Fact-checked by Grok 2 weeks ago

Butterworth filter

The Butterworth filter is a type of filter designed to have a maximally flat in the , providing uniform gain across the desired frequency range without ripples. This characteristic makes it ideal for applications requiring smooth attenuation from the to the , with the transition sharpness controlled by the filter order N. First described in 1930 by British engineer Stephen Butterworth in his seminal paper "On the Theory of Filter Amplifiers," published in Experimental Wireless and the Wireless Engineer, the filter derives its name from its inventor and emphasizes optimal flatness in magnitude response for analog amplifier circuits. The magnitude squared of the transfer function for a low-pass Butterworth filter is given by |H(\omega)|^2 = \frac{1}{1 + (\omega / \omega_c)^{2N}}, where \omega_c is the (typically defined at -3 attenuation) and N is the , which determines the rate of -20N per beyond \omega_c. Unlike filters with ripples, such as Chebyshev types, the Butterworth response is monotonic, ensuring no overshoot in the but a gentler transition to the . These filters can be realized in analog form using passive components like resistors and capacitors or active circuits with operational amplifiers, and in digital form as (IIR) filters via bilinear transformation. Butterworth filters are widely applied in audio processing for equalization and , communications for band-limiting signals, biomedical signal to smooth physiological , and control systems for stable frequency selection due to their predictable and ease of design. Variants include high-pass, band-pass, and band-stop configurations, all sharing the core maximally flat property, making them a foundational tool in despite trade-offs in steeper compared to elliptic filters.

History

Invention and Original Publication

The Butterworth filter was invented by British physicist and engineer Stephen Butterworth, who introduced the concept in his seminal 1930 paper titled "On the Theory of Filter Amplifiers," published in the journal Experimental Wireless and the Wireless Engineer (volume 7, pages 536–541). Working at the Admiralty Research Laboratory, Butterworth addressed the limitations of existing electrical wave filter theories, which were primarily focused on sharp cutoffs but often resulted in irregular responses unsuitable for practical applications. The primary motivation for Butterworth's work stemmed from the demands of early radio engineering, where amplifiers required frequency responses that were as smooth and uniform as possible to avoid in for wireless communication systems. At the time, radio technology was rapidly advancing, and filter amplifiers needed to maintain consistent gain across the to ensure reliable and reception of signals without unwanted variations. Butterworth's approach prioritized this uniformity over abrupt transitions, marking a shift toward filters optimized for integration in radio circuits. In the paper, Butterworth derived the foundational squared magnitude response of the filter, given by |H(j\omega)|^2 = \frac{1}{1 + \left( \frac{\omega}{\omega_c} \right)^{2N}} where \omega is the , \omega_c is the cutoff angular frequency, and N is the filter order. This expression ensures a maximally flat response in the , with the filter attenuating higher frequencies gradually based on the order N. Butterworth illustrated this with plots for various orders, demonstrating how increasing N steepens the while preserving passband flatness, a concept he explicitly tied to practical design needs in .

Development and Naming

Following its introduction in 1930, the Butterworth filter saw widespread adoption in and during the ensuing decades. Butterworth's original work on filter amplifiers for radio applications provided a foundational basis for subsequent developments in . The filter became known as the "Butterworth filter" in recognition of Stephen Butterworth's pioneering contribution, with the nomenclature formalized and consistently applied in engineering literature by the mid-20th century. This evolution extended to its integration into techniques during the 1940s, which embedded the Butterworth approach within systematic methodologies, shaping subsequent theoretical progress in the field.

Fundamental Characteristics

Overview of Filter Behavior

The Butterworth filter is a linear time-invariant filter designed to provide a maximally flat in the , ensuring minimal distortion of signals within the desired frequency range. This characteristic makes it particularly suitable for applications requiring smooth amplitude preservation without ripples or oscillations in the . Invented by British engineer Stephen Butterworth in 1930, it serves as a foundational in analog and filtering. Butterworth filters are available in several configurations, including low-pass, high-pass, band-pass, and band-stop types, each derived from transformations of the fundamental low-pass prototype. The low-pass variant, which attenuates frequencies above a specified while passing lower frequencies, is the most commonly referenced and serves as the basis for designing the other forms. These configurations allow for versatile frequency shaping in various systems. In operation, the Butterworth filter exhibits a smooth, monotonic transition from the passband to the stopband, avoiding abrupt changes or ripples that could introduce unwanted artifacts in the output signal. This gradual provides an intuitive balance between selectivity and simplicity, making it ideal for audio processing—such as equalizers and —and general tasks where preserving the natural waveform integrity is essential, particularly in environments intolerant to passband variations.

