Fact-checked by Grok 2 weeks ago

Minimum phase

In and , a minimum-phase system is defined as a linear time-invariant (LTI) system that is causal and , and whose is also causal and . This property distinguishes minimum-phase systems from other LTI systems with the same , as they exhibit the smallest possible (or group delay) among all causal systems sharing that response. For discrete-time systems, the minimum-phase condition is equivalent to all poles and zeros of the system's H(z) lying strictly inside the unit circle in the z-plane, ensuring both the system and its $1/H(z) are causal and stable. In continuous-time systems, the analogous condition requires all poles and zeros to reside in the open left half of the s-plane. Minimum-phase systems also feature impulse responses that are "energy-concentrated" toward the beginning, meaning a larger fraction of the total energy accumulates earlier in time compared to non-minimum-phase counterparts with identical magnitude responses. The significance of minimum-phase systems lies in their practical advantages, particularly the existence of a stable inverse, which enables applications such as signal equalization, , and inverse filtering to recover distorted inputs without . In control systems, they offer improved margins and faster transient responses, making them preferable for high-gain designs over non-minimum-phase systems that introduce additional . Additionally, spectral factorization techniques can uniquely recover a minimum-phase from its magnitude-squared response alone, a property exploited in and .

Fundamentals

Definition via Inverse System

A minimum-phase is characterized by the property that both the and its are causal and . This definition applies to linear time-invariant (LTI) systems in both continuous and domains, ensuring that the can recover the input signal without introducing or non-causality. In contrast, non-minimum-phase systems possess zeros outside the region of (right half-plane in continuous time or outside the unit circle in time), rendering their inverses either unstable or requiring non-causal implementations. In discrete-time systems, the provides the framework for deriving this condition. Consider a rational H(z) = \frac{B(z)}{A(z)}, where B(z) and A(z) are polynomials with roots representing , respectively. For H(z) to be causal and stable, all poles (roots of A(z)) must lie inside the unit circle in the z-plane, i.e., |p_k| < 1 for each pole p_k. The system is given by G(z) = 1/H(z) = \frac{A(z)}{B(z)}, so the poles of G(z) are the zeros of H(z) (roots of B(z)). For G(z) to be causal and stable, these zeros must also satisfy |z_m| < 1 for each zero z_m, ensuring the region of convergence includes the exterior of the unit circle while maintaining stability. Thus, a minimum-phase H(z) requires all zeros inside the unit circle, guaranteeing a causal and stable . For continuous-time systems, the Laplace transform yields an analogous derivation. The transfer function is H(s) = \frac{B(s)}{A(s)}, with causality and stability requiring all poles in the open left half-plane (LHP), i.e., \operatorname{Re}(p_k) < 0. The inverse G(s) = 1/H(s) = \frac{A(s)}{B(s)} has poles at the zeros of H(s). Stability and causality of G(s) demand these zeros also lie in the open LHP, \operatorname{Re}(z_m) < 0, so the region of convergence is the right half-plane. Therefore, minimum-phase systems in continuous time have all zeros in the LHP, ensuring the inverse filter $1/H(s) is both causal and stable. The concept of minimum phase originated in the context of filter design and control theory during the mid-20th century, with H. W. Bode formalizing it in his seminal work on feedback amplifiers. Bode introduced the term to describe systems amenable to stable equalization in network analysis, emphasizing the implications for phase minimization in stable designs.

