DATA receivers for digital transmission and storage systems

Similar documents
Parameter Derivation of Type-2 Discrete-Time Phase-Locked Loops Containing Feedback Delays

IN THIS PAPER, we consider a class of continuous-time recurrent

THE problem of phase noise and its influence on oscillators

Fundamentals of PLLs (III)

ELECTRODYNAMIC magnetic suspension systems (EDS

A conjecture on sustained oscillations for a closed-loop heat equation

R a) Compare open loop and closed loop control systems. b) Clearly bring out, from basics, Force-current and Force-Voltage analogies.

ONE can design optical filters using different filter architectures.

Article (peer-reviewed)

798 IEEE TRANSACTIONS ON CIRCUITS AND SYSTEMS II: ANALOG AND DIGITAL SIGNAL PROCESSING, VOL. 44, NO. 10, OCTOBER 1997

AN EXTENSION OF GENERALIZED BILINEAR TRANSFORMATION FOR DIGITAL REDESIGN. Received October 2010; revised March 2011

Lyapunov Stability of Linear Predictor Feedback for Distributed Input Delays

PERIODIC signals are commonly experienced in industrial

Sensitivity Analysis of Coupled Resonator Filters

BECAUSE this paper is a continuation of [9], we assume

An Alternative Proof for the Capacity Region of the Degraded Gaussian MIMO Broadcast Channel

OVER THE past 20 years, the control of mobile robots has

Motivation for CDR: Deserializer (1)

Switched-Capacitor Circuits David Johns and Ken Martin University of Toronto

Chapter 2 Application to DC Circuits

Digital Phase-Locked Loop and its Realization

Fast Seek Control for Flexible Disk Drive Systems

H State-Feedback Controller Design for Discrete-Time Fuzzy Systems Using Fuzzy Weighting-Dependent Lyapunov Functions

An Iteration-Domain Filter for Controlling Transient Growth in Iterative Learning Control

Robust Gain Scheduling Synchronization Method for Quadratic Chaotic Systems With Channel Time Delay Yu Liang and Horacio J.

MANY algorithms have been proposed so far for head-positioning

LECTURE 3 CMOS PHASE LOCKED LOOPS

Signal Processing for Digital Data Storage (11)

Optimal Placement of Training Symbols for Frequency Acquisition: A Cramér-Rao Bound Approach

/$ IEEE

IN SIGNAL processing theory, a linear, time-invariant (LTI),

Adaptive Inverse Control based on Linear and Nonlinear Adaptive Filtering

SYNTHESIS OF BIRECIPROCAL WAVE DIGITAL FILTERS WITH EQUIRIPPLE AMPLITUDE AND PHASE

DUE to its practical importance in communications, the

Research Article On Complementary Root Locus of Biproper Transfer Functions

Issues with sampling time and jitter in Annex 93A. Adam Healey IEEE P802.3bj Task Force May 2013

Jerk derivative feedforward control for motion systems

IN the multiagent systems literature, the consensus problem,

KINGS COLLEGE OF ENGINEERING DEPARTMENT OF ELECTRONICS AND COMMUNICATION ENGINEERING

POSITIVE REALNESS OF A TRANSFER FUNCTION NEITHER IMPLIES NOR IS IMPLIED BY THE EXTERNAL POSITIVITY OF THEIR ASSOCIATE REALIZATIONS

Notes on DPLL. Haiyun Tang Most of material is derived from Chistian s report[4]

Maximally Flat Lowpass Digital Differentiators

Performance of an Adaptive Algorithm for Sinusoidal Disturbance Rejection in High Noise

Analysis of Communication Systems Using Iterative Methods Based on Banach s Contraction Principle

HOPFIELD neural networks (HNNs) are a class of nonlinear

Estimation of the Capacity of Multipath Infrared Channels

Chapter 1. Introduction

458 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 16, NO. 3, MAY 2008

THE clock driving a digital-to-analog converter (DAC)

MANY adaptive control methods rely on parameter estimation

C(s) R(s) 1 C(s) C(s) C(s) = s - T. Ts + 1 = 1 s - 1. s + (1 T) Taking the inverse Laplace transform of Equation (5 2), we obtain

388 Facta Universitatis ser.: Elec. and Energ. vol. 14, No. 3, Dec A 0. The input-referred op. amp. offset voltage V os introduces an output off