Maximal Flatness Property

The maximal flatness property of the Butterworth low-pass filter refers to the condition where the squared magnitude response |H(j\omega)|^2 and its first $2N-1 derivatives with respect to the normalized \omega are all zero at \omega = 0 for an Nth-order filter. This ensures that the passband response remains as constant as possible near (DC), with no ripples or variations in gain, distinguishing it from other filter approximations like Chebyshev types that introduce equiripple behavior. The property arises from the filter's design criterion, originally proposed by Stephen Butterworth, to maximize the number of derivatives that vanish at the origin, thereby providing the "flattest" possible monotonic response in the . This flatness approximates the ideal brick-wall low-pass filter, which has a perfectly constant magnitude of unity for \omega < 1 and zero otherwise, but the Butterworth achieves it through a rational all-pole transfer function without finite zeros, leading to a smooth, gradual roll-off beyond the cutoff. Unlike filters with zeros that can introduce sharper transitions but potential ripples, the all-pole structure of the Butterworth prioritizes passband uniformity, making it suitable for applications requiring minimal distortion in the signal's low-frequency components, such as audio processing or anti-aliasing. Conceptually, the maximal flatness can be illustrated through the Taylor series expansion of the squared magnitude response around \omega = 0: |H(j\omega)|^2 = \frac{1}{1 + \omega^{2N}} = 1 - \omega^{2N} + O(\omega^{4N}), where the expansion shows that all terms up to order $2N-1 match those of the ideal constant response (unity), with the first deviation occurring at the $2Nth power of \omega. This structure maximizes the order of contact between the actual and ideal responses at DC, enhancing flatness as N increases, though at the cost of a slower transition to the stopband compared to sharper approximations.

Mathematical Description

Transfer Function

The transfer function for a continuous-time low-pass Butterworth filter of order N and cutoff angular frequency \omega_c takes the form H(s) = \frac{1}{B_N\left( \frac{s}{\omega_c} \right)}, where B_N(s) denotes the Nth-order Butterworth polynomial with B_N(0) = 1. This polynomial is constructed such that its roots (the poles of H(s)) lie in the left half of the complex s-plane, ensuring stability and the characteristic maximally flat frequency response near \omega = 0. The squared magnitude of the frequency response, |H(j\omega)|^2, is given by |H(j\omega)|^2 = \frac{1}{1 + \left( \frac{\omega}{\omega_c} \right)^{2N}}. This form arises from the design requirement for maximal flatness in the passband, where the first $2N-1 derivatives of |H(j\omega)|^2 vanish at \omega = 0. To derive the transfer function H(s), consider the analytic continuation via H(s)H(-s) = 1 / [1 + (-s^2 / \omega_c^2)^N] for the normalized case (\omega_c = 1). The poles of this expression are the $2N roots of $1 + (-1)^N s^{2N} = 0, located at s_k = e^{j \left( \frac{\pi}{2} + \frac{(2k-1)\pi}{2N} \right)} for k = 1, \dots, 2N, which are equally spaced around the unit circle in the s-plane due to the geometric symmetry of the roots of unity. Stability dictates selecting only the N poles in the left half-plane for H(s), with the constant adjusted for unity DC gain. High-pass Butterworth filters are obtained by applying the frequency transformation s \to \omega_c / s to the low-pass prototype transfer function, inverting the frequency axis while preserving the magnitude-squared response shape but shifting the cutoff to high frequencies. This substitution maps low frequencies to high frequencies and ensures the resulting H(s) retains the same order N and flatness properties.