Discrete-Time Example

A simple example of a minimum-phase system in discrete time is the first-order finite impulse response (FIR) filter with transfer function H(z) = 1 + 0.5 z^{-1}. This system has no poles, as it is FIR, and a single zero found by solving $1 + 0.5 z^{-1} = 0, which yields z = -0.5. Since the magnitude of this zero is |z| = 0.5 < 1, it lies inside the unit circle, confirming that all zeros are inside the unit circle and thus the system is minimum phase. As per the definition via inverse system, the reciprocal G(z) = 1 / H(z) must be stable and causal. Here, G(z) = 1 / (1 + 0.5 z^{-1}) = z / (z + 0.5), which has a pole at z = -0.5 inside the unit circle, ensuring stability for a causal realization. To illustrate the distinction, consider a non-minimum-phase counterpart obtained by flipping the zero outside the unit circle: H'(z) = 1 + 2 z^{-1}. The zero is at z = -2, with |z| = 2 > 1. The inverse G'(z) = 1 / H'(z) = z / (z + 2) has a pole at z = -2 outside the unit circle, rendering it unstable for causal implementation. The impulse response of the minimum-phase system h = \delta + 0.5 \delta[n-1] concentrates energy early (at n=0 and n=1), with values [1, 0.5]. In contrast, the non-minimum-phase version has h' = \delta + 2 \delta[n-1], with values [1, 2], shifting more energy later despite the finite duration. This front-loading in minimum-phase systems arises from the inward zero placement. Numerical evaluation of the frequency response for H(z) shows a magnitude |H(e^{j\omega})| = |1 + 0.5 e^{-j\omega}|, which peaks at 1.5 for \omega = 0 and dips to 0.5 at \omega = \pi. The phase \arg(H(e^{j\omega})) starts at 0 and reaches a minimum lag of about -0.4636 radians at \omega = \pi/2, demonstrating the minimal phase distortion characteristic. For the non-minimum-phase H'(z), the magnitude scales larger (peaking at 3 at \omega = 0), and the phase lag is more negative, exceeding -1.107 radians at \omega = \pi/2.

System Properties

Causality Requirement

In linear time-invariant (LTI) systems, causality is defined as the property where the output at any given time depends solely on the current input and all past inputs, with no dependence on future inputs. This ensures that the system's impulse response h(t) (or h in discrete time) is zero for negative time arguments, i.e., h(t) = 0 for t < 0 or h = 0 for n < 0. For minimum-phase systems, the placement of zeros plays a critical role in guaranteeing a causal inverse. In discrete-time systems, all zeros lie inside the unit circle (|z| < 1), while in continuous-time systems, all zeros are in the left-half s-plane (\Re(s) < 0). This configuration ensures that the inverse transfer function $1/H(z) or $1/H(s) has its poles (the original system's zeros) positioned such that a causal region of convergence (ROC) can be selected without introducing non-causal components in the time-domain response. Specifically, the causal ROC for the inverse is exterior to the outermost pole, and since these poles are inside the unit circle (discrete) or left-half plane (continuous), the ROC includes the unit circle, maintaining both causality and stability. A mathematical demonstration of this causality in the inverse uses partial fraction expansion of the Z-transform for discrete-time systems. Consider a rational transfer function H(z) = \frac{\prod (z - z_k)}{\prod (z - p_m)} with all |z_k| < 1 and |p_m| < 1. The inverse is H^{-1}(z) = \frac{\prod (z - p_m)}{\prod (z - z_k)}, with poles at z_k. Assuming a proper expansion, the partial fraction form is H^{-1}(z) = \sum_k \frac{A_k}{1 - z_k z^{-1}} (for the causal ROC |z| > \max |z_k| < 1), where residues A_k are computed as A_k = \prod_m (z_k - p_m) / \prod_{j \neq k} (z_k - z_j). The inverse Z-transform yields h^{-1} = \sum_k A_k z_k^n u, a sum of causal geometric sequences starting at n = 0, with no terms for n < 0 (future-dependent or anti-causal components). This absence of negative-time terms confirms the causal nature, as the ROC exterior to all poles ensures right-sided sequences. A similar expansion applies in the Laplace domain for continuous-time, where terms like e^{z_k t} u(t) (with \Re(z_k) < 0) are causal. In contrast, non-minimum-phase systems with at least one zero outside the unit circle (discrete) or in the right-half plane (continuous) result in a non-causal inverse. For example, consider the discrete-time system H(z) = 1 - 2z^{-1}, with a zero at z = 2 (|z| = 2 > 1). The inverse H^{-1}(z) = \frac{z}{z - 2} = \frac{1}{1 - 2z^{-1}} has a at z = 2. To achieve (ROC including |z| = 1), the anti-causal ROC |z| < 2 must be chosen, yielding h^{-1} = -2^n u[-n-1], an anti-causal sequence extending to n \to -\infty and requiring future inputs for realization. This illustrates how zeros outside the stability region force non-causal filtering in the inverse.

