferrite bead are DC resistance and AC
impedance at a given frequency. In general, assuming no issues with PCB space, it is
best to select a device with the lowest DC resistance and highest impedance at the
operating frequency. This yields the lowest IR drop across the ferrite bead while
providing the highest noise rejection.
DCR
Ferrite-
168 13 Power Integrity
The ideal way to decouple the supply noise is to have one capacitor between each of
the power and ground pins of the DSP. Normally, this is physically not practical
because the DSP package area is too small. So, designers must compromise by
reducing the number of decoupling capacitors to fit in the general area underneath or
above the DSP. Refer to the device data manual for a recommended method. But in
general, here are the important considerations for decoupling:
• Add as many decoupling capacitors as space allows but do not put more capac-
itors than the DSP power pins.
• Add 8 bulk capacitors, 4 for Core and 4 for IO supplies. Place each bulk capacitor
at each region of the DSP, with the region being defined as an edge or a corner of
the DSP. Bulk capacitors act as a low-frequency noise filter and a charge storage
device for the smaller decoupling capacitors. The use of four bulk capacitors is
preferable to one large discrete component because this guarantees a shorter
recharge path and a lower parasitic inductance path between the bulk and the
decoupling capacitors.
• Keep in mind that all capacitors have equivalent series inductance (ESL) and
equivalent series resistance (ESR). ESL and ESR reduce filtering effectiveness.
So, select the smallest surface mount capacitors that can be used.
Figure 13.8 demonstrates a good scheme for decoupling a particular DSP. Refer
to the device data manual to find more details. As shown in this figure, 0.01 μF
ceramic capacitors are used for the decoupling capacitors and 10 μF tantalum
capacitors are used as low-frequency filtering components. Typically, designers
must go back and change the values to optimize them for their applications. A
good approach is changing the capacitor values to achieve less than 50 mV power
supply ripple for the IO rail and less than 20 mV for the Core rail. Another good rule
is to use ceramic capacitors for high-frequency decoupling and tantalum capacitors
for low-frequency filtering. This is because tantalum capacitors come in higher
values than ceramic capacitors as shown in Table 13.1. These two types of capacitors
13.1 Power Supply Decoupling Techniques 169
provide the low ESR and ESL which are needed for low noise and low EMI designs
operating over a wide range of voltages, temperatures, and frequencies.
These general rules are only applied to the digital Core and IO power pins of the
DSP. PLL and other analog power pins need to include better filtering schemes
(Pi filters and or linear voltage regulators) to prevent the low-and high-frequency
noise from affecting the performance of these circuits. Refer to the PLL (Chap. 6)
and data converters (Chap. 7) for more details.
Another method of decoupling power supply noise from a DSP system is calculating
the total capacitance required to keep the power supply ripple under a certain limit.
Similar to the general rules of decoupling, this method provides a starting value that
must typically be optimized. The large ball grid array (BGA) package typically used
for DSPs behaves like a PCB itself with long traces routing from the die out to the
balls. These traces can generate interference and are susceptible to crosstalk, power
supply droop, and other electrical noise. The asymmetry analytical decoupling
technique begins by dividing the DSP into 4 regions and then decoupling each
region separately. Providing fewer decoupling capacitors in the low-speed section
leads to a uniform reduction in noise and electromagnetic radiation. The rules for this
decoupling technique are:
• Divide the DSP package into 4 regions by drawing 2 diagonal lines across the
4 corners of the DSP as shown in Fig. 13.10. Be sure to keep a group of signals
together, for example, keeping all the DDR signals in one region. The boundaries
do not have to be diagonally divided as shown in Fig. 13.10.
170 13 Power Integrity
2 I CoreMax
I CRegion ¼ M: ð13:4Þ
N
4 I CoreTyp
I CRegion ¼ M: ð13:5Þ
N
• Calculate the total decoupling capacitance for the region by applying Eq. (13.7)
dV Core
I CRegion ¼ CCore , ð13:6Þ
dt
dt
C Core ¼ I CRegion , ð13:7Þ
dV Core
where dt is the fastest rise time in the region and dV is the maximum ripple
allowed for the Core voltage, assuming 10 mV ripple.
• Now, calculate the total bulk capacitance for the region by multiplying the total
decoupling capacitance by 40. The rule recommended for bulk capacitance is at
least 10 times the total decoupling capacitance [4]. Use one bulk capacitor per
region to minimize the parasitic inductance between the bulk and the decoupling
capacitors.
• To figure out the number of decoupling capacitors, review the PCB area to see
how many capacitors can be placed within 0.5 in. or 1.25 cm of the power supply
pins. It is preferable to use smaller size capacitors in order to have more capacitors
in a region. If the DSP package is a full grid array as shown in Fig. 13.9, escape all
the signals uniformly out in 4 different directions (Northwest, Northeast, South-
west, and Southeast) to create two lanes across the package. Now, use these two
lanes to place the capacitors near the DSP Core and IO power pins. To find the
decoupling capacitor value, divide the total capacitance by the number of capac-
itors allowed for the region. It is good to select a capacitor with a self-resonant
frequency equal to the maximum frequency of the particular region. For example,
if the SDRAM port runs at 100 MHz, then add at least one capacitor with the
resonant frequency of 100 MHz in this region. The other capacitors within each
region should have the highest possible resonant frequency. This helps mitigate
EMI over a wide frequency range.
13.1 Power Supply Decoupling Techniques 171
• Divide the DSP package into 4 regions by drawing 2 diagonal lines across the
4 corners of the DSP as shown in Fig. 13.10. Be sure to keep the signal groups
together as indicated in Core voltage decoupling section.
• Count the number of IO voltage, inputs and outputs of each region.
• Conservatively estimate the IO current consumption of the DSP itself in the
region, IIORegion, as shown in Eq. (13.8) by taking the maximum device current
specification, IIO, divided by the total number of IO voltage pins, K, and multiplied
by the number of IO voltage pins, J, within a region. Do not need to add margin to
the IO current consumption here as the margin will be added in the next step.
I IO
I IORegion ¼ J: ð13:8Þ
K
• The total IO current is not equal to the IO current sourcing and sinking defined in
the DSP datasheet. Most of the total IO current depends on the external loads, for
example, resistive, capacitive, or transmission line. In this design, let us add a lot
of margins by assuming a worst case scenario where all the IOs are outputs and
172 13 Power Integrity
are loaded with transmission lines. In this case, each output current, IIOTrans, is
equal to the output voltage divided by the characteristic impedance of the
transmission line, Zo, as shown in Eq. (13.9).
V IO
I IOTrans ¼ : ð13:9Þ
Zo
• In Eq. (13.10), the total IO current for the region is equal to the IO current of the
DSP itself plus the IO current driving the transmission lines.
V IO
I IOTotal ¼ I IORegion þ J : ð13:11Þ
Zo
• Calculate the total decoupling capacitance for the region by applying Eq. (13.13).
dV IO
I IOTotal ¼ C IO , ð13:12Þ
dt
dt
C IO ¼ I IOTotal , ð13:13Þ
dV IO
13.1 Power Supply Decoupling Techniques 173
where dt is the fastest rise time in the region and dV is the maximum ripple
allowed for the IO voltage, assuming 50 mV ripple.
• Now, calculate the total bulk capacitance for the region by multiplying the total
decoupling capacitance by 40. The rule recommended for bulk capacitance is at
least 10 times the total decoupling capacitance [4]. Use one bulk capacitor per
region to minimize the parasitic inductance between the bulk and the decoupling
capacitors.