Normalized Polynomials and Poles

The normalized Butterworth low-pass filter is defined with a cutoff frequency of ω_c = 1 rad/s, where the transfer function takes the form H(s) = 1 / B_N(s), and B_N(s) is the Nth-order with all coefficients real and positive. These polynomials are derived from the requirement of maximally flat magnitude response at ω = 0 and are monic (leading coefficient 1). For practical design, they are often expressed in factored form using first- and second-order factors corresponding to real and complex-conjugate poles, respectively. The normalized polynomials for orders N = 1 to 5 are as follows:
  • N = 1: B_1(s) = s + 1
  • N = 2: B_2(s) = s^2 + \sqrt{2} s + 1
  • N = 3: B_3(s) = (s + 1)(s^2 + s + 1) = s^3 + 2s^2 + 2s + 1
  • N = 4: B_4(s) = (s^2 + 0.765s + 1)(s^2 + 1.848s + 1) = s^4 + 2.613s^3 + 3.414s^2 + 2.613s + 1
  • N = 5: B_5(s) = (s + 1)(s^2 + 0.618s + 1)(s^2 + 1.618s + 1) = s^5 + 3.236s^4 + 5.236s^3 + 5.236s^2 + 3.236s + 1
These forms facilitate cascading in filter realizations. The roots of B_N(s) = 0, which are the poles of H(s), lie on the unit circle in the left half of the s-plane, ensuring stability since all have negative real parts. There are no finite zeros; all zeros are at infinity, making the Butterworth filter an all-pole structure. The N poles are equally spaced angularly with separation of π/N radians (180°/N), located at angles φ_k = (2k + 1)π / (2N) measured from the positive imaginary axis (or equivalently, from the negative real axis in the clockwise direction), for k = 0, 1, ..., N-1. The pole coordinates are thus s_k = e^{j(π/2 + φ_k)} = -\sin(φ_k) + j \cos(φ_k). For example, in the N=2 case, the poles are at s = -(\sqrt{2}/2) \pm j (\sqrt{2}/2). This geometric configuration arises directly from solving 1 + (-s)^{2N} = 0 for the normalized case and selecting the stable poles.

Frequency Response and Roll-off

The magnitude response of a Butterworth filter is characterized by a maximally flat passband, transitioning smoothly to the stopband with an attenuation of 3 dB precisely at the cutoff frequency \omega_c. This design ensures minimal variation in amplitude within the passband, providing a gentle initial roll-off that becomes more pronounced beyond \omega_c. For frequencies significantly higher than the cutoff (\omega \gg \omega_c), the magnitude approximates an asymptotic behavior where the attenuation increases at a rate of -20N dB per decade, with N denoting the filter order; this steepening roll-off enhances rejection of high-frequency noise while maintaining the filter's overall smoothness. The phase response of the Butterworth filter is inherently nonlinear across the frequency spectrum, which introduces variations in signal timing. Specifically, the associated group delay—defined as the negative derivative of phase with respect to angular frequency—exhibits a peak near the cutoff frequency, leading to potential distortion for broadband signals that span this region. This nonlinearity arises from the filter's pole configuration on a circular locus in the s-plane, contrasting with filters optimized for constant delay, though it remains suitable for applications prioritizing amplitude flatness over phase linearity. Asymptotically, the Butterworth filter's frequency response approximates an ideal low-pass behavior with a perfectly flat passband extending to \omega_c and an infinitely steep transition to zero gain in the stopband; in practice, higher-order implementations narrow the transition band, achieving a steeper roll-off closer to this ideal while preserving the monotonic, ripple-free profile introduced in the original formulation. This balance makes it particularly effective for scenarios requiring uniform gain in the passband without overshoot or ringing upon step inputs.

Design Principles

Determining Filter Order

The order of a Butterworth filter is selected as the smallest integer that ensures the magnitude response meets or exceeds the specified attenuation levels at the passband edge frequency \omega_p and stopband edge frequency \omega_s. This determination is based on the filter's magnitude squared response, |H(j\omega)|^2 = 1 / [1 + (\omega / \omega_c)^{2N}], where the cutoff frequency \omega_c is adjusted after selecting N to satisfy the passband requirement. The minimum order N is calculated using the formula N \geq \frac{\log \left[ \frac{10^{A_s/10} - 1}{10^{A_p/10} - 1} \right]}{2 \log \left( \frac{\omega_s}{\omega_p} \right)}, where A_p (in dB) is the maximum allowable attenuation in the passband (typically small, e.g., 0.5–3 dB), and A_s (in dB) is the minimum required attenuation in the stopband (e.g., 20–60 dB or more). This formula arises from setting the response equal to the attenuation specifications at \omega_p and \omega_s, solving for the ratio that eliminates \omega_c, and taking the ceiling to the next integer for practical implementation. To illustrate, consider specifications with A_p = 1 dB, A_s = 40 dB, \omega_p = 1 rad/s (normalized), and \omega_s = 2 rad/s. First, compute $10^{A_p/10} - 1 = 10^{0.1} - 1 \approx 0.2589 and $10^{A_s/10} - 1 = 10^4 - 1 = 9999. The ratio is $9999 / 0.2589 \approx 38630.8. Taking the base-10 logarithm gives \log_{10}(38630.8) \approx 4.587. The denominator is $2 \log_{10}(2/1) = 2 \times 0.3010 = 0.602. Thus, N \geq 4.587 / 0.602 \approx 7.62, so round up to N = 8. With this order, the actual stopband attenuation will exceed 40 dB at \omega_s, providing excess rejection that improves filter performance beyond the minimum requirements. The roll-off rate beyond the cutoff is -20N dB per decade, so selecting a higher N than the minimum yields a steeper transition band at the cost of increased complexity.