Stability Implications

For linear time-invariant (LTI) systems, bounded-input bounded-output (BIBO) stability requires that every bounded input produces a bounded output. In discrete time, this is achieved when all poles of the transfer function H(z) lie inside the unit circle in the z-plane (|z| < 1), ensuring the impulse response is absolutely summable. In continuous time, BIBO stability holds when all poles of H(s) are in the open left-half s-plane (real part < 0), making the impulse response absolutely integrable. The minimum-phase property extends this stability by requiring all zeros to also reside in the stable region: inside the unit circle for discrete-time systems or in the left-half plane for continuous-time systems. Assuming the system is already stable (poles in the stable region), the minimum-phase condition ensures the inverse system H_{\text{inv}}(z) = 1/H(z) or H_{\text{inv}}(s) = 1/H(s) has poles precisely at the original system's zeros, which are thus in the stable region, guaranteeing the inverse is also BIBO stable. This follows directly from the pole-zero structure: the denominator of the inverse transfer function consists of the numerator factors of the original H, relocating the zeros to pole positions without altering their locations relative to the stability boundary. The stability margin of the inverse is tied to the locations of these zeros; for instance, in discrete time, the minimum distance from any zero to the unit circle serves as a quantitative indicator, where greater separation enhances robustness against perturbations that could push poles toward instability. In continuous time, analogous margins relate to the minimum real part distance of zeros from the imaginary axis. This property has key practical implications, enabling stable equalization and deconvolution in signal processing applications, as the inverse can be implemented without amplifying instabilities. For contrast, all-pass systems, which are stable but possess zeros outside the unit circle (or in the right-half plane), yield unstable inverses, precluding such operations.

Frequency Domain Analysis

Discrete-Time Response

In discrete-time systems, the frequency response of a minimum-phase filter is characterized by its discrete-time Fourier transform (DTFT), H(e^{j\omega}), obtained by evaluating the z-transform H(z) on the unit circle |z| = 1. For such filters, all zeros of H(z) lie inside the unit circle in the z-plane, ensuring causality, stability, and the minimal phase shift for a given magnitude response |H(e^{j\omega})|. This placement of zeros minimizes the phase distortion compared to non-minimum-phase counterparts with equivalent magnitude characteristics. The phase response \phi(\omega) of a minimum-phase filter is uniquely determined by its magnitude response through the Hilbert transform relation in the discrete domain:
\phi(\omega) = -\mathcal{H} \left[ \ln |H(e^{j\omega})| \right],
where \mathcal{H} denotes the discrete Hilbert transform. This relation arises from the analytic properties of the log-frequency response on the unit circle and guarantees that the phase is the negative Hilbert transform of the log-magnitude, enabling direct computation of the phase from magnitude specifications in filter design.
A representative example is a second-order infinite impulse response (IIR) Butterworth lowpass filter designed via the bilinear transform with a normalized cutoff frequency of \omega_c = \pi/2. Its transfer function is
H(z) = \frac{1}{2 + \sqrt{2}} \cdot \frac{(1 + z^{-1})^2}{1 + \frac{2 - \sqrt{2}}{2 + \sqrt{2}} z^{-2}},
featuring a double zero at z = -1 (on the unit circle) and complex conjugate poles inside the unit circle. The phase response at the cutoff frequency is \phi(\pi/2) = -90^\circ, which can be computed by evaluating H(e^{j\omega}) and taking the argument, illustrating the smooth phase transition typical of such designs.
In digital filter design for sampling applications, the minimum-phase property facilitates stable reconstruction by ensuring the filter's inverse is causal and stable, allowing effective phase compensation in decimation or interpolation stages to suppress aliasing artifacts without introducing instability or excessive delay. This is particularly beneficial in multirate systems, where precise timing alignment aids in faithful signal recovery from oversampled data.