• To figure out the number of decoupling capacitors, review the PC board area to
see how many capacitors can be placed within 0.5 in. of the pins. If the DSP
package being used is a full ball grid array, then apply the same technique
outlined in the Core decoupling section to create two lanes for placing the
decoupling capacitors nearby the power pins. To find the decoupling capacitor
value, take the total capacitance value just calculated and divide it by the number
of capacitors allowed for the region. It is good to select a capacitor with a self-
resonant frequency equal to the maximum frequency of the particular region. For
example, if the video port IO runs at 100 MHz, then add at least one capacitor
with the resonant frequency of 100 MHz at this region. For the rest of the
capacitors within that region, select the highest possible resonant frequency
value.
This analytical decoupling method provides designers with a good starting point.
As mentioned earlier, designers need to optimize the decoupling capacitors to ensure
low noise and EMI during the board characterization process. The following exam-
ple shows how this process can be applied to a typical design.
Example 13.1 Let us use a 289-pin BGA (Ball Grid Array) DSP [5]. Now, divide
the 289-pin package into 4 regions by drawing two symmetry lines across the part as
shown in Fig. 13.11. Then count the number of Core voltage pins, I/O voltage pins
and signals, not including the ground pins, in each region. Also, pay special attention
to the critical sections, such as external memory interface fast (EMIFF), phase-
locked loop (PLL), and other high-speed serial/parallel ports. Assume all IOs outputs
driving a 60-Ω transmission line and all the signal groups falling within a region.
These are reasonable assumptions but there are cases where the boundaries of the
regions had to be altered in order to keep the signal groups together. For the PLL and
other analog power pins, the decoupling schemes are covered in Chap. 11.
The next step is to conservatively estimate the switching current requirements for
each Region.
Table 13.2 shows the calculations of switching currents for all 4 Regions. The
conservative assumptions used to calculate the capacitors in Table 13.2 are:
– Maximum Core current ¼ Typical Current 2 plus 100% margin
¼ 170 mA 2 2 ¼ 680 mA.
– Device IO current ¼ Typical IO current 2 plus 100% margin
¼ 45 mA 2 2 ¼ 180 mA.
– Total IO Current ¼ Device IO Current plus IO Current Driving a Transmission Line.
174 13 Power Integrity
A1 A21
– Assuming that half of the inputs and outputs in the Region switching at the same
time driving 60-Ω transmission lines, this is a conservative assumption since
many of the signals in the 4 regions are too slow to be considered as transmission
lines.
Since the Core and I/O voltage operate at different frequencies, they require
separate decoupling calculations. The following shows the steps needed to calculate
and select the decoupling capacitors for both Core and I/O supplies.
To find the decoupling capacitance, plug the peak current, the rise time, and the
maximum ripple voltage parameters into Eq. (13.7) and solve for C. It is acceptable
to assume that the maximum ripple voltage is 10 mV for Core and 50 mV for IO and
the typical rise time is 2 nS.
dt
CCore ¼ I CRegion
dV Core
Use the capacitor Eq. (13.13) to calculate the total capacitance for the IO voltage
decoupling:
dt
C IO ¼ I IOTotal
dV IO
Now let us calculate the total capacitance required for each region.
13.1
Region Total peak core current, IIOTotal Device IO current, IIORegion transmission lines, IIOTrans plus IIOTrans
Region 1 680 mA 180 mA 3:3 69 mA þ 2.97 A ¼ 3 A
13 3 ¼ 157 mA 21 8 ¼ 69 mA 60 54 ¼ 2:97 A
Region 2 Same as 1, 157 mA 180 mA 3:3 3.3 A þ 34 mA ¼ 3.3 A
21 4 ¼ 34 mA 60 60 ¼ 3:3 A
Region 3 Same as 1, 157 mA 180 mA 3:3 3.3 A þ 26 mA ¼ 3.3 A
21 3 ¼ 26 mA 60 59 ¼ 3:3 A
Region 4 680 mA 180 mA 3:3 3 A þ 51 mA ¼ 3.1 A
13 4 ¼ 209 mA 21 6 ¼ 51 mA 60 55 ¼ 3 A
175
176 13 Power Integrity
Region 1
ð2 nSÞ
Total Core capacitance, C Core ¼ 157 mA ¼ 0:03 μF,
ð10 mVÞ
ð2 nSÞ
Total I=O capacitance, C IO ¼3 A ¼ 0:08 μF:
ð50 mVÞ
There are 3 Core voltage pins operating at 150 MHz (CPU frequency) and 8 I/O
voltage pins operating at 40 MHz (EMIFS frequency). It would be desirable to use
multiple capacitors for the multiple supply pins, but there is a physical limitation due
to the limited space available around the device. For the DSP [5] package, there is
enough board space to place about 4 or 5 capacitors per region. In this case, select
two capacitors with a total capacitance of around 0.03 μF. At least one of the
capacitors should have a self-resonant frequency of around 150 MHz to decouple
the Core voltage pins in region 1. Then, select three capacitors with a total capac-
itance of around 0.08 μF with at least one of the capacitors having the self-resonant
frequency of around 75 MHz to decouple the I/O voltage pins in region 1.
In summary, for Core voltage in region 1, use two 0.022 μF (0.044 μF total)
ceramic capacitors and, for the I/O voltage, use three 0.033 μF (0.099 μF total)
ceramic capacitors.
The next step is calculating the bulk capacitors for both Core and IO. Bulk
capacitor placement is not as critical as decoupling capacitor placement. But bulk
capacitors are needed to filter the low-frequency ripple typically generated by
switching power supply and to recharge the decoupling capacitors.
A rule-of-thumb is to select bulk capacitors with at least ten times the total
decoupling capacitance. Let us use 40 times to be conservative. For the Core voltage,
As mentioned earlier in this chapter, the best technique is adding 4 bulk capac-
itors to 4 regions of the DSP and the smallest tantalum capacitor available is 4.7 μF.
In this case, select 4.7 μF tantalum bulk capacitors for both IO and Core voltages in
region 1.
In summary, Fig. 13.12 shows the complete schematic diagram for decoupling
Region 1 of the DSP. Next is to repeat the same steps for regions 2, 3, and 4.
Region 2
ð2 nSÞ
Total Core capacitance, CCore ¼ 157 mA ¼ 0:03 μF
ð10 mVÞ
ð2 nSÞ
Total I=O capacitance, CIO ¼ 3:3 A ¼ 0:13 μF:
ð50 mVÞ
13.1 Power Supply Decoupling Techniques 177
0.033uF REGION 2
0.022uF
REGION 1
0.033uF R REGION 3
0.022uF
REGION 4
0.033uF
4.7uF
There are 3 Core voltage pins operating at 150 MHz (CPU frequency) and 4 I/O
voltage pins operating at 40 MHz (EMIFS frequency). For the DSP [5] package,
there is enough board space to place about 4 or 5 capacitors per region. In this case,
select two capacitors with a total capacitance of around 0.03 μF. At least one of the
capacitors should have a self-resonant frequency of around 150 MHz to decouple the
Core voltage pins in region 2. Then, select three capacitors with a total capacitance of
around 0.13 μF with at least one of the capacitors having the self-resonant frequency
of around 75 MHz to decouple the I/O voltage pins in region 2.
In summary, for Core voltage in region 2, use two 0.022 μF (0.044 μF total)
ceramic capacitors and for the I/O voltage, use three 0.047 μF (0.14 μF total) ceramic
capacitors.
The next step is calculating the bulk capacitors for both Core and IO. A rule-of-
thumb is to select bulk capacitors with at least ten times the total decoupling
capacitance. Let us use 40 times to be conservative.
For the Core voltage,
In this case, select 4.7 μF tantalum capacitor for the Core voltage and 6.8 μF
tantalum capacitor for the IO voltage in region 2. Figure 13.13 shows the complete
decoupling schematic of region 2.
178 13 Power Integrity
6.8uF
4.7uF
0.047uF
0.022uF
0.047uF
0.022uF
0.047uF
2 decoupling capacitors
REGION 2
REGION 1 R REGION 3
REGION 4
Region 3
Region 3 has the same Core and IO currents as region 2. Therefore, the Core and IO
capacitors have the same values as the capacitors in region 2; there are two 0.022 μF
capacitors and three 0.047 μF capacitors. And for the bulk capacitors, one 4.7 μF
tantalum capacitor is for the Core voltage and one 6.8 μF tantalum capacitor is for the
IO voltage as shown in Fig. 13.14.
Region 4
ð2 nSÞ
Total Core capacitance, C Core ¼ 209 mA ¼ 0:042 μF,
ð10 mVÞ
ð2 nSÞ
Total I=O capacitance, CIO ¼ 3:1 A ¼ 0:124 μF:
ð50 mVÞ
There are 4 Core voltage pins operating at 150 MHz (CPU frequency) and 6 I/O
voltage pins operating at 40 MHz (EMIFS frequency). For the DSP [5] package,
there is enough board space to place about 4 or 5 capacitors per region. In this case,
select two capacitors with a total capacitance of around 0.042 μF. At least one of the
13.1 Power Supply Decoupling Techniques 179
REGION 2 0.047uF
0.022uF
0.022uF
REGION 4
0.047uF
4.7uF
capacitors should have a self-resonant frequency of around 150 MHz to decouple the
Core voltage pins in region 4. Then, select three capacitors with a total capacitance of
around 0.124 μF with at least one of the capacitors having the self-resonant fre-
quency of around 75 MHz to decouple the I/O voltage pins in region 4.
In summary, for Core voltage in region 4, use two 0.027 μF (0.054 μF total)
ceramic capacitors and for the I/O voltage, use three 0.047 μF (0.14 μF total) ceramic
capacitors.
The next step is calculating the bulk capacitors for both Core and IO. A rule-of-
thumb is to select bulk capacitors with at least ten times the total decoupling
capacitance. Let us use 40 times to be conservative.
For the Core voltage,
In this case, select 4.7 μF tantalum capacitor for the Core voltage and 6.8 μF
tantalum capacitor for the IO voltage in region 4 as shown in Fig. 13.15.
Table 13.3 shows a summary of all the capacitors calculated for the 4 regions of
the DSP and Fig. 13.16 shows the complete schematic.
180 13 Power Integrity
REGION 1 R REGION 3
REGION 4
0.047uF
0.027uF
0.047uF
0.027uF
0.047uF
4.7uF
6.8uF
NOTE: SHADED COMPONENTS ARE FOR IO POWER AND
NON-SHADED COMPONENTS ARE FOR CORE POWER
Another method widely used in the industry is target impedance method which is the
maximum impedance allowed while maintaining the ripple voltage on a power
supply rail below the ripple specifications. The equation for target impedance, Ztar, is
V dd Ripple
Z tar ¼ , ð13:14Þ
I transient
where Vdd is the supply voltage rail, Ripple is the maximum percentage of the ripple
voltage, and Itransient is the worst case transient current.
As discussed, the function of all the power supply decoupling capacitors is to
create a low AC current path to ground to prevent noise from affecting the system
13.1 Power Supply Decoupling Techniques 181
6.8uF
4.7uF
0.047uF
0.022uF
0.047uF
0.022uF
0.047uF
schematic
4.7uF 6.8uF
0.022uF 0.022uF
REGION 1
0.033uF R REGION 3 0.047uF
0.022uF 0.022uF
REGION 4
0.033uF 0.047uF
4.7uF 4.7uF
0.047uF
0.027uF
0.047uF
0.027uF
0.047uF
4.7uF
6.8uF
NOTE: SHADED COMPONENTS ARE FOR IO POWER AND
NON-SHADED COMPONENTS ARE FOR CORE POWER
It is important to place all the decoupling capacitors as close as possible to the pins,
no more than 0.25 in. in most cases. The bulk capacitors should be placed as close as
possible to the decoupling capacitors. This reduces the trace lengths, reducing the
current loops and in turn lowering radiation while minimizing parasitic inductance.
The best strategy is placing the decoupling capacitors on the bottom of the PCB right
underneath the device being decoupled, and the bulk capacitors on top or bottom of
the PCB close to the decoupling capacitors. Also, using via-in-pads to connect the
capacitors to ground and power planes is the best way to keep parasitic inductances
as low as possible. Due to costs, via-in-pad method has rarely been implemented in
commercial products.
182 13 Power Integrity
In summary, there should be two bulk capacitors per region, one for Core and one
for IO, and as many decoupling capacitors as space allows. Figure 13.18 shows a
good example of the capacitor’s placement on the bottom side of the PCB. The Core
decoupling capacitors and four large bulk capacitors are placed on the interior of the
BGA package in the open space right under the DSP. The IO decoupling and bulk
capacitors are placed on the perimeter of the BGA package. This is possible because
this particular BGA package is not a full BGA package where all the balls are fully
populated on the bottom of the package.
If the package being used is a full BGA package, then it is necessary to route the
signals from the DSP package out to external circuits as demonstrated in Fig. 13.19
that creates two lanes underneath the DSP for decoupling capacitors. Now, use these
two lanes and populate as many decoupling capacitors as the lanes allow. In this
case, the bulk and some of the IO capacitors can be placed on the perimeter of the
DSP package. The recommended rules for creating the lanes are as follows:
• Power and ground pins need to be closest to the lanes. This allows the shortest
connection paths to the capacitors.
• The lanes do not have to be symmetrical as shown in Fig. 13.19.
• Designers can replace the lanes with capacitor islands underneath the DSP if this
allows placing the decoupling capacitors near the power pins. For example,
instead of having two lanes, designers may want to create many islands and
each island can hold one or more decoupling capacitors.
13.2 High-Frequency Noise Isolation 183
Core
Caps
IO
Caps
The decoupling methods described up to now filter noise locally at the DSP. There
are cases where the whole power supply plane for some critical sections needs to be
isolated. This may be required to prevent external noise from entering these sections
or to prevent noisy circuits such as oscillators from coupling onto the power plane.
The power supply plane is generally isolated using either Pi or T filters. A Pi filter is
constructed with two capacitors and one ferrite bead while a T filter requires one
capacitor and two ferrite beads. Each of these filters is commonly used in series with
the signals exiting and entering the system or the power supply to reduce the radiated
emissions. The pass-band of the filter must be calculated precisely to ensure that the
bandwidth is wide enough to pass the desired signals without degrading signal
quality, especially critical parameters such as rise and fall times and amplitude.
te te te te te te
xt xt xt xt xt xt
Via Via Via Via
te te te te
CAP
xt xt xt xt
te te te te te te
Signal Signal
xt xt xt xt xt xt
Via Via Via Via
te te te te
CAP
xt xt xt xt
te te te te te te
Signal Signal
xt xt xt xt xt xt
CAP
te te te te te te
Signal Signal
xt xt xt xt xt xt
Via Via Via Via
te te te te
CAP
xt xt xt xt
te te te te te te
Signal Signal
xt xt xt xt xt xt
Via Via Via Via
te te te te
CAP
xt xt xt xt
te te te te te te
xt xt xt xt xt xt
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Lz
, ð13:15Þ
2 C1
1
f C ¼ pffiffiffiffiffiffiffiffiffiffi : ð13:16Þ
π Lz C 1
C1 C2
7
DSP
1
Ferrite Bead, Z1 Ferrite Bead, Z2 Output
3
2
C1
7
DSP
Ferrite Bead
VDD
Power TI DSPA
14
+ 1
C1 C2 3
- 2
Load Cap
7
Let us design a Pi filter for a graphic controller’s Red, Green, and Blue (RGB) analog
signal outputs driving a computer monitor. Assuming that RGB signals have a
100 MHz analog bandwidth, calculate the filter components as follows.
Let fC ¼ 200 MHz. Setting the filter at 200 MHz provides a 100 MHz margin
(200 MHz minus 100 MHz) to make sure that the filter is not affecting the video
signal bandwidth.
The filter corner frequency is
1
f C ¼ pffiffiffiffiffiffiffiffiffiffi ¼ 200 MHz:
π Lz C 1
186 13 Power Integrity
Let us select a ferrite bead with 100 Ω impedance at 100 MHz and calculate Lz.
The impedance, Z, of the ferrite bead is
Z ¼ 2πfL ¼ 100 Ω,
100 100
Lz ¼ ¼ ¼ 0:16 μH:
2πf 2π 100 106
Now, calculate C1 by substituting Lz and fC into the Eq. (13.16) and solve for C1.
1
f C ¼ pffiffiffiffiffiffiffiffiffiffi ¼ 200 MHz,
π Lz C1
1
200 106 ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi,
π 0:16 106 C 1
C1 ¼ 15:8 pF:
Therefore, the Pi filter has two 15.8 pF capacitors and one 0.16 μH inductor.
Now, let us use an analog circuit simulator [7] to verify the design.
To match the source and load impedance, use Eq. (13.15) and calculate R1 and R2
values. In this case, R1 ¼ R2 ¼ 71 Ω for C1 ¼ 15.8 pF and Lz ¼ 0.16 μH.
Figure 13.23 shows the simulation results of the circuit model shown in
Fig. 13.24. In the pass-band from DC to 100 MHz, the circuit shows a 6 dB
attenuation. This is because the voltage divider is formed by the 71-Ω source resistor
and 71-Ω load resistor. In this case, the attenuation is
V F2
¼ 20 log 10 , ð13:17Þ
V F1
where
71
V F2 ¼ V : ð13:18Þ
71 þ 71 F1
Now, substitute Eq. (13.18) into Eq. (13.17) and solve to calculate the
attenuation.
71
Attenuation ¼ 20 log 10 ¼ 6 dB:
71 þ 71
AMPLITUDE
in dB
fc FREQUENCY
in Hz
VF1
VF2
R1 71 L1 160n
C2 15.8p
C1 15.8p
VG1
+
R2 71
Fig. 13.24 Pi filter circuit model for simulation
frequency of each pole is at 200 MHz and the Pi filter has 3 poles (2 capacitors and
1 ferrite bead), the combined corner frequency at 200 MHz has a 9 dB attenuation
as shown in the simulation.
For this special T filter assuming L1 equal to L2, the corner frequency of the 3-pole
filter [6] is
1
f C ¼ pffiffiffiffiffiffiffiffiffiffi : ð13:20Þ
π Lz C 1
Let us design a T filter for a graphic controller’s Red, Green, and Blue (RGB) analog
signal outputs driving a computer monitor. Assuming that RGB signals have a
100 MHz analog bandwidth, calculate the filter components as follows:
Let fC ¼ 200 MHz. Setting the filter at 200 MHz provides a 100 MHz margin
(200 MHz minus 100 MHz) to make sure that the filter is not affecting the video
signal bandwidth.
The filter corner frequency is
1
f C ¼ pffiffiffiffiffiffiffiffiffiffi ¼ 200 MHz:
π Lz C 1
Let us select a ferrite bead with 100 Ω impedance at 100 MHz and calculate Lz.
The impedance, Z, of the ferrite bead is
13.2 High-Frequency Noise Isolation 189
AMPLITUDE
in dB
fc FREQUENCY
in Hz
Z ¼ 2πfL ¼ 100 Ω,
100 100
Lz ¼ ¼ ¼ 0:16 μH:
2πf 2π 100 106
Now, calculate C1 by substituting Lz and fC into the Eq. (13.20) and solve for C1.
1
f C ¼ pffiffiffiffiffiffiffiffiffiffi ¼ 200 MHz,
π Lz C1
1
200 106 ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi,
π 0:16 106 C 1
C1 ¼ 15:8 pF:
Therefore, the T filter has two 0.16 μH ferrite beads and one 15.8 pF capacitor.
Now, let us use an analog circuit simulator [7] to verify the design.
To match the source and load impedance, use Eq. (13.18) and calculate R1 and R2
values. In this case, R1 ¼ R2 ¼ 71 Ω for C1 ¼ 15.8 pF and L1 ¼ L2 ¼ 0.16 μH.
Figure 13.27 shows the simulation results of the circuit model shown in
Fig. 13.28. In the pass-band from DC to 100 MHz, the circuit shows a 6 dB
attenuation. This is because the voltage divider is formed by the 71-Ω source resistor
and 71 Ω load resistor. In this case, the attenuation is
V F2
¼ 20 log 10 , ð13:21Þ
V F1
where
71
V F2 ¼ V : ð13:22Þ
71 þ 71 F1
190 13 Power Integrity
VF1
VF2
R1 71 L1 160n L2 160n
C1 15.8p
VG1
+
R2 71
Fig. 13.27 Pi filter circuit simulation results
Now, substitute Eq. (13.22) into Eq. (13.21) and solve to calculate the
attenuation.
71
Attenuation ¼ 20 log 10 ¼ 6 dB:
71 þ 71
13.3 Summary
Power supply decoupling and noise isolation techniques discussed in this chapter
provide practical design considerations and theoretical approaches to design the
optimum filters for noise isolation. Poor decoupling techniques are the number one
root cause of random DSP system failures. So, to improve the probability of design
success and to prevent random logic delete/remove line spacing failures caused by
excessive system noise, designers need to do the following:
• Apply the General Rules for decoupling or the Analytic Decoupling methods
described in this chapter.
• Use power integrity tool to simulate decoupling capacitors to make sure that the
decoupling method provides good low impedance paths to ground, less than
target impedance.
• Follow the guidelines shown in this chapter to place the decoupling capacitors
properly. Select the right components (ferrite beads, inductors, capacitors, or
resistors) for the design.
• Use Pi filters to isolate noise and place the capacitor as close to the connector as
possible. This minimizes the RF current loops while providing some ESD
protection.
• Use an analog simulator [7] and simulate the design to verify all the calculations
before going into layout.
References
4. O. Henry, Electromagnetic Compatibility Engineering (John Wiley and Sons, Hoboken, NJ,
2009)
5. Texas Instruments Inc., OMAP5910 Dual-Core Processor Data Manual (2002). http://focus.ti.
com/lit/ds/symlink/sm320c6713b-ep.pdf
6. K. Kenneth, Electromagnetic Compatibility Handbook (CRC Press, Boca Raton, FL, 2005)
7. Texas Instruments Inc., Spice-Based Analog Simulation Program (2008). http://focus.ti.com/
docs/toolsw/folders/print/tina-ti.html
Chapter 14
Printed Circuit Board (PCB) Layout
Once all the circuits have been designed and simulated, the next step is board layout.
This is one of the most critical steps in the development process because the
effectiveness of the circuits designed depends on where the components are placed
relative to the DSP/CPU pins and the traces that are being routed to those pins. Also,
the board layout has a big effect on noise, crosstalk, and transmission line effects so
optimizing the layout can minimize these effects. This chapter covers the printed
circuit board stackup, routing guidelines, and layout techniques for low noise and
EMI.
First, designers must determine the minimum number of PCB layers and then
configure the board stackup. Here are some general guidelines:
• Perform layout experiments and refer to the DSP/CPU reference design package
to find the minimum number of layers required to escape the signals out from the
device’s package. Typically, DDR layout guidelines dictate the number of PCB
layers to allow for escaping all the signals out from the DSP/CPU.
• Consider the need for high-speed signals to be shielded between the ground and
power planes.
• Are there buses, such as USB, Ethernet, and RapidIO, that require a tight
differential impedance specification? If so, designers need to follow the industry
guidelines to control the differential impedance of these buses.
• Does the PCB manufacture require a certain trace width and spacing? This
determines whether a trace can be routed between the balls of a small pitch
BGA package. For good signal integrity with minimum skin effect losses, keep
the trace width between 4 and 12 mils. A common choice is a 5-mil trace and
5-mil spacing.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 193
T. T. Tran, High-Speed System and Analog Input/Output Design,
https://doi.org/10.1007/978-3-031-04954-5_14
194 14 Printed Circuit Board (PCB) Layout
εr A
C pp ¼ k pF, ð14:1Þ
d
where k is 0.00885 m, εr, dielectric constant ¼ 4.1–4.7 for FR4 type PCB, A is area
of the power and ground planes in mm2, and d is the distance between the power and
ground planes in mm.
When using this topology, designers need to consider these points:
• As shown in the equation for Cpp, the distance, d, between the power and ground
planes determine the board capacitance. Reducing the distance increases the
capacitance and reduces high-frequency impedance. The limiting factor is how
closely the layers can be packed together while still maintaining the quality and
reliability of the design. Refer to the specifications from PCB manufacturers to
understand the minimum requirements between the layers.
• Route the high-speed signals on the planes next to the power and ground planes.
If possible, route all the high-speed signals next to the ground plane. If not, then it
is also acceptable to route them next to the power plane.
• In Fig. 14.1, the best routing layer is Layer 2 because it is next to the ground
plane. This provides optimal current return paths which help reduce radiation.
Therefore, the adjacent power and ground topology is recommended for DSP
systems operating at high frequency.
14.2 Microstrip and Stripline 195
NON-ADJACENT P/G
• The adjacent power and ground topology are not useable for DSP systems that
require many layers to route the signals out from the DSP and interface with other
circuits.
Figure 14.2 [1] shows a typical PCB stackup for the non-adjacent power and
ground topology. The power and ground planes are placed in Layer 5 and Layer
2, respectively. Layer 3 is best for routing high-speed traces while Layer 1, Layer
4, and Layer 6 are acceptable. As shown in the figure, each of the routing layers is
next to either a ground or power plane. Layer 3 is best because it is not only next to a
ground plane but is also guarded by a power plane below it. This scheme is best for
difficult-to-route DSP systems that do not operate at high frequency. One thing to
keep in mind is that board capacitance becomes important for systems operating
above 300 MHz [2].
Here are rules for doing non-adjacent board stackup design:
• For non-adjacent topology, the board capacitance, as shown in capacitance
equation Cpp, is low and the board impedance is high between the power and
ground planes. This is the opposite of what is needed for systems to have low
noise and low EMI.
• This topology requires more high-frequency decoupling capacitors to compensate
for the board characteristics.
Table 14.1 shows the advantages and disadvantages of the two main signal routing
technologies, Microstrip is shown in Fig. 14.3 where H is the height, W is the width
of the signal trace, W1 is the width including the outside edge, T is the thickness of
the trace, and εr is the dielectric constant. The Stripline is shown in Fig. 14.4.
Designers generally make compromises by using both topologies where some of
the critical signals are routed between the ground and power planes.
196 14 Printed Circuit Board (PCB) Layout
where L is the trace length, W is the width, and T is the thickness. All dimensions are
in mm.
14.3 PCB Traces and Vias 197
For DC, the current flows through the entire area of the trace and the DC resistance is
where L is the length of the trace, σ is the conductivity of the metal, W is the width of
the trace, and T is the thickness of the trace.
For AC, the impedance is
where L is the length, σ (sigma) is the conductivity, W is the width, and δ is the skin
depth. Skin depth is small compared to the thickness of the trace and is shown in
Eq. (14.5).
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
δðdeltaÞ ¼ 1=ðπf μσ Þ, ð14:5Þ
where f is the frequency, μ (Mu) is the permeability of free space (4π 10 7 H/m),
and σ is the conductivity (copper ¼ 5.6 107). Or
pffiffiffiffiffiffiffi
δ ¼ 66 106 m 1=f , ð14:6Þ
What is the minimum external Trace Width to guarantee that the design is
IPC-2221 compliant? The trace can be a microstrip or a stripline, depending on a
particular design.
Answer: Using PCB Trace Width Calculator [4], the answer is 15.5 mils mini-
mum width. In this case, the voltage drop is 0.398 V which may be out of the range
required by the receiver. Keep in mind that the IPC 2221 only gives a reliability
limit, it is up to the engineers to determine whether the voltage drop is acceptable.
Design Example 14.2: Characteristic Impedance Calculations
Given:
A 4-layer PCB layout in Fig. 14.7 with two middle ground layers,
External trace thickness ¼ 1.4 mils,
Height (from trace to the adjacent ground) ¼ 6.7 mils,
Trace width ¼ 13 mils, and
Dielectric constant (prepreg material) ¼ 3.66
Calculate the characteristic impedance of the trace.
14.3 PCB Traces and Vias 199
14.3.3 Vias
Vias, Fig. 14.8, have both inductance and capacitance components [3] that can cause
signal and power integrity issues. Vias need to be carefully designed to minimize
parasitics and verified using 3D EM Advanced Solver [6].
4T
Via Inductance ¼ 2T ln þ 1 nH, ð14:7Þ
D0
where T is the thickness of the board and D0 is the via diameter, all in centimeters.
0:55 εr T D1
Via Capacitance ¼ pF, ð14:8Þ
D2 D1
where εr is the dielectric constant of the board, T is the thickness of the board (via
depth), D1 is the via pad diameter, and D2 is the anti-pad diameter (void area). All
dimensions are in centimeters.
The parasitics of vias and traces must be included in the system simulations to get
accurate results, and the method is to use an Advanced 3D Solver [6] tool to create
s-parameter models for the simulations.
14.6 Summary
PCB routing and board stackup are major contributors to design issues related to
EMI, noise, signal integrity, and power integrity, so designers must apply best
practices, including extensive simulations before releasing the design to fabrication,
to improve the probability of having a first-pass success. The guidelines listed in
Sect. 14.5 must be followed to ensure proper routings of single-end or differential
signals.
Also, high-speed design requires having image planes in the PCB stackup to
control the trace characteristic impedance as shown in Design Example 14.2. The
use of an image plane, a ground or power plane located next to the routing layer,
provides low inductance current return paths for high-speed signals. The image
plane helps reduce current loop areas and minimizes the potential differences on
the ground plane. Experiments conducted in [7] compare the EMI for PCBs with and
without an image plane. They demonstrated that a PCB with an image plane shows
around 15 dB reduction in EMI across the frequency spectrum.
References
1. M. Montrose, Printed Circuit Board Design Techniques for EMC Compliance (The Institute of
Electrical and Electronics Engineers, New York, 2000)
2. Texas Instruments Inc., Design Guidelines: Integrated Circuit Design for Reduced EMI. Appli-
cation Note (2000)
3. J. Ardizzoni, A Practical Guide to High-Speed Printed Circuit Board Layout, Analog Dialogue
39-09, Sept 2005
4. PCB Trace Width Calculator, https://www.4pcb.com/trace-width-calculator.html
5. Trace Resistance Calculator, https://www.allaboutcircuits.com/tools/trace-resistance-calculator/
6. Mentor Graphics (VX.2.7) HyperLynx Signal Integrity Simulation Software, http://www.
mentor.com/products/pcb-system-design/circuit-simulation/hyperlynx-signal-integrity/
7. M. Montrose, Analysis on the Effectiveness of Image Planes Within a Printed Circuit Board (The
Institute of Electrical and Electronics Engineers, New York, 1996)
8. Signal Integrity (SI) in High-Speed PCB Designs, Intel Corporation, ID: 683864, 18 Mar 2021
Chapter 15
Electromagnetic Interference (EMI)
Radiated emissions in high-speed systems are caused by fast switching currents and
voltages propagating through printed circuit board traces. As system speed increases,
printed circuit board traces are becoming more effective antennas, and these anten-
nas are radiating unwanted energies that interfere with other circuitry and with other
systems located nearby. This section outlines different ways to design for low EMI
and find the root cause of EMI problems when they occur. It only covers the
electrical design aspects of EMI even though shielding, cabling, and other mechan-
ical fixes can also be used to help reduce the emissions below the maximum
allowable limits. In general, mechanical solutions are expensive for high-volume
designs. Even worse, the mechanical solutions may have to change when the system
speed increases.
To prevent systems from interfering with each other, the FCC sets maximum limits
known as FCC Part 15 A for commercial products and FCC Part 15 B [1] for
consumer devices as shown in Fig. 15.1.
The following lists some of the most common sources of EMI in high-
speed DSPs:
• Fast switching digital signals such as clocks, memory buses, PWM (switching
power supplies)
• Large current return loops
• Not having an adequate power supply decoupling scheme around large DSPs
• Transmission lines
• Printed circuit board layout and stackup, lack of power, and ground planes
• Unintentional circuit oscillations
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 203
T. T. Tran, High-Speed System and Analog Input/Output Design,
https://doi.org/10.1007/978-3-031-04954-5_15
204 15 Electromagnetic Interference (EMI)
55
Field Strength (dBuV/m)
50
Class A
48
45
43
Class B
40
37
35
34 960
216
30
88
25
The five main sources of radiation are digital signals propagating on traces, current
return loop areas, inadequate power supply filtering or decoupling, transmission line
effects, lack of power, and ground planes.
Radiation is classified into two modes, differential mode radiation and common
mode radiation. It is important for engineers to understand the differences between
the two modes in order to develop an effective scheme to mitigate the problem. In
DSP systems, all electrical currents propagate from the source to the load and return
to the original source. This mechanism generates a current loop which creates
differential mode radiation as shown in Fig. 15.2 [2].
Differential mode radiation is directly related to the length of the signal trace, the
driving current, and the operating frequency. The electric field caused by differential
mode radiation is
E ¼ 87:6 1016 f 2 AI , ð15:1Þ
where f is the operating frequency; A is the current loop area created by the trace
length and the board stackup; and I is the driving source current.
Common mode radiation is generated by a differential voltage between two
points on a ground plane. It typically radiates from cables connected to the board
or chassis. In theory, 100% of the source current returns to the source but a small
portion of the current spreads over the entire plane before finding its way back to the
source. This current creates an imbalance in the ground potential and causes com-
mon mode radiation as shown in Fig. 15.3 [2].
15.2 EMI Fundamentals 205
High-Speed Current, I
DSP
Current, I
SDRAM
Cable
System
ground
plane Vgnd
de
eca
/D
dB 20
20 dB
/D
eca
de
fo frequency f
1
f = , where Tr is risetime
π Tr
where f is the frequency, L is the length of the cable in meter, and I is the source
current.
The relationship between the common mode and differential mode radiation for a
given signal is shown in Fig. 15.4 [2]. In general, the differential mode dominates at
a higher frequency spectrum while the common mode radiates more energy around
the operating frequency.
206 15 Electromagnetic Interference (EMI)
1
f 3dB ¼ , ð15:3Þ
πPW
1
f 3dB ¼ : ð15:4Þ
πT r
PW and Tr are the width and the rise time of the signal, respectively. Therefore,
increasing the rise time increases attenuation of the harmonics which leads to lower
Tr Pw
T
1
20dB/decade
π Pw 1
π Tr
Amplitude
40dB/decade
radiation. This method is not always practical because the slower rise time reduces
the timing margin and may violate electrical requirements such as setup and hold
times.
The best technique to minimize EMI generated by digital signals is keeping the
high-speed signal traces as short as possible. It is a good practice for engineers to go
through a design and analyze the traces to see if they are effective antennas or not. A
good rule-of-thumb is keeping the length of the trace less than the wavelength (λ)
divided by 20. Here is the equation
λ c
max trace length ¼ ¼ , ð15:5Þ
20 20f
Current loops are the dominant sources of EMI, so it is important for designers to
understand high-speed and low-speed current return paths and optimize the design to
reduce the loop areas. Figure 15.6 shows two possible current return paths from
points A and B; for high-speed current (>10 MHz), the return is right underneath the
signal and, for low-speed current, the return is the shortest path back to the source.
Should high speed and low speed have hyphen in this figure and title?
As shown in Fig. 15.7 [3], current return creates a loop area that is directly related
to the radiated electric field, so reducing the loop area lowers radiation. Skin effect
modifies the current distribution within a conductor and increases resistance, so the
high-speed current return is right underneath the signal. Skin effect is negligible at
lower frequencies but increases as frequency rises. For a typical conductor used in
A B
Load
Signal Low speed current
return (purple)
U19 A U19 A
14
14
+ 1
3 + 100nF
1
3
2 2
- -
7
1408 1/SO 1408 1/SO
I I
and has no decoupling capacitor, the parasitic inductance is large and requires some
time to charge up. This delay is the root cause of the power supply droop problem.
Power supply droop occurs when the output buffer switches at a fast rate but is
starved for the current needed to drive the load since the parasitic inductance
between the power supply and the DSP becomes an open circuit.
Example 15.1
– A DSP BGA (ball grid array) package has a trace inductance of 1.44 nH.
– This output is driving a 3 in. trace with 1 nS rise time signal.
– This trace is being routed on a typical FR4 printed circuit board. Line character-
istic impedance and IO voltage are 68 Ω and 3.3 V, respectively.
To estimate the power supply droop caused by the parasitic inductance, first let us
estimate the peak current as follows. The dynamic IO current is the current transient
for transmission line load, not steady-state resistive load.
ΔV 3:3 V
I ðpeakÞ ¼ ¼ ¼ 48:5 mA
Zo 68
Since the package inductance is 1.44 nH for 1 nS rise time signal, the internal
voltage droop is
dI 48:5 mA
V ðdroopÞ ¼ L ¼ ð1:44 nHÞ ¼ 70 mV
dt 1 nS
Typically, one DSP power supply pin is shared by many output buffers. This
creates a larger droop and leads to higher radiation. This helps explain why good
power supply decoupling is required for low EMI design.
minimize EMI, the series termination resistors should also be as large as possible
without violating AC timing. A parallel termination resistor as shown in Fig. 15.10 is
commonly used in RF and analog designs but is not practical for digital signals due
to the amount of DC current drained by the 50-Ω resistor. If parallel termination is
required, use a DC blocking capacitor in series with the resistor as demonstrated in
Sect. 15.6.
Table 15.1 [4] shows the source current for different values of the series termi-
nation resistor. Changing the value from 10 to 39 Ω does not have much effect on the
waveform [4], showing about 1 nS degradation, but dramatically reduces the source
current which greatly lowers the radiated emissions. Figure 15.11 shows a DSP
board with a 47-Ω series resistor added to the memory clock, reducing the radiated
emissions 3 dB compared to the emissions of the signal without termination.
Overall, if slower rise time signals are acceptable and do not violate AC timing
specifications, designers should use the largest resistor value to terminate high-speed
signals to optimize the design from an EMI standpoint.
For high-speed DSP systems, it is getting more and more difficult to meet EMI
regulations without using multiple layer PCBs and dedicating some of the layers as
power and ground planes. Compared to a trace, a power or ground plane has a lower
parasitic inductance and provides a shielding effect for high-speed signals. Power
and ground planes also provide natural decoupling capacitance. As described in the
PCB layout section of this document, natural decoupling capacitance occurs when
power and ground planes are spaced closely, yielding higher capacitance. This effect
becomes important at 300 MHz speed or higher. So, adding power and ground
planes simplifies PCB routing and reduces the number of high-frequency decoupling
capacitors required for the DSP.
Another important consideration for the PCB is layer assignment. Refer to the
board layout chapter, Chap. 10, to determine the best board stackup for your
application. Keep in mind that adding a ground plane directly underneath the high-
speed signal plane creates an image plane that provides the shortest current return
paths. Studies in [5, 6] show that image planes greatly reduce radiated emissions.
The comparison between PCB with and without image plane is shown in Figs. 15.12
and 15.13.
212 15 Electromagnetic Interference (EMI)
In summary, here are the guidelines for low EMI system design:
• Add image planes wherever possible.
• Create ground planes if there are spaces available on the routing layers. Connect
these ground areas to the ground plane with vias. Creating a quarter inch via grid
is ideal.
• Add guard traces to high-speed signals if possible.
• Reduce the rise time of the signal if the timing is not critical. This can be
accomplished by including series termination resistors on high-speed buses and
fine-tuning the resistors for optimal signal integrity and EMI. Series termination
resistors lower the source current, increase the signal rise time and reduce
transmission effects. Substantial benefits can be achieved with this approach at
a low cost.
• Keep the current loops as small as possible. Add as many decoupling capacitors
as possible. Always apply current return rules to reduce loop areas.
• Keep high-speed signals away from other signals and especially away from input
and output ports or connectors.
• Avoid isolating the ground plane. If this is required for performance reasons, such
as with audio ADCs and DACs, apply current return rules to connect the grounds
together.
• Avoid connecting the ground splits with a ferrite bead. At high frequencies, a
ferrite bead has a high impedance and creates a large ground potential difference
between the planes.
• Use multiple decoupling capacitors with different values. Every capacitor has a
self-resonant frequency so be careful. Refer to Sect. 13.1 in Power Integrity for
more information.
• For PC board stackup, add as many power and ground planes as possible. Keep
the power and ground planes next to each other to ensure low impedance stackup
or large natural capacitance stackup.
• Add an EMI pi filter on all the signals exiting the box or entering the box.
• If the system fails EMI tests, find the source by tracing the failed frequencies to
their source. For example, assume the design fails at 300 MHz but there is nothing
on the board running at that frequency. The source is likely a third harmonic of a
100 HMz signal.
• Determine if the failed frequencies are common mode or differential mode.
Remove all the cables connected to the box. If the radiation changes, it is common
mode, if not, then it is differential mode. Then, go to the clock source and use
termination or decoupling techniques to reduce the radiation. If it is common
mode, add pi filters to the inputs and outputs. Adding a common choke onto the
cable is an effective but expensive method of reducing EMI.
214 15 Electromagnetic Interference (EMI)
References
A B
AC-Coupled, 19–25, 40, 44 Backward crosstalk, 105, 107
AC timing, 210 Ball grid array, 169, 173, 209
Active Filter Design Bandwidth, 7, 27–30, 47, 49, 50, 57, 59, 62, 63,
dynamic range limitation, 28 73, 137, 138, 183, 184, 186, 188, 189
Adjacent power and ground, 194, 195 BGA package, 182, 193
Algorithms, 59 Biasing, 20, 27, 34–38, 57
Amplifier circuit, 5, 37, 38, 47, 75, 150 Bit error rate (BER), 7, 127
Amplitude modulating, 62 Bluetooth, 1, 7, 8
Analog circuit simulator, 13, 186, 190 Board capacitance, 194, 195
Analog Filter Design Boost converter, 149
Butterworth, 27, 31 Broadband, 7
Cauer and Besser, 27 BT.656, 6
Chebyshev, 27 BT.1120, 6
filter topology, 27, 57 Buck converter, 148–150, 152
highpass filter, 43 Bulk capacitor, 118, 139, 168, 170, 173,
Inverse Chebyshev, 27 176–179, 182
inverting lowpass, 46, 47
non-inverting lowpass, 46
Sallen-Key, 52–56 C
Analog-to-digital Capacitance, 3, 67, 82, 88, 93, 154, 157,
Analog-to-Digital Converter (ADC), 5, 29, 162–165, 167, 169, 170, 172–174, 176,
30, 57, 60, 73–76 177, 179, 194, 195, 199, 211, 213
anti-aliasing filter, 5, 29, 59, 62, 63 Capacitive load, 67, 88, 162
Differential Non-Linearity (DNL), 73, Capacitor characteristics, 163–165
75, 78 Central Processing Unit (CPU), 1, 3, 4, 59, 60,
Integral Non-Linearly (INL), 76 109–111, 147, 152–157, 159, 176, 177,
Antenna, 108, 109, 111 179, 193
Anti-Pad, 128, 200 Ceramic, 164, 168, 176, 177, 179, 181
Artifacts, 5, 6, 154 Characteristic impedance, 12, 81–83, 88, 91,
Attenuation, 40, 41, 44, 45, 57, 69, 73, 139, 93, 172, 198, 201, 202, 209
140, 186, 187, 190, 191, 206 Charge Pump, 132
Audio Clock oscillator, 133
filter, 46, 50 Clock speed, 116
© The Editor(s) (if applicable) and The Author(s), under exclusive license to 215
Springer Nature Switzerland AG 2023
T. T. Tran, High-Speed System and Analog Input/Output Design,
https://doi.org/10.1007/978-3-031-04954-5
216 Index
M
I Magnitude, 42, 43
IEEE802.11, 8 Maximum crosstalk, 107
Image planes, 196, 200, 202, 211, 213 Maximum dynamic range, 31
Imbalance, 204 Maximum swing, 34
Impedance, 5, 19, 20, 28, 29, 31, 32, 37, 40–42, Memory signal integrity, 118–120
44, 54, 82, 83, 93, 94, 101, 112, 125, Memory sub-system
128, 139, 161, 163, 164, 166, 167, 180, read cycle, 115, 118, 120, 123
181, 184, 186, 188–190, 192–195, SDRAM, 116
197–199, 213 write cycle, 116, 117
Impedance mismatch, 87, 208, 209 Methodologies, 11, 13, 17, 46
Inductances, 3, 6, 82, 93, 104, 153, 162, 164, Microphone, 5
185, 188, 196, 199, 202, 209 Microstrip, 12, 93, 98, 99, 195, 196, 198, 199
Inductor characteristics, 165–167 Minimizing radiation, 108
Initial voltage, 81, 83 MPEG-2 decode, 109
Input impedance, 24, 31–33, 40, 41, 43, 44, MPEG-2 encoder, 110
67, 88 Multi-Channel Buffered Serial Port (McBSP), 4
Insertion loss, 97, 98, 101, 128
Insulating core, 166
Interference reduction, 112 N
IPC 2221, 197, 198 NEXT, 97
Isolation, 5–7, 28, 133, 138–142, 148, 183–191 NMOS, 25
Noise characteristics
avalanche noise, 4
J burst noise, 4
Jitter flicker noise, 3, 4
cycle-to-cycle jitter, 133, 135 shot noise, 3
long term jitter, 133, 134 thermal noise, 3
period jitter, 133, 135, 137 Non-adjacent topology, 195
Non-embedded designs, 119
NTSC spectrum, 111
K
Kirchhoff’s Current Law, 32, 33
O
Open ended load, 82
L Operating frequency, 9, 24, 47, 81, 95, 133,
Lattice diagram, 83 167, 181, 204, 206
LC filter, 149, 154 Operational amplifier
LDO versus switching regulators, 154 AC couple, 39–46
Least significant bit, 60, 61 AC gain, 32
Linear and non-linear distortions, 59 AC ground, 36, 44
218 Index
Q
P Quantization, 59, 63–67, 77, 136
Parallel capacitance, 194 error, 60, 64, 67
Parallel combination, 31, 35, 37, 50, 55 step, 64
Parallel termination, 84–88, 91, 94, 210
Parasitic inductance, 155, 162, 164, 168, 170,
173, 182, 209, 211 R
Passband, 27, 30, 56 Radiated emissions, 183, 200, 203, 210, 211
Passive and active filters, 27–28 RapidIO, 193
Passive filter Reconstruction filters, 59, 71, 78
passive lowpass filter, 28–29 Redriver, 99–101, 125, 127
Pi filter, 136, 138–142, 169, 183–188, Reflected voltage, 84
190–192, 213 Reflection coefficients, 82, 96
T filter, 138–140, 183, 188–191 Reflections, 1, 3, 71, 81–83, 87, 93, 96,
PCB manufacture, 193 128, 208
PCI Express, 7, 8 Reliability, 28, 152–154, 194, 197, 198
Peripherals Component Interface (PCI), 6 Reset, 154, 156, 158
Phase-frequency detector (PFD), 132, 137 Resolution, 73–74, 110
Phase-locked loop (PLL) Retimer, 99–101
APLL, viii, 131–132, 136–138, 174 Return losses, 97, 98, 101
DPLL, 131, 136–138, 142, 174 RF filter, 13
Divide-by-M counter, 131, 132, 137 RF spectrum, 8
PLL isolation, 133, 138 Ripple rejection, 141, 147
Phase relationship, 47, 50 Rise and fall-times, 23, 84, 95, 115
Phase-shift, 116 RMS, 65, 66
Pixel clocks, 6 Routability, 196
Placing Decoupling Capacitors, 182–183 Routing layer, 108, 194–196, 200, 202, 213
Power consumption, 1, 67, 97, 116, 136,
142, 161
Power dissipation, 91, 140, 146, 148, 149, 154 S
Power efficiency and dissipation, 145 Sampling
Power planes, 182, 183, 193–195, 200, 202 aliasing, 57
Power sequencing, 153, 154, 156–159 frequency, 27, 62, 73, 74
Power sequencing verification, 158 noise, 57, 59, 71
Power spectral density, 112 Nyquist, 62, 77
Power supply architectures, 145–152, 155, 156, Self-resonant frequency, 163–165, 170, 173,
159 176, 177, 179, 213
Power supply decoupling, 3, 20, 153, 161–183, SERDES, 13, 71, 95, 96, 98, 125
191, 203, 208, 209 Series termination, 87, 88, 91, 94, 111, 112,
Power supply droop, 115, 161, 163, 169, 133, 209, 210, 213
196, 209 Serpentines, 129, 201
Index 219
Shielding, 7, 57, 152, 203, 211 Transient response, 146, 151, 153
Shielding layer, 200 Transistor, 3, 25, 145, 146, 149, 150
Signal integrity, 6, 8, 16, 17, 88, 98, 102, 115, Transmission coefficients, 96
118, 162, 193, 194, 197, 202, 209, 213 Transmission line (TL)
Signal quality, 59, 101, 183, 196 lossy TLs, 81, 94
Signal-to-noise (SNR), 3, 7, 30, 66, 67, 71, Transmission line effects, 4, 69, 101, 112, 118,
73–74, 77, 78 155, 193, 204
Simulations, 1, 8, 11–17, 20, 27, 30, 41, 45, 47, Two-port network, 96
50, 55, 85–93, 98, 101, 102, 105, 107,
120, 125, 127, 129, 130, 138–140, 186,
187, 190, 191, 200–202, 209 U
Single rail supply, 34, 37 UI, 127, 128
Sinusoidal, 65, 66 Undershoots, 6, 81, 92, 95, 118, 209
SiWave, 17 Unit Interval, 127
Skin depth, 197 Universal serial bus (USB), 3, 8, 71, 78,
S-parameter, 13, 14, 95–101, 125, 200 125–130, 136, 193
Spice, 13–16, 20 Unstable, 47, 51, 138, 154, 157
Spread spectrum, 7, 112, 113
Squarewave, 19, 137, 206
Stability analysis, 47 V
Stackup, 98, 99, 128, 193–195, 200, 202–204, Vias, 8, 11–13, 101, 102, 128, 196–201, 213
211, 213 Via structure, 199
Standard definition, 6 Video, 1, 3–8, 22, 27, 31, 35, 59, 70, 73,
Stripline, 12, 195, 196, 198 109–112, 136, 145, 154, 159, 167, 173,
Switching current estimation, 175 186, 189
Symmetrical, 22, 23, 34, 57, 74, 183 Voltage controlled oscillator (VCO), 132, 136
System start-up, 158 Voltage droop, 145, 153, 154, 162, 209
VREF, 61, 66, 67, 71, 116, 117, 120, 148
T
Tantalum capacitors, 164, 168, 176, 178, 180 W
Target impedance, 8, 180–181, 192 Waveforms at the terminated and unterminated
Termination, 6, 81, 85, 88–89, 91–92, 96, 116, loads, 85
119, 120, 210, 213 Wavelength, 108, 112, 207
Thermal noise, 3 Wireless local area network (WLAN), 1, 7
Threshold voltage, 118
Time shifting, 109–111
Timing Interval Analyzer (TIA), 135 Z
Timing violations, 71, 85, 95, 112, 120, 131 Zero-to-peak voltage, 65
TINA, 11, 13, 16, 20, 25 Zig zag pattern, 93
Transceivers, 22