Cutoff Frequency Adjustment

The design of a Butterworth filter begins with a normalized low-pass prototype, where the cutoff frequency is set to 1 rad/s, ensuring the magnitude response exhibits maximal flatness up to this point with a -3 dB attenuation exactly at the cutoff. To scale this prototype to an arbitrary low-pass cutoff frequency \omega_c, the transformation s \to s / \omega_c is applied to the normalized transfer function H_p(s), resulting in the adjusted transfer function H(s) = H_p(s / \omega_c). This linear frequency scaling shifts the entire frequency response such that the 3 dB point occurs at \omega_c, while preserving the maximally flat passband characteristic and the -20N dB/decade roll-off rate for an Nth-order filter, as the pole locations are simply scaled by \omega_c in the s-plane. For band-pass Butterworth filters, the normalized low-pass prototype is transformed using the quadratic substitution s \to (s^2 + \omega_0^2) / (B s), where \omega_0 is the geometric center frequency and B represents the desired bandwidth (typically defined between the -3 dB points). This transformation maps the unit cutoff of the prototype to a symmetric passband around \omega_0 with width B, doubling the filter order to 2N and positioning the poles such that the magnitude response remains maximally flat in the central passband region. The 3 dB attenuation points are thereby adjusted to \omega_0 \pm B/2, maintaining the essential Butterworth properties of monotonicity and smoothness without introducing ripples./02:_Filters/2.09:_Filter_Transformations) These transformations are directly applied to the normalized Butterworth polynomials, which define the prototype denominator, enabling efficient computation of the scaled filter coefficients for both low-pass and band-pass realizations.

Non-Standard Attenuation Handling

In standard Butterworth filter design, the cutoff frequency \omega_c is defined as the point where the power attenuation reaches 3 dB, corresponding to a magnitude response of $1/\sqrt{2}. However, practical specifications often define the cutoff at a different attenuation level A dB (e.g., 1 dB or 0.5 dB) to ensure minimal deviation within the passband. To accommodate this, the filter's effective cutoff frequency must be adjusted while preserving the maximally flat response property. The adjustment involves scaling the standard 3 dB cutoff frequency \omega_c' relative to the specified cutoff frequency \omega_c (where attenuation is A dB) using the formula: \omega_c' = \omega_c \left(10^{A/10} - 1\right)^{-1/(2N)} where N is the filter order. This scaling ensures that the magnitude response |H(j\omega)| = 1 / \sqrt{1 + (\omega / \omega_c')^{2N}} meets the A dB attenuation precisely at \omega_c. The derivation follows from setting the attenuation equation A = 10 \log_{10} \left(1 + (\omega_c / \omega_c')^{2N}\right) and solving for \omega_c'. Additionally, if the design requires a specific passband gain G_0 other than unity (common in active implementations), the overall transfer function can be scaled by multiplying by G_0, without affecting the frequency response shape. This gain adjustment is applied post-pole placement and is independent of the attenuation scaling. For example, consider a first-order low-pass (N=1) with a specified cutoff \omega_c = 1 rad/s at 1 dB attenuation. Here, $10^{1/10} - 1 \approx 0.2589, so \omega_c' = 1 \times (0.2589)^{-1/2} \approx 1.965 rad/s. The adjusted filter then exhibits exactly 1 dB attenuation at 1 rad/s and 3 dB at approximately 1.965 rad/s, maintaining the standard roll-off of -20 dB/decade beyond the cutoff. This approach allows precise compliance with non-standard specifications while leveraging precomputed .

Implementation Techniques

Analog Passive Designs

Analog passive Butterworth filters are realized through LC ladder networks, where the Cauer topology provides a systematic approach to synthesizing the circuit by performing a continued fraction expansion on the driving-point impedance derived from the filter's Hurwitz polynomial. This method ensures a doubly terminated structure with source and load resistances equal to the characteristic impedance, optimizing power transfer and minimizing reflections. The ladder configuration alternates series inductors and shunt capacitors, starting with a shunt capacitor for odd-order low-pass prototypes to achieve the maximally flat passband response. For normalized low-pass designs with a cutoff frequency of 1 rad/s and characteristic impedance Z = 1 Ω, component values are obtained directly from the continued fraction coefficients, scaled such that shunt capacitors are in farads and series inductors in henries. Representative values for a third-order filter include a first shunt capacitor C_1 = 1 F, series inductor L_2 = 2 H, and terminating shunt capacitor C_3 = 1 F; these are denormalized later by frequency and impedance scaling factors for practical implementation. Higher-order filters follow similarly, with values tabulated from the expansion to maintain the required attenuation characteristics. These passive LC realizations offer advantages such as the absence of active components, enabling reliable operation at high frequencies beyond the bandwidth limits of transistors or op-amps, and inherent stability without feedback-related issues. However, they suffer from disadvantages including sensitivity to component tolerances and parasitics, which can degrade the filter's frequency response, particularly in higher-order designs requiring precise element matching.