Continuous-Time Response

In continuous-time linear time-invariant systems, the frequency response is characterized by the H(j\omega), obtained by evaluating the H(s) along the imaginary axis s = j\omega. For a minimum-phase system, which is causal and stable, all zeros of H(s) lie in the open left-half of the s-plane, ensuring that H(j\omega) has no zeros in the right-half plane and thus maintains analyticity in the upper half-plane consistent with causality. This placement of zeros minimizes the phase contribution from the numerator, distinguishing minimum-phase systems from those with right-half-plane zeros that introduce additional phase lag. The Paley-Wiener criterion, adapted to minimum-phase systems, posits that the logarithm of the magnitude response \ln |H(j\omega)| must be Hilbert transformable without singularities in the complex plane, guaranteeing that the phase can be uniquely recovered from the magnitude via a causal relationship. This condition arises from the analyticity of \ln H(s) in the right-half plane for minimum-phase transfer functions, preventing logarithmic branch points that would occur with right-half-plane zeros. Consequently, the criterion ensures the frequency response satisfies the necessary integrability for the Hilbert transform to converge, linking the real part (log-magnitude) and imaginary part (phase) of \ln H(j\omega). The phase \phi(\omega) of a minimum-phase system is derived directly from the magnitude as the negative Hilbert transform of the log-magnitude: \phi(\omega) = -\frac{1}{\pi} \mathrm{P.V.} \int_{-\infty}^{\infty} \frac{\ln |H(j\nu)|}{\nu - \omega} \, d\nu, where P.V. denotes the Cauchy principal value integral. This formula encapsulates the minimum-phase property, as it yields the unique phase that concentrates the system's energy earliest in time among all causal stable systems with the same |H(j\omega)|. The derivation follows from the , which stem from the contour integral of \ln H(s) around the imaginary axis, assuming no singularities in the right-half plane. In analog filter design, the exemplifies a minimum-phase response, featuring all poles in the left-half plane and no finite zeros, resulting in a maximally flat magnitude approximation with inherently minimal phase distortion. Compared to non-minimum-phase designs, such as those augmented with all-pass networks to achieve linear phase (e.g., for reduced group delay variation), exhibit lower overall phase lag for equivalent magnitude roll-off, though at the cost of nonlinear phase across the passband that introduces some transient distortion. This trade-off highlights the minimum-phase advantage in applications prioritizing early signal arrival over constant delay, such as in audio or control systems where phase linearity is secondary to magnitude fidelity.

Magnitude-Phase Relationship

In causal and stable linear time-invariant systems, the phase response \phi(\omega) is uniquely determined by the magnitude response |H(j\omega)| through Bode's gain-phase relationship, which arises from the analyticity of the transfer function in the right-half complex plane. This connection is expressed via the , where the logarithm of the magnitude and the phase form a , ensuring that the phase at any frequency is an integral over the entire magnitude spectrum. For minimum-phase systems specifically, this relationship achieves uniqueness because all phase information is fully encoded in the magnitude response, without additional contributions from non-minimum-phase factors. In contrast, non-minimum-phase systems incorporate extra phase lag through all-pass filters, which have unit magnitude but non-trivial phase, altering the overall response beyond what the magnitude alone dictates. The minimum-phase property guarantees the smallest possible phase lag (in absolute value) for a given magnitude, as any zeros in the right-half plane would introduce irreducible delays. The precise mathematical form of this relation is given by the for the phase: \phi(\omega) = -\frac{1}{\pi} \mathcal{P} \int_{-\infty}^{\infty} \frac{\ln |H(j\nu)|}{\nu - \omega} \, d\nu, where \mathcal{P} denotes the Cauchy principal value. This integral directly links \ln |H(j\omega)| and \phi(\omega) as , valid under the assumptions of causality, stability, and minimum phase. The proof of uniqueness follows from the unique factorization theorem for analytic functions: any stable causal transfer function H(s) can be decomposed as H(s) = H_{\min}(s) \cdot H_{\text{ap}}(s), where H_{\min}(s) is minimum-phase (with all zeros in the left-half plane) and H_{\text{ap}}(s) is an ; for minimum-phase systems, H_{\text{ap}}(s) = 1, making the phase solely dependent on the magnitude. A modern extension of this relationship applies to broadband signals, where direct Hilbert transform computation can be inefficient due to the wide frequency range. Cepstral analysis provides an alternative for phase recovery by exploiting the minimum-phase property in the time domain: the complex cepstrum of a minimum-phase signal is causal and real-valued, allowing reconstruction of the phase from the magnitude spectrum via inverse Fourier transform of the log-magnitude, followed by liftering to isolate minimum-phase components. This method is particularly effective for broadband applications, such as seismic or acoustic signal processing, where it enables minimum-phase equivalent reconstruction without explicit integration over infinite limits.

