Fact-checked by Grok 2 weeks ago

RLC circuit

An RLC circuit is an electrical circuit consisting of a resistor (R), an inductor (L), and a capacitor (C) interconnected in series or parallel configurations. These components interact to control current flow and energy storage, making the circuit a fundamental building block in electrical engineering. As a second-order linear system, it features two energy storage elements—the inductor storing energy in its magnetic field and the capacitor in its electric field—leading to dynamic behaviors such as oscillations. In a series RLC circuit driven by an alternating current (AC) source, the total impedance Z combines resistance R with frequency-dependent reactances from the inductor (X_L = ωL) and capacitor (X_C = 1/(ωC)), given by Z = √(R² + (X_L - X_C)²), where ω is the angular frequency. The current amplitude reaches a maximum at resonance, when X_L = X_C or ω_0 = 1/√(LC), minimizing impedance and enabling selective frequency response. Parallel RLC circuits exhibit similar principles but with voltages across components in common, often used for bandpass filtering. When subjected to a direct current (DC) source or transient conditions, such as closing a switch on a charged capacitor, series RLC circuits produce damped electromagnetic oscillations analogous to a mechanically damped spring-mass system. The damping depends on the resistance relative to √(L/C): underdamped cases yield decaying sinusoidal currents, critically damped provides fastest non-oscillatory return to equilibrium, and overdamped results in exponential decay without oscillation. Energy in the circuit oscillates between the inductor and capacitor, gradually dissipating as heat in the resistor. RLC circuits serve critical roles in applications including radio frequency tuning, signal filtering, and oscillation generation in electronic devices from simple receivers to advanced communication systems. Their resonant properties allow precise frequency selection, essential for bandpass filters and impedance matching in antennas. In modern contexts, they model phenomena in power electronics, sensor design, and even biological signal processing.

Fundamentals

Components and Configuration

An RLC circuit is defined as a second-order linear circuit comprising a resistor (R), an inductor (L), and a capacitor (C), which collectively model dynamic electrical behavior through their interactions. The resistor dissipates electrical energy as heat, opposing the flow of current and converting it into thermal energy. In contrast, the inductor stores energy in a magnetic field generated by the current passing through it, while the capacitor stores energy in an electric field between its plates, established by accumulated charge. These storage mechanisms in the inductor and capacitor enable oscillatory energy exchange, with the resistor introducing dissipation that influences the circuit's response. The resistance R is measured in ohms (\Omega), the inductance L in henries (\text{H}), and the capacitance C in farads (\text{F}). Standard configurations include the series RLC circuit, where the components are connected end-to-end in a single path, such that the same current flows through each, and the parallel RLC circuit, where all components share common nodes, allowing the same voltage across each while currents divide. In textual terms, a series setup can be visualized as R → L → C in a loop with a voltage source, whereas parallel arranges them branching from two nodes like spokes. Analyses of RLC circuits typically assume ideal components, meaning the inductor and capacitor exhibit no internal resistance or leakage currents beyond the explicit resistor R, ensuring losses are solely attributable to the resistor.

Governing Equations