Analog Active Designs

Analog active designs for Butterworth filters typically employ operational amplifiers (op-amps) to realize the filter transfer function, offering advantages such as high input impedance, low output impedance, and ease of integration compared to passive realizations. These designs are particularly suited for low-frequency applications where inductors are impractical, using resistor-capacitor (RC) networks combined with op-amps to approximate the desired maximally flat frequency response. The most common approach involves cascading second-order sections, as Butterworth filters of order greater than one can be factored into quadratic terms corresponding to complex conjugate pole pairs. The Sallen-Key topology, introduced by R. P. Sallen and E. L. Key in 1955, serves as the foundational building block for second-order active Butterworth filters. This configuration uses a single op-amp with an RC network to implement low-pass, high-pass, or bandpass responses, valued for its simplicity and minimal component count. For a low-pass Sallen-Key filter, the circuit consists of two resistors (R_1 and R_2) and two capacitors (C_1 and C_2) arranged such that the input signal passes through R_1 and R_2 to the non-inverting input of the op-amp, with C_1 providing feedback from the output to the junction of R_1 and R_2, and C_2 connected from the non-inverting input to ground; the op-amp is configured as a non-inverting amplifier with gain K = 1 + R_f / R_g (or unity gain if R_f = 0). The transfer function is given by H(s) = \frac{K \omega_0^2}{s^2 + \frac{\omega_0}{Q} s + \omega_0^2}, where \omega_0 is the natural frequency and Q is the quality factor. For a second-order Butterworth low-pass filter, the poles lie on the unit circle at angles of \pm 45^\circ from the negative real axis, yielding Q = 1 / \sqrt{2} \approx 0.707 to achieve the maximally flat passband response. In the unity-gain variant (K=1), component values are selected to meet this Q while setting the cutoff frequency f_c = \omega_0 / (2\pi). The natural frequency is \omega_0 = 1 / \sqrt{R_1 R_2 C_1 C_2}, and the Q-factor is Q = \sqrt{R_1 R_2 C_1 C_2} / [C_2 (R_1 + R_2)]. For example, with equal resistors R_1 = R_2 = R, the feedback capacitor is set to C_1 = 2C and the shunt capacitor to C_2 = C, yielding Q = 1 / \sqrt{2} and \omega_0 = 1 / (R C \sqrt{2}). However, the unity-gain Sallen-Key exhibits sensitivity to component tolerances for Q > 0.5, potentially leading to peaking or instability if variations exceed 5-10%; stability is improved by using the non-unity gain configuration, where K = 3 - 1/Q \approx 1.586 for Butterworth, allowing equal resistors (R_1 = R_2 = R) and capacitors (C_1 = C_2 = C) with \omega_0 = 1 / (R C), though this introduces a dc gain of 1.586 that may require additional attenuation. Higher-order Butterworth filters are constructed by cascading multiple Sallen-Key second-order sections, with buffering between stages to prevent loading effects; each section is assigned a conjugate pole pair from the filter's pole locations on the unit circle, ensuring the overall response matches the desired order N. For even orders, all sections are second-order, while odd orders include a RC stage; the Q-factors vary across sections—for instance, in a fourth-order filter, typical values are approximately 0.541 and 1.307—calculated as Q_k = 1 / [2 \sin((2k-1)\pi / (2N))] for the k-th pair, with component values scaled accordingly to the desired . This cascading approach maintains the Butterworth characteristic of -20 N dB/decade roll-off beyond the cutoff while preserving passband flatness, though careful op-amp selection is needed to handle the varying Q sensitivity, favoring low-noise, high-bandwidth devices for minimal .