Time Domain Characteristics

Impulse Response Features

The impulse response of a minimum-phase system exhibits a distinctive concentration of energy toward the beginning of the response, distinguishing it from other systems with the same magnitude response. In discrete-time systems, this property is quantified by the partial energy E = \sum_{k=0}^{n} |h|^2, which is maximized for each n among all causal stable systems sharing the same |H(e^{j\omega})|. This rapid energy accumulation implies a faster initial buildup in the impulse response compared to non-minimum-phase counterparts. In continuous-time systems, the analogous characteristic holds, where the energy \int_{0}^{T} |h(t)|^2 \, dt is maximized for a given duration T and fixed magnitude response |H(j\omega)|, concentrating the response energy near the time origin. This energy concentration arises from fundamental mathematical properties observable through tools like the cepstrum and autocorrelation. The complex cepstrum of a minimum-phase impulse response is causal and real-valued (for real signals), with its energy predominantly in low quefrencies, reflecting the minimal phase lag and leading to a compact, early-peaking time-domain profile. Similarly, the autocorrelation function of a minimum-phase response inherits minimum-phase characteristics, resulting in a narrower main lobe and reduced sidelobes, which contributes to a shorter rise time relative to systems with the same power spectrum. A representative example illustrates these features in finite impulse response (FIR) filters. Consider a low-pass FIR filter designed with a fixed magnitude response: the minimum-phase version concentrates its impulse response energy immediately after the onset, exhibiting a sharp rise without precursors, whereas the linear-phase counterpart displays symmetric ringing that extends before the main peak (pre-ringing), delaying the energy buildup. This absence of pre-ringing in minimum-phase responses enhances temporal sharpness, making them preferable in applications requiring quick signal onset, such as audio processing.

Group Delay Minimization

The group delay of a system is defined as \tau(\omega) = -\frac{d\phi(\omega)}{d\omega}, where \phi(\omega) is the phase response, representing the time delay experienced by the envelope of a narrowband signal at frequency \omega. In continuous-time systems, minimum-phase systems have all poles and zeros in the open left-half plane, ensuring both the system and its inverse are causal and stable while achieving the minimum possible group delay for a given magnitude response. This minimality arises from the all-pass decomposition of a general stable causal transfer function H(s), expressed as H(s) = H_{\min}(s) \cdot H_{\text{ap}}(s), where H_{\min}(s) is the minimum-phase component with the same magnitude response |H_{\min}(j\omega)| = |H(j\omega)| and all zeros in the left-half plane, and H_{\text{ap}}(s) is an all-pass factor with poles and zeros symmetric across the imaginary axis. The all-pass component introduces only positive group delay \tau_{\text{ap}}(\omega) > 0 for all \omega, as its phase \phi_{\text{ap}}(\omega) is monotonically decreasing, such that the total group delay \tau(\omega) = \tau_{\min}(\omega) + \tau_{\text{ap}}(\omega) is strictly greater than \tau_{\min}(\omega) unless H_{\text{ap}}(s) is constant. For minimum-phase systems, the minimum group delay is given by \tau(\omega) = \mathcal{H}\left[ \frac{d \ln |H(j\omega)|}{d\omega} \right], where \mathcal{H} denotes the , linking the group delay directly to the rate of change of the log-magnitude response via the minimum-phase property. Minimum-phase systems are applied in audio processing to minimize in real-time equalization and enhancement, where low group delay preserves temporal alignment without the added delay of linear-phase alternatives.

Contrasting Systems

Maximum-Phase Systems

Maximum-phase systems represent the counterpart to minimum-phase systems in linear time-invariant (LTI) signal processing. In the discrete-time domain, a causal and stable maximum-phase system has all its zeros located outside the unit circle in the z-plane, while in the continuous-time domain, all zeros lie in the right-half s-plane. This configuration ensures that the system's inverse is stable but anti-causal, meaning the inverse impulse response extends backward in time while remaining bounded. A key property of maximum-phase systems is their maximum phase lag among all causal stable systems sharing the same magnitude response; this lag is greater than that of their minimum-phase equivalents. The impulse response of such systems concentrates energy toward the tail, delaying the onset of significant output relative to minimum-phase counterparts. For systems with identical magnitude responses |H(ω)|, the phase response satisfies φ(ω) = -φ_min(ω), where φ_min(ω) is the phase of the corresponding minimum-phase system, reflecting the reflected zero locations across the unit circle. An illustrative example is the zero-flipped version of a minimum-phase , obtained by reflecting all zeros across the (replacing each zero at a with 1/\bar{a}). For instance, starting from a minimum-phase with zeros inside the , the maximum-phase variant exhibits the same frequency magnitude but with reversed in sign, resulting in an where oscillations and ringing are shifted to the latter part of the response, increasing perceived delay and tail energy.

