Frequency response
Frequency response refers to the steady-state output of a linear system when subjected to a sinusoidal input signal, where the output is also sinusoidal at the same frequency but with potentially altered magnitude and phase.[1] This characteristic is quantified by the system's transfer function evaluated along the imaginary axis in the s-plane (s = jω), where ω represents the angular frequency.[2] For linear time-invariant (LTI) systems, the frequency response is equivalent to the Fourier transform of the system's impulse response, providing a complete description of how the system amplifies, attenuates, or shifts the phase of different frequency components in the input signal.[3] The magnitude of the frequency response indicates the gain or attenuation applied to each frequency component, while the phase response captures the time delay or advance introduced by the system.[2] These properties are typically visualized using Bode plots, which separately graph the magnitude (in decibels) and phase (in degrees) against the logarithm of frequency, facilitating analysis of system behavior across a wide range of frequencies.[1] Alternative representations, such as Nyquist plots, display the complex-valued frequency response in the complex plane to assess stability margins like gain and phase margins.[2] This frequency-domain approach contrasts with time-domain methods by revealing periodic steady-state performance without simulating transient responses. Frequency response analysis is fundamental across engineering disciplines, including electrical and electronics engineering for designing filters that selectively pass or reject specific frequency bands.[4] In control systems, it enables the prediction of closed-loop stability, disturbance rejection, and tracking performance by evaluating open-loop characteristics.[1] Applications extend to mechanical engineering for vibration analysis, acoustics for audio system design, and biomedical engineering for modeling physiological responses like hearing.[4] In power systems, it supports monitoring grid frequency deviations to ensure reliable operation.[5] Overall, this tool is essential for system identification, performance optimization, and ensuring robustness to varying input frequencies.Fundamentals
Definition
In signal processing and control systems, frequency response characterizes the steady-state behavior of a system when subjected to sinusoidal inputs of varying frequencies, specifically detailing how the system's output amplitude and phase relate to those of the input.[4] This measure reveals the system's ability to amplify, attenuate, or shift the phase of different frequency components in the input signal.[6] Unlike transient response, which captures the initial, time-varying dynamics following a sudden input change, frequency response emphasizes the long-term, periodic output after any initial transients have subsided, making it particularly suitable for analyzing periodic or steady-state signals.[7] For sinusoidal inputs, the output settles into a sinusoid of the same frequency, with modifications only in magnitude and phase determined by the input frequency.[8] A representative example is a simple RC low-pass filter, consisting of a resistor and capacitor in series, where low-frequency sine waves pass through with minimal attenuation and phase shift, approximating the input, while high-frequency sine waves experience substantial amplitude reduction and a progressive phase lag due to the capacitor's impedance.[9] This behavior illustrates how frequency response quantifies filtering effects in practical circuits. Frequency response analysis primarily applies to linear systems, where the superposition principle ensures that the overall response to a complex input can be predicted by summing the individual responses to its sinusoidal components.[10] In such systems, the response remains proportional and additive, enabling reliable decomposition of signals into frequency domains.[11]Linear time-invariant systems
A linear time-invariant (LTI) system is characterized by two fundamental properties: linearity and time-invariance. Linearity implies adherence to the superposition principle, where the system's response to a linear combination of inputs equals the linear combination of the individual responses; this includes both additivity (response to the sum of inputs is the sum of responses) and homogeneity (scaling the input scales the output proportionally). For instance, if inputs x_1(t) and x_2(t) produce outputs y_1(t) and y_2(t), then for scalars \alpha and \beta, the output to \alpha x_1(t) + \beta x_2(t) is \alpha y_1(t) + \beta y_2(t).[11][12] Time-invariance means that the system's behavior does not change over time; specifically, if an input x(t) yields output y(t), then a time-shifted input x(t - \tau) produces the correspondingly shifted output y(t - \tau) for any delay \tau. This property ensures that the system's characteristics remain consistent regardless of when the input is applied. Systems satisfying both linearity and time-invariance, such as those described by linear differential equations with constant coefficients, form the class of LTI systems.[13][11] These properties are essential for frequency response analysis because they enable the decomposition of complex signals into sinusoidal components using Fourier methods. In LTI systems, sinusoidal inputs produce sinusoidal steady-state outputs at the same frequency, with only amplitude and phase alterations, as complex exponentials are eigenfunctions of the system; this allows the overall response to be computed by summing the responses to each frequency component.[14][11] A classic example is the mass-spring-damper system, modeled by m \ddot{x}(t) + c \dot{x}(t) + k x(t) = f(t), where constant parameters m, c, and k ensure linearity and time-invariance, making it amenable to frequency-domain analysis of vibrations under harmonic forcing.[13] The conceptual foundations of LTI systems and frequency response trace back to Joseph Fourier's early 19th-century investigations into heat conduction, where he introduced Fourier series in 1822 to solve the linear partial differential equation governing heat flow in solids, establishing the basis for spectral decomposition in linear systems.[15]Mathematical Description
Transfer function
In linear time-invariant (LTI) systems, the transfer function H(s) provides a compact representation of the system's input-output relationship in the Laplace domain, defined as the ratio of the Laplace transform of the output signal Y(s) to the Laplace transform of the input signal X(s), assuming zero initial conditions:H(s) = \frac{Y(s)}{X(s)}.
This formulation applies specifically to single-input single-output (SISO) systems and facilitates analysis by converting differential equations into algebraic ones.[16] The transfer function is generally expressed as a rational function of the complex variable s, taking the form of a ratio of two polynomials:
H(s) = \frac{N(s)}{D(s)} = \frac{b_m s^m + b_{m-1} s^{m-1} + \cdots + b_0}{s^n + a_{n-1} s^{n-1} + \cdots + a_0},
where N(s) is the numerator polynomial of degree m and D(s) is the denominator polynomial of degree n (typically m \leq n for physical systems). To derive H(s), the system's time-domain differential equation is transformed via the Laplace operator. For instance, consider a second-order mass-spring-damper system governed by the equation
m \frac{d^2 x(t)}{dt^2} + c \frac{dx(t)}{dt} + k x(t) = f(t),
where m is the mass, c the damping coefficient, k the spring constant, x(t) the displacement, and f(t) the input force. Applying the Laplace transform with zero initial conditions yields
(m s^2 + c s + k) X(s) = F(s),
so the transfer function is
H(s) = \frac{X(s)}{F(s)} = \frac{1}{m s^2 + c s + k}.
This derivation highlights how the s-domain encapsulates the system's dynamics.[17][18] The poles of H(s) are the roots of the denominator polynomial D(s) = 0, which correspond to the system's eigenvalues and govern its natural response modes; for stability, all poles must lie in the open left half of the complex s-plane. The zeros are the roots of the numerator polynomial N(s) = 0, which influence the system's response by attenuating or emphasizing specific input frequencies without affecting stability directly. Additionally, the impulse response h(t), which fully characterizes the LTI system's behavior under convolution with any input, is obtained as the inverse Laplace transform of H(s): h(t) = \mathcal{L}^{-1}\{H(s)\}.[17][19]