Digital Realizations

Digital realizations of Butterworth filters typically involve transforming the analog prototype into a discrete-time (IIR) filter suitable for implementation on digital hardware. This conversion preserves key characteristics of the maximally flat while accounting for the discrete nature of sampled signals. Common methods include the and , each with distinct advantages in mapping the s-domain to the z-domain. The provides a one-to-one mapping from the continuous-time s-plane to the discrete-time z-plane, ensuring preservation since the left-half s-plane maps inside the unit circle in the z-plane. The transformation is given by s = \frac{2}{T} \frac{1 - z^{-1}}{1 + z^{-1}}, where T is the sampling period. To compensate for the nonlinear frequency warping introduced by this method, prewarping is applied to the : \omega_c' = \frac{2}{T} \tan\left(\frac{\omega_c T}{2}\right), where \omega_c is the desired digital . This approach is particularly effective for Butterworth filters as it avoids and maintains the monotonic without phase distortion beyond the warping effect. In contrast, the impulse invariance method designs the digital filter by sampling the impulse response of the analog prototype, resulting in h = T \cdot h_a(nT), where h_a(t) is the analog impulse response. This technique preserves the time-domain at sampling instants but is less suitable for Butterworth low-pass filters due to significant effects, as the analog Butterworth response is not strictly bandlimited and its extends to infinite frequencies. distorts the frequency response, particularly in the , making impulse invariance preferable only for applications where the analog filter's bandwidth is well below the or for approximating step responses in control systems. Once the digital transfer function H(z) is obtained, implementation often uses structures like direct form II or a cascade of second-order biquad sections to minimize and enhance . Cascading biquads decomposes the higher-order filter into paired complex-conjugate poles, each realized as H_k(z) = \frac{b_0 + b_1 z^{-1} + b_2 z^{-2}}{1 + a_1 z^{-1} + a_2 z^{-2}}, with coefficients derived from the analog poles via the chosen . This form guarantees for the digital Butterworth filter, as the analog prototype's poles in the left-half plane map to poles inside the unit circle, and the modular structure reduces sensitivity to finite-word-length effects in .

Comparisons

Versus Chebyshev Filters

The Chebyshev Type I filter features an equiripple response in the , which allows for a steeper compared to the Butterworth filter of the same order, but this comes at the cost of introducing in the passband, typically specified as 0.5 or 1 depending on design requirements. In contrast, the Butterworth filter maintains maximal flatness in the passband with no ripple, providing a smoother overall . In the transition band, the Chebyshev Type I filter achieves a sharper , enabling it to meet specifications with a lower filter order than the Butterworth filter; for instance, a third-order Chebyshev may suffice where a fourth-order Butterworth is needed for similar and performance. This sharper transition in Chebyshev filters results from the equiripple approximation, which trades uniformity for improved selectivity, whereas the Butterworth's monotonic response leads to a more gradual . Chebyshev Type I filters are preferred in bandwidth-critical applications, such as communication systems requiring sharp frequency separation to maximize , while Butterworth filters are favored in low-ripple scenarios like audio processing where passband flatness minimizes and preserves signal fidelity.

Versus Bessel Filters

Bessel filters are characterized by a maximally flat group delay, which approximates a response over a wide range in the passband, enabling better preservation of signal waveforms such as pulses or square waves without significant . In comparison, Butterworth filters exhibit a nonlinear that varies more rapidly near the , potentially introducing greater time-domain for transient signals. The of a features a broader transition band and gentler beyond the , prioritizing over sharp . This contrasts with the Butterworth filter's steeper initial , which delivers superior for equivalent filter orders, making it preferable for applications requiring effective frequency rejection. Consequently, demand higher orders than Butterworth filters to match the same level of suppression, as their magnitude response decays more slowly.