Mixed-Phase Systems

Mixed-phase systems in are characterized by transfer functions with zeros located both inside and outside the unit circle in the z-plane, leading to a that is only partially recoverable from the response alone. This hybrid zero placement distinguishes them from purely minimum-phase systems, where all zeros lie inside the unit circle, and maximum-phase systems, where all zeros are outside. A key property of mixed-phase systems is their decomposition into a minimum-phase component and an all-pass factor, expressed as H(z) = H_{\min}(z) \cdot H_{ap}(z), where H_{\min}(z) contains all zeros reflected inside the unit circle for minimum phase, and H_{ap}(z) is a stable all-pass filter that preserves the magnitude response while adjusting the phase to match the original system. This factorization is applicable to any causal and stable filter without zeros on the unit circle. Such systems represent a compromise in terms of and for inversion: while the forward system remains causal and stable (with poles inside the ), the inverse inherits non-minimum-phase characteristics from the all-pass component, complicating full recovery without additional delays or approximations. In practical applications, mixed-phase behavior is prominent in acoustics modeling, where the overall combines direct sound paths with multipath reflections, necessitating hybrid equalization techniques to address both magnitude and distortions. A representative example is the acoustic in an enclosed space, where early arrivals—such as the direct and initial reflections—exhibit minimum-phase properties due to their proximity and , while later reverberant tails introduce non-minimum-phase elements from diffuse and overlapping paths, yielding a mixed-phase profile overall.

Linear-Phase Systems

Linear-phase systems represent a specific of non-minimum-phase filters in where the is a of , expressed as \phi(\omega) = -\alpha \omega, with \alpha denoting a positive constant delay parameter. This linear phase characteristic arises from the symmetry of the system's , which is even about its midpoint at n = \alpha, ensuring that the response h = h[2\alpha - n] for discrete-time systems. Such symmetry distinguishes linear-phase systems from purely minimum-phase ones, as it imposes constraints on the placement of zeros in the . In relation to minimum-phase systems, linear-phase behavior can be achieved by augmenting a minimum-phase —whose zeros lie inside the —with an that introduces additional zeros in reciprocal conjugate pairs across the unit circle. For instance, if a minimum-phase system has a zero at z = r e^{j\theta} with r < 1, the all-pass component adds a corresponding zero at z = (1/r) e^{-j\theta} outside the unit circle, preserving the magnitude response while linearizing the . This construction ensures the overall system is not minimum phase but maintains exact linear through balanced zero-pole reciprocity. A key property of linear-phase systems is their constant group delay, given by \tau(\omega) = \alpha, which introduces a uniform time shift across all frequencies without distorting the waveform's shape. This constant delay, derived from the negative derivative of the phase with respect to frequency, contrasts with the variable group delay in minimum-phase systems and minimizes dispersion effects. Linear-phase systems find applications in phase-sensitive domains such as image processing, where symmetry preservation is crucial to avoid visually distorting artifacts from nonlinear phase shifts. A representative example is the design of symmetric finite impulse response (FIR) filters, such as Type I filters with odd-length symmetric coefficients, which are employed for low-pass filtering in image enhancement while ensuring constant delay and exact linear phase.