Closed-Form Design of Maximally Flat IIR Half-Band Filters

A FEEDBACK STRUCTURE WITH HIGHER ORDER DERIVATIVES IN REGULATOR. Ryszard Gessing

A Log-Frequency Approach to the Identification of the Wiener-Hammerstein Model

Computing running DCTs and DSTs based on their second-order shift properties

Roundoff Noise in Digital Feedback Control Systems

Control Systems. University Questions

THIS paper is aimed at designing efficient decoding algorithms

STABILITY ANALYSIS OF HVDC CONTROL LOOPS

Determining Appropriate Precisions for Signals in Fixed-Point IIR Filters

PAPER A Low-Complexity Step-by-Step Decoding Algorithm for Binary BCH Codes

Filter Design for Linear Time Delay Systems

The output voltage is given by,

Nonlinear Discrete-Time Observer Design with Linearizable Error Dynamics

CLASSICAL error control codes have been designed

Professional Portfolio Selection Techniques: From Markowitz to Innovative Engineering

Adaptive Inverse Control

. pl Ifler CH29645/91/ $ IEEE 66 CONTROL THEORY APPLIED TO THE DESIGN OF AGC CIRCUFFS

Shifted-modified Chebyshev filters

Quantitative Feedback Theory based Controller Design of an Unstable System

Analysis and Synthesis of State-Feedback Controllers With Timing Jitter

QFT Framework for Robust Tuning of Power System Stabilizers

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 58, NO. 5, MAY invertible, that is (1) In this way, on, and on, system (3) becomes

High-Speed Low-Power Viterbi Decoder Design for TCM Decoders

Automatic Stabilization of an Unmodeled Dynamical System Final Report

Optimization of Phase-Locked Loops With Guaranteed Stability. C. M. Kwan, H. Xu, C. Lin, and L. Haynes

Optimal Polynomial Control for Discrete-Time Systems

R10 JNTUWORLD B 1 M 1 K 2 M 2. f(t) Figure 1

An Adaptive LQG Combined With the MRAS Based LFFC for Motion Control Systems

Stability of Switched Linear Hyperbolic Systems by Lyapunov Techniques

Behavior of Phase-Locked Loops

Here represents the impulse (or delta) function. is an diagonal matrix of intensities, and is an diagonal matrix of intensities.

An Invariance Property of the Generalized Likelihood Ratio Test

INSTRUMENTAL ENGINEERING

IN this paper, we consider the capacity of sticky channels, a

The Generalized Nyquist Criterion and Robustness Margins with Applications

Implementation Issues for the Virtual Spring

Predictive Cascade Control of DC Motor

Fundamentals: Frequency & Time Generation

Adaptive MMSE Equalizer with Optimum Tap-length and Decision Delay

ME 132, Dynamic Systems and Feedback. Class Notes. Spring Instructor: Prof. A Packard

IMPROVEMENTS IN ACTIVE NOISE CONTROL OF HELICOPTER NOISE IN A MOCK CABIN ABSTRACT

A Novel Integral-Based Event Triggering Control for Linear Time-Invariant Systems

A Strict Stability Limit for Adaptive Gradient Type Algorithms

On the Frequency-Domain Properties of Savitzky-Golay Filters

Time-Varying, Frequency-Domain Modeling and Analysis of Phase-Locked Loops with Sampling Phase-Frequency Detectors

AdaptiveFilters. GJRE-F Classification : FOR Code:

EC CONTROL SYSTEM UNIT I- CONTROL SYSTEM MODELING

LED lamp driving technology using variable series-parallel charge pump

Transcription:

IEEE TRANSACTIONS ON CIRCUITS AND SYSTEMS II: EXPRESS BRIEFS, VOL. 52, NO. 10, OCTOBER 2005 621 Effect of Loop Delay on Phase Margin of First-Order Second-Order Control Loops Jan W. M. Bergmans, Senior Member, IEEE Abstract This paper analyzes the phase margin of first-order second-order control loops in the presence of a loop delay establishes rules of thumb on the maximum permissible delay for a given phase margin. Both discrete-time continuous-time loops are considered. Results are applicable, for example, to adaptive filters to first-order second-order phase-locked loops. Index Terms Adaptive filter, control loop, loop delay, phase margin, phase-locked loop (PLL). I. INTRODUCTION DATA receivers for digital transmission storage systems normally contain various control loops (e.g., automatic gain control, adaptive dc compensation, adaptive equalization, timing recovery) that jointly act to permit reliable data recovery in spite of varying or uncertain system conditions [1], [2]. Data rates computational requirements in these systems tend to outpace Moore s law [3], [4]. This spurs increasing use of techniques such as pipelining parallelization. As a result, loop delays increase can become a limiting factor, especially during acquisition. Control loops in data receivers are often decision-directed, i.e., control information is derived with the aid of the bit decisions. Bit detectors necessarily become increasingly powerful,, as a consequence, their detection delay increases, thus further increasing loop delay. Especially during acquisition, the compound loop delay can become prohibitively large. A typical approach to mitigate this problem involves the use of auxiliary detectors that produce tentative decisions with minimum delay [5], [6]. Against the above background, it is increasingly important to underst how loop delay affects the properties of the loop. Previous studies have been directed at identifying the edge of the stability region of first-order second-order discrete-time loops in the presence of a loop delay [7], [8]. In practice, it is desirable to operate loops well within the stability region. In this respect, the so-called phase margin is often used as a measure of the degree of loop stability, it is desirable to be able to dimension the loop for a prescribed phase margin. This paper analyzes the impact of loop delay on the phase margin of first second-order control loops of both the discrete-time the continuous-time variety, with the remainder of the paper organized as follows. Section II analyzes the first-order loop, which Manuscript received October 9, 2003; revised April 3, 2004. This work was supported in part by the European Union under the TwoDOS IST Project IST- 2001-34168. This paper was recommended by Associate Editor X. Yu. The author is with Eindhoven University of Technology, 5600 MB Eindhoven, The Netherls (e-mail: J.W.M.Bergmans@tue.nl). Digital Object Identifier 10.1109/TCSII.2005.852003 Fig. 1. First-order discrete-time control loop with delay M total open-loop gain K. is representative, for example, of automatic gain control (AGC) loops, dc control loops, adaptive equalizers. Section III focuses on the second-order discrete-time high-gain loop. Since second-order control loops are mainly found in phase-locked loops (PLLs), this section is cast in PLL terms. This also applies to Section IV, which focuses on the second-order continuous-time high-gain loop. Behavior of this loop can approximate that of its discrete-time counterpart, yet analytical results are comparatively simple hence insightful. For both loops, an adequate phase margin is required to limit jitter. Section V establishes rules of thumb for accomplishing a prescribed phase margin draws conclusions. Detailed analysis for the secondorder discrete-time continuous-time loops is relegated to Appendices I II. II. FIRST-ORDER CONTROL LOOP We first consider the discrete-time first-order loop (Fig. 1). The ideal value of the control parameter is denoted, the actual value is denoted, the error is denoted (the subscript denotes the time index expressed in sampling intervals ). The error is delayed by sampling intervals scaled by a compound loop gain which determines loop bwidth tracking speed. The loop is closed via a first-order ideal integrator (the symbol denotes a delay of one sampling interval ). The model of Fig. 1 is illustrative for, e.g., AGC dc compensation loops. A typical gain for such loops is during acquisition. After acquisition, a substantially lower value (e.g., ) is normally used for parameter tracking. In the absence of a delay, the loop has a first-order exponential impulse response with a time constant (expressed in seconds). We can think of as the normalized time constant, i.e., the time constant expressed in symbol intervals. Clearly,. The loop has closed-loop transfer function 1057-7130/$20.00 2005 IEEE

622 IEEE TRANSACTIONS ON CIRCUITS AND SYSTEMS II: EXPRESS BRIEFS, VOL. 52, NO. 10, OCTOBER 2005 is the open-loop transfer function is the transfer function of the discrete-time integrator in Fig. 1 (we use capitals to denote the transforms of the corresponding lower case sequences). In frequency-domain notation, we may write with Here, is a normalized measure of frequency, with corresponding to the sampling rate. The phase margin PM is determined at the unity-gain frequency of, i.e., at the frequency for which. The system will be unstable if the phase of exceeds rad at. Bydefinition [9, Sec. 6.4], PM is the margin that is left with respect to this stability limit, i.e., In practical systems one typically requires a phase margin of around 45 to 60, i.e., around to radians. For the loop at h we evidently have. Correspondingly For loop gains of practical interest (i.e., for, whence ) we have Fig. 2. Discrete-time phase-domain model of second-order high-gain PLL with a loop delay of M symbol intervals. discrete loop delay of sampling intervals is replaced by a continuous delay. Upon retracing the above steps for this continuous-time model, one readily verifies that all approximate equality signs above become exact equalities. This reflects the fact that the continuous-time discrete-time loops behave essentially identically for [2, Ch. 11]. Hence, the above results carry over directly to the continuous-time case. III. SECOND-ORDER DISCRETE-TIME PHASE-LOCKED LOOP This loop is of the high-gain second-order type has the discrete-time model of Fig. 2. The input phase is tracked by a voltage-controlled oscillator (VCO) that is modeled as an ideal integrator with output phase. The phase error (normalized in sampling intervals ) serves as the VCO input after being delayed across sampling intervals filtered by the loop filter. This filter has a proportional an integrating path with total open-loop gains, respectively. Any gain of the phase detector /or VCO is accommodated in. These two parameters together determine the normalized natural frequency damping factor of the PLL according to [2, Ch. 11] Evidently PM decreases as increases. The contribution 0.5 can be regarded as the effective delay introduced by the integrator in Fig. 1, we can think of as the compound effective loop delay. The edge of the stability region is demarcated by. Here. Phase margins of (45 ) (60 ) are obtained for, respectively. In terms of the normalized time constant,wehave This equation reveals the largest loop delay that is permissible to achieve a prescribed phase margin. In particular for a phase margin of 45, should be no larger than times the normalized time constant, versus times for a phase margin of 60. Continuous-Time Loop: The model of this loop is that of Fig. 1, but with all discrete-time quantities replaced by their continuoustime counterparts. Specifically, the discrete-time integrator is replaced by a continuous-time integrator (with transfer function denotes the angular frequency), the (1) A typical loop uses during acquisition during tracking. In both cases, the damping factor is in the order of unity. It is worth mentioning that the notions of natural frequency damping factor pertain to a second-order system are, hence, strictly speaking, only applicable in the absence of a delay. For the sake of consistency, we will nevertheless use them here, defined as in (1), for any value of. The phase margin of the loop is derived in Appendix I. For various values of, Fig. 3 portays the combinations of that yield a phase margin of 60. Similar graphs are shown in Fig. 4 for phase margins of 30 0. The latter margin corresponds to the edge of the stability region, which was derived earlier in [7]. Irrespective of, for phase margins of practical interest, the curves all have a maximum for damping factors in the order of unity. They level off only slowly as increases beyond this optimum, yet decline rapidly as decreases. This suggests that a proper engineering choice for would be a value somewhat above unity.

BERGMANS: EFFECT OF LOOP DELAY ON PHASE MARGIN OF FIRST-ORDER AND SECOND-ORDER CONTROL LOOPS 623 Fig. 3. Normalized natural frequency! T versus damping factor for second-order discrete-time PLL with a phase margin of 60. The loop delay M is used as a parameter. Fig. 5. Normalized natural frequency! T versus damping factor for second-order continuous-time PLL with phase margin of 60. The loop delay amounts to (M +0:5)T. Fig. 4. As Fig. 3, but for phase margins of 30 0. IV. SECOND-ORDER CONTINUOUS-TIME PHASE-LOCKED LOOP The phase margin of the continuous-time PLL is derived in Appendix II has a considerably simpler analytical form than that of the discrete-time PLL. Specifically, is a monotonically increasing function of the damping factor is the effective loop delay in seconds. In terms of the discrete-time loop of (2) Fig. 6. Same as Fig. 5, but for phase margins of 30 0. Fig. 2, we may equate with the contribution accounts for the effective delay of the discrete-time integrator that models the VCO. In the absence of a loop delay (i.e., for ), PM is fully determined by does not depend on. The presence of a loop delay causes PM to decrease in linear proportion to the normalized loop delay. Figs. 5 6 are the counterparts of Figs. 3 4. Only for small loop delays in conjunction with a small phase margin is there a significant difference between the characteristics of both

624 IEEE TRANSACTIONS ON CIRCUITS AND SYSTEMS II: EXPRESS BRIEFS, VOL. 52, NO. 10, OCTOBER 2005 Fig. 7. Minimum damping factor that supports a prescribed phase margin PM. PLLs. For phase margins of practical interest, we can use the simple analytical results of Appendix II as a close approximation for those of the discrete-time PLL. Since PM depends on the product of, the curves of Fig. 5 ( similarly for Fig. 6) are identical except for a -axis scaling factor. Accordingly, for a given phase margin, their global maximum occurs for the same damping factor, irrespective of. Similarly, the minimum damping factor that is needed to achieve a prescribed phase margin PM is independent of is the solution of the equation. An equivalent (though slightly more complex) equation was derived earlier for loops without delay in [9, eq. 6.31]. Fig. 7 depicts for phase margins between 0 90. For small phase margins (say below 45 ), increases linearly with PM. At higher phase margins, it increases ever more rapidly, very large damping factors are required for phase margins close to 90. Fig. 8, computed from (2), depicts the maximum normalized loop delay that is permissible for a given phase margin PM. For a practical damping factor in the order of unity or somewhat higher, the normalized loop delay should apparently be of the order of 0.1 to 0.2 to achieve practical phase margins on the order of 45 to 60. The points at which the curves of Fig. 8 cross the axis define the highest phase margins that are achievable with a given damping factor are consistent with the minimum damping factor that is needed for a prescribed phase margin as in Fig. 7. Fig. 8. Maximum normalized loop delay (M +0:5)! T as a function of the phase margin PM. The damping factor is used as a parameter. Equivalently, loop dynamics should be dimensioned for to be at least twice as large as. 2) In high-gain second-order loops, the damping factor is preferably selected somewhat larger than unity, irrespective of. Here, loop delay should be at least 5 to 10 times smaller than the inverse of the normalized natural frequency in order for the loop to have an adequate phase margin. Equivalently, should be chosen to be at least 5 to 10 times smaller than. These rules are likely to be helpful in the design of the concerned loops. APPENDIX I PHASE MARGIN OF HIGH-GAIN SECOND-ORDER DISCRETE-TIME PLL The loop of Fig. 2 has transfer function with notation, we may write. In frequency-domain V. FINAL REMARKS For loop conditions of practical interest, we have found that the discrete-time continuous-time loops behave essentially identically in the presence of a loop delay. This is true both for first-order second-order loops. The above results permit us to formulate the following simple rules of thumb. 1) In first-order loops, loop delay should be less than half of the normalized loop time constant in order for the loop to have an adequate phase margin.