References

  1. [1]
    [PDF] Butterworth filters
    Butterworth filters were invented in 1930 by Stephen Butterworth, introduced them in his paper titled: "On the Theory of Filter Amplifiers", Wireless Engineer, ...
  2. [2]
    2.7 Chebyshev and Butterworth filters | OpenLearn - Open University
    The first is the Butterworth filter. The gain function of a Butterworth filter has the familiar flat passband and roll-off you would expect. However, this ...Missing: definition | Show results with:definition
  3. [3]
    Butterworth Lowpass Design - Stanford CCRMA
    The transfer function is defined for LTI filters as the z transform of the filter ... Butterworth filters (as well as Chebyshev and Elliptic Function filter ...
  4. [4]
    [PDF] On the Theory of Filter Amplifiers - changpuak.ch
    October, 1930. T. 536. EXPERIMENTAL WIRELESS &. On the Theory of Filter Amplifiers.*. By S. Butterworth, M.Sc. (Admiralty Research Laboratory). HE orthodox ...
  5. [5]
    [PDF] Analog Filter Design
    ➢ 1930 – Stephen Butterworth introduces maximally flat filters. (with amplifier to separate stages). ➢ 1917 – Edwin Howard Armstrong: First Superhetorodyne ...Missing: Bode | Show results with:Bode
  6. [6]
    Pro Audio Reference (B) - Audio Engineering Society
    Bode plot or Bode diagram A method developed by Hendrick Wade Bode to ... Butterworth filter A type of electronic filter characterized by having a ...Missing: Hendrik | Show results with:Hendrik<|control11|><|separator|>
  7. [7]
    [PDF] Mini Tutorial | Analog Devices
    Because it has no ripple in the pass band or the stop band, it is sometimes referred to as a maximally flat filter. The Butterworth filter achieves its.Missing: definition | Show results with:definition
  8. [8]
  9. [9]
    butter - Butterworth filter design - MATLAB - MathWorks
    Butterworth filters have a magnitude response that is maximally flat in the passband and monotonic overall. This smoothness comes at the price of decreased ...<|control11|><|separator|>
  10. [10]
    Butterworth Filter - an overview | ScienceDirect Topics
    A Butterworth filter is defined as a type of signal processing filter designed to have a flat frequency response in the passband, characterized by its ...Missing: invention motivation radio
  11. [11]
    [PDF] Ideal filter approximation and synthesis - Scholars' Mine
    The function yields a very flat characteristic for small values of w, and, for this reason, the Butterworth filter is known as maximally flat.Missing: explanation | Show results with:explanation
  12. [12]
    [PDF] Active Low-Pass Filter Design (Rev. D) - Texas Instruments
    This report focuses on active low-pass filter design using operational amplifiers. ... SLOA049D – JULY 2000 – REVISED FEBRUARY 2023. Submit Document Feedback.
  13. [13]
    [PDF] Theory of filter amplifiers
    Feb 3, 2010 · October, 1930. T. 536. EXPERIMENTAL WIRELESS &. On the Theory of Filter Amplifiers.*. By S. Butterworth, M.Sc. {Admiralty Research Laboratory) ...
  14. [14]
    Filter Design - Sidney Burrus Style - Rice Research Repository (R-3)
    2 Butterworth Filter Properties ... of flatness" at ω = 0 (DC) and at ω = ∞. ... How can I explain how strange it was to have something like Microsoft suddenly ...
  15. [15]
    [PDF] Lecture 24: Butterworth filters - MIT OpenCourseWare
    The frequency response of these filters is monotonic, and the sharpness of the transition from the passband to the stop- band is dictated by the filter order.
  16. [16]
    Butterworth Lowpass Poles and Zeros | Introduction to Digital Filters
    . These poles may be quickly found graphically by placing $ 2N$ poles uniformly distributed around the unit circle (in the $ s$ plane, not the $ z$ plane ...
  17. [17]
    Butterworth Filter Design - Electronics Tutorials
    Butterworth Filter Design has a maximally flat frequency response with no ripple and a more linear phase response in the passband.
  18. [18]
    Understanding Butterworth Filter Poles and Zeros - Technical Articles
    Sep 23, 2019 · This article explores the Butterworth low-pass filter, also known as the maximally flat filter, from the perspective of its pole-zero diagram.Missing: unit | Show results with:unit
  19. [19]
    [PDF] "Analog Reconstruction Filter for HDTV Using THS8133,THS8134 ...
    The Butterworth filter is designed to have a maximally flat ... The group delay of these filters increases during the pass band and peaks near the cutoff.
  20. [20]
    [PDF] LECTURE 3 FILTERING
    The pass band in a Butterworth filter is virtually ripple-free. However, there are two problems: 1. A non-linear phase response in the pass band rules out its ...
  21. [21]
    [PDF] Analog Filters (II) Butterworth Filters - UCSB ECE
    The Butterworth Low pass filter order N is given by ... 1. 10. 1. 10 log. 10. 10 where Ap and As are, respectively, the pass band ripple ...Missing: A_s/ A_p/
  22. [22]
    [PDF] 10.3.4 Characteristics of Commonly Used Analog Filters
    Lowpass Butterworth filters are all-pole filters characterized by ... The pole locations are most easily determined for a filter of order N by first locating.
  23. [23]
    [PDF] AI Accelerated Digital Filter Design: Butterworth, Chebyshev, Elliptic ...
    The filter cutoff frequency is 1 rad/s, and the sampling rate for ... where the normalization ˜s = s/ωc maps the desired corner frequency ωc to 1, and the.<|control11|><|separator|>
  24. [24]
    [PDF] CHAPTER 8 ANALOG FILTERS
    Butterworth and Bessel filters are examples of all-pole filters with no ripple in the pass band. Typically, one or more of the above parameters will be variable ...
  25. [25]
    [PDF] Synthesis of Optimal Ladder Networks. - DTIC
    Cauer's method is based on continuous fraction expansions. Cauer also ... Butterworth filter given the Butterworth polynomial. B6(s) s6+3.8637s5+7.464 ...
  26. [26]
    A Review and Modern Approach to LC Ladder Synthesis - MDPI
    For analogue domain filtering doubly terminated LC ladder based filter topologies ... and the Cauer 2 values are given by a continued fraction expansion around ...
  27. [27]
    [PDF] Welcome to the Cookbook Filter Guide!
    Design a Normalized Low-Pass Filter using a Table: ... These ratios are usually kept is handy tables like the one below. Element Values for Butterworth (Maximally ...
  28. [28]
    [PDF] Designing Narrow-Bandwidth Ladder Filters - Mikrocontroller.net
    Apr 7, 2019 · A N=4 Butterworth low pass would have the normalized values of C1=0.7654 Farad, L2=1.85 Henry, C3=1.85 Farad, and L4=0.7654 Henry. These ...
  29. [29]
  30. [30]
    A practical method of designing RC active filters - IEEE Xplore
    A practical method of designing RC active filters | IEEE Journals & Magazine ... R. P. Sallen; E. L. Key. All Authors. Sign In or Purchase. 334. Cites in.
  31. [31]
    [PDF] Mini Tutorial - MT-222 - Analog Devices
    The design equations for the Sallen-Key low-pass filter are shown in the Sallen-Key Low-Pass Design. Equations section. While the Sallen-Key filter is widely ...
  32. [32]
    Lecture 16: Digital Butterworth Filters | Digital Signal Processing
    Topics covered: Design of digital Butterworth filter using impulse invariance, design of digital Butterworth filter using the bilinear transformation.
  33. [33]
    [PDF] 2.161 Signal Processing: Continuous and Discrete
    (4) Use the bilinear transform to transform Hp(s) to H(z). (5) Realize the digital filter as a difference equation. Example 5. Use the bilinear transform ...
  34. [34]
    [PDF] Design of an IIR by Impulse Invariance and Bilinear Transformation
    Mar 2, 2015 · In the impulse invariant method, aliasing occurs when the prototype analog filter is transformed back into the digital filter. To reduce the ...
  35. [35]
    IIR Bandpass Filters Using Cascaded Biquads - Neil Robertson
    Apr 20, 2019 · We implemented lowpass IIR filters using a cascade of second-order IIR filters, or biquads. This post provides a Matlab function to do the same for Butterworth ...
  36. [36]
    Implementation Structures for Recursive Digital Filters
    Butterworth Lowpass Filter Example. This example illustrates the design of a 5th-order Butterworth lowpass filter, implementing it using second-order sections.The Four Direct Forms · Direct-Form I · Series And Parallel Filter...
  37. [37]
    TUTORIAL: Introduction to Filter Design
    But on the other hand Butterworth filters have a more linear phase-characteristic than do Chebyshev filters, and this may affect the choice of design for ...<|control11|><|separator|>
  38. [38]
    Butterworth Filter - an overview | ScienceDirect Topics
    The Butterworth is a class of filters that provides a maximally flat response in the passband. The pole locations for an Nth-order Butterworth filter are ...
  39. [39]
    (PDF) Chebyshev filter and Butterworth filters: Comparison and ...
    The Chebyshev filter is suited for applications that call for a high frequency response, and Chebyshev filters provide a steeper roll-down slope, which can ...
  40. [40]
    [PDF] 2.161 Signal Processing: Continuous and Discrete
    The prototype low-pass filter is based upon the magnitude-squared of the frequency response function |H(jΩ)|2. , or the frequency response power function. The ...Missing: ² | Show results with:²
  41. [41]
    [PDF] Characteristics of Commonly Used Analog Filters
    Nov 13, 1996 · Butterworth filters are maximally flat in the passband and ... Bessel filters have a linear phase response in the passband in the analog.<|control11|><|separator|>
  42. [42]
    [PDF] Lecture 3 - Transient Response and Transforms
    The main specification for Bessel-Thomson filters is usually that of a constant delay (or linear phase response) over as large a band of frequencies as possible.
  43. [43]
    [PDF] How to compare your circuit requirements to active-filter ...
    Butterworth low-pass active filter. Stephen Butterworth described the Butterworth filter in his 1930 paper titled, “On the Theory of Filter Amplifiers.” As ...
  44. [44]
    [PDF] TYPES OF ACTIVE FILTERS
    Both Butterworth and Chebyshev filters exhibit large phase ... The Bessel filter provides ideal phase characteristics with an approximately linear phase response.