References

  1. [1]
    [PDF] All-Pass Systems - MIT OpenCourseWare
    1. Page 2. Minimum-Phase Systems. The basic definition of a minimum-phase system is as follows: Stable & causal and has a stable & causal inverse.
  2. [2]
    [PDF] Digital Signal Processing Minimum-Phase Systems - spinlab
    When we say a system is “minimum phase”, we mean that it has the least phase delay (or least phase lag) among all systems with the same magnitude response. H1(z)= ...
  3. [3]
    [PDF] ECE 380: Control Systems - Purdue Engineering
    Note that for minimum phase systems, the magnitude plot uniquely determines the transfer function, but for nonminimum phase systems, we need both the magnitude ...
  4. [4]
    What is the true meaning of a minimum phase system?
    May 2, 2012 · The relation of "minimum" to "phase" in a minimum phase system or filter can be seen if you plot the unwrapped phase against frequency.On what it means for a system to be minimum phaseanalytic signal - Definition of minimum-phase systemMore results from dsp.stackexchange.com
  5. [5]
    [PDF] Loop Analysis - ResearchGate
    The distinction is important in practice because minimum phase systems are easier to control than systems with larger phase lag. We will now give a few ...<|control11|><|separator|>
  6. [6]
    [PDF] Minimum Phase and Allpass Systems
    A system function H(z) is said to be a minimum phase system if all of its poles and zeros are within the unit circle. Consider a causal and stable LTI ...
  7. [7]
    [PDF] EE120 - Fall'15 - Lecture 1 Notes1 - People @EECS
    Nov 23, 2015 · Minimum Phase Systems. A stable and causal LTI system is called minimum phase if all of its zeros are in the open left half-plane. Page 101 ...<|control11|><|separator|>
  8. [8]
    [PDF] Network Analysis and Feedback Amplifier Design
    This book was originally written as an informal mimeographed text for one of the so-called “ Out-of-Hour ” courses at Bell Telephone Labora¬ tories.
  9. [9]
    Minimum-Phase Filters | Introduction to Digital Filters
    Since poles on the unit circle are sometimes called ``marginally stable,'' we could say that zeros on the unit circle are ``marginally minimum and/or maximum ...
  10. [10]
    [PDF] ECE503: Digital Signal Processing Lecture 5 - spinlab
    Feb 13, 2012 · A causal stable system H with transfer function H(z) with all zeros inside the unit circle is called minimum phase. Definition. A causal stable ...
  11. [11]
    [PDF] Discrete-Time Systems, LTI Systems, and Discrete-Time Convolution
    Definition: A system is said to be causal if and only if its output depends ... The class of causal discrete-time Linear, Time-Invariant (LTI) systems ...
  12. [12]
    [PDF] Lecture Notes 5 - ECEN 314: Signals and Systems
    Property 5 (Causality). A DT LTI system is causal if and only if its unit impulse response h[n]=0 for all n < 0. Proof. If a DT LTI system is causal, then ...
  13. [13]
    [PDF] Chapter 7 z-Transform
    Applying a partial-fraction expansion to H(z) give. 1. 1. )4/3(1. 1. )2/1(1. 2. )( ... 7.18, is termed a minimum-phase system. ♧ The phase response of a minimum-phase ...
  14. [14]
    [PDF] SIGNALS, SYSTEMS, and INFERENCE — Class Notes for 6.011
    3.3 ALL-PASS AND MINIMUM-PHASE SYSTEMS. Two particularly interesting classes of stable LTI systems are all-pass systems and minimum-phase systems. We define ...
  15. [15]
    [PDF] Theory Problems
    It can be shown that the minimum-phase spectrum corresponding to any magnitude spectrum is given by the Hilbert transform of the log-magnitude spectrum.2 The ...
  16. [16]
    Example: Second-Order Butterworth Lowpass - DSPRelated.com
    The digital frequency response at one-fourth the sampling rate is $\displaystyle H_d(j) = \frac{(1-j)^2}{2+\sqrt{2} - (2-\sqrt{2})} = -\frac
  17. [17]
    [PDF] Transform Representation of Signals and LTI Systems
    If we further know or assume that the system is minimum-phase, then the left-half-plane zeros from each mirrored pair are assigned to H(s), and the right-half- ...
  18. [18]
    [PDF] Name: Answers Grade: Q1 Q2 Q3 Q4 Total Mean: 66.1, Stdev: 14.4
    Since Hc(s) is minimum phase s0 = σ + jΩ and σ < 0. Therefore |z0| < 1 and all zeros and poles will be inside the unit circle, so the DT filter will be minimum.
  19. [19]
    Hilbert transform relations for complex signals - ScienceDirect
    The log-magnitude and phase, or real and imaginary parts of the Fourier transform of a minimum phase causal signal are related through Hilbert transform.
  20. [20]
    [PDF] Hilbert Transform and Applications - IntechOpen
    Apr 25, 2012 · Note that any transfer function H(jω) can be decomposed logarithmically into magnitude and phase: log H(jω) = log |H(jω)| + jZH(jω).
  21. [21]
    [PDF] The Importance of Phase in Signals
    It is well known that this is possible under certain assumptions, such as when the signal is minimum phase. Under this assumption, the Hilbert transform can be.
  22. [22]
    Section 4: Linear and Minimum Phase Filters, Recursive Digital ...
    Equivalently, only LTI filters exhibit no intermodulation distortion ... Lecture Videos. Butterworth Filter Properties: Maximum Flatness at Infinity, Low Ringing, ...
  23. [23]
    [PDF] Kramers–Kronig, Bode, and the meaning of zero
    Because causality precludes a phase advance, we conclude that the Bode relation gives the minimum phase lag. Such minimum-phase response functions have the ...<|control11|><|separator|>
  24. [24]
    [PDF] Minimum-phase signal calculation using the real cepstrum - CREWES
    We wish to have an alternative method to the Hilbert transform to convert a signal of arbi- trary phase to its minimum phase equivalent, while retaining the ...
  25. [25]
    Stability of Equation Error Designs - Stanford CCRMA
    In continuous time, the impulse is a narrow, unit-area ... minimum phase gives the maximum concentration of impulse-response energy near the time origin ...
  26. [26]
    Linear Phase Audio Filters - Stanford CCRMA
    In the minimum-phase case, all filter ringing occurs after the main pulse, while in the zero-phase case, it is equally divided before and after the main pulse ( ...
  27. [27]
    [PDF] Digital Signal Processing Phase and Group Delay of LTI Systems
    where τg(x) := − dθ(ω) dω ω=x is called the “group delay” (in samples) at normalized frequency x. D. Richard Brown III. 5 / 7. Page 6. DSP: Phase ...Missing: τ( | Show results with:τ(
  28. [28]
    [PDF] Minimum-Phase & All-Pass Filters - Multimedia Signal Processing
    Mar 11, 2011 · Since an all-pass filter has a positive group delay, a minimum-phase filter has the smallest group delay in the family of filters with the same ...
  29. [29]
    [PDF] Design of Optimal Minimum Phase Digital FIR Filters Using Discrete ...
    Mar 23, 2000 · A linear phase FIR filter has zeros inside, on, and outside the unit circle, whereas a minimum phase FIR filter has all of its zeros on or ...
  30. [30]
    Performance Analysis of IIR and FIR Filters for 5G Wireless Networks
    Jun 5, 2020 · We evaluate these filters in terms of magnitude response, phase response and group delay, and identify the minimum filter order that ...
  31. [31]
    [PDF] NUMERICAL ANALYSIS OF RIGtrF-HALF PLANE ZEROS FOR A ...
    As mentioned before, a system is considered nonminimum phase if there are system zeros or poles located in the right half of the Laplace plane. Figures 2. la ...<|control11|><|separator|>
  32. [32]
    Illustrations of spectra - SEG Wiki
    May 3, 2021 · ... negative of the other. In particular, the phase-lag spectrum is the ... The minimum-phase spectrum (thick line) and the maximum-phase ...
  33. [33]
  34. [34]
    [PDF] ELC 4351: Digital Signal Processing - Baylor University
    ▷ A system is called a minimum-phase system if it has the minimum group delay of the set of systems that have the same magnitude response. ▷ A system is called ...
  35. [35]
    Minimum-Phase/Allpass Decomposition - Stanford CCRMA
    Every causal stable filter $ H(z)$ with no zeros on the unit circle can be factored into a minimum-phase filter in cascade with a causal stable allpass filter.Missing: mixed | Show results with:mixed
  36. [36]
    [PDF] Digital Signal Processing Minimum-Phase All-Pass Decomposition
    Does it matter which one we pick? To satisfy the minimum phase delay property, recall that we require. ZHmin(ej0)=0. Note that. Hmin(ej0) = 1 - 2. 1 + 1. 3. < 0 ...Missing: group | Show results with:group
  37. [37]
    Minimum Phase - REW
    Identifying Minimum Phase Regions. Room responses are mixed phase, meaning there are some minimum phase regions and some regions that are not minimum phase.Missing: modeling | Show results with:modeling<|control11|><|separator|>
  38. [38]
  39. [39]
    [PDF] Allpass, Minimum Phase, and Linear Phase Systems
    In this brief note we discuss three special cases: allpass, minimum phase, and linear phase systems. ... This means that H(ejω), the DTFT of h[n], would be the ...Missing: definition | Show results with:definition<|control11|><|separator|>
  40. [40]
    [PDF] LINEAR-PHASE FIR FILTERS 1. The amplitude response 2. Why ...
    The LTI system has the effect of scaling the cosine signal and de- laying it by −θ(ω1)/ω1. When does the system delay cosine signals with different frequencies ...