BERGMANS: EFFECT OF LOOP DELAY ON PHASE MARGIN OF FIRST-ORDER AND SECOND-ORDER CONTROL LOOPS 625 Clearly, if only if therefore, i.e., if is. The corresponding open-loop transfer function or, equivalently, if be denoted. The solution of this equation is. This may alternatively (3) We first identify the angular frequency at which. Clearly, so that if only if (6) We recall that. Condition (6) may be recast in terms of as It can be observed that only the ratio of comes into play. The absolute value of does not matter. Equation (7) has only one positive root, which is given by (7) For practical values of, will always be positive,, since the desired value of is also positive, the solution will be, i.e., the root with the sign in (3). It should be noted that is four times the square of a sine, for this reason can fundamentally not become larger than 4. In cases exceeds 4, there exists no frequency for which for the given values of,,. The phase margin is determined by the phase of at the frequency that was just identified. The phase of depends on according to Correspondingly,,. Having identified, we next turn our attention to the phase characteristics of. Clearly (8) The phase margin may be expressed in terms of as It follows that APPENDIX II PHASE MARGIN OF HIGH-GAIN SECOND-ORDER CONTINUOUS-TIME PLL The model of this PLL is the one of Fig. 2 but with all discretetime operations replaced by their continuous-time counterparts. Specifically, discrete-time integrators are replaced by continuous-time integrators (with transfer function is the angular frequency), the discrete loop delay of sampling intervals is replaced by a continuous delay (4) (5) REFERENCES [1] H. Meyr, M. Moeneclaey, S. A. Fechtel, Digital Communication Receivers, 2nd ed. New York: Wiley, 1997. [2] J. W. M. Bergmans, Digital Baseb Transmission Recording. Boston, MA: Kluwer, 1996. [3] J. Rabaey, Low-power silicon architectures for wireless communications, in Proc. ASP-DAC 2000, Yokohama, Japan, Jan. 2000, pp. 377 380. [4] Y. Miura, Information storage for the broad-b network era Fujitsu s challenge in hard disk drive technology, Fujitsu Sci. Tech. J., vol. 37, no. 2, pp. 111 125, Dec. 2001. [5] R. Cideciyan, F. Dolivo, R. Hermann, W. Hirt, W. Schott, A PRML system for digital magnetic recording, IEEE J. Sel. Areas Commun., vol. 10, no. 1, pp. 38 56, Jan. 1992. [6] R. Alini et al., A 200 MSample/s trellis-coded PRML channel with analog adaptive equalizer digital servo, IEEE J. Solid-State Circuits, vol. 32, no. 11, pp. 1824 1838, Nov. 1997. [7] J. W. M. Bergmans, Effect of loop delay on stability of discrete-time PLL, IEEE Trans. Circuits Syst. I, Fundam. Theory Appl., vol. 42, no. 4, pp. 229 231, Apr. 1995. [8] A. De Gloria, D. Grosso, M. Olivieri, G. Restani, A novel stability analysis of a PLL for timing recovery in hard disk drive, IEEE Trans. Circuits Syst. I, Fundam. Theory Appl., vol. 46, no. 8, pp. 1026 1031, Aug. 1999. [9] G. F. Franklin, J. D. Powell, A. Emami-Naeini, Feedback Control of Dynamic Systems, 4th ed. Upper Saddle River, NJ: Prentice-Hall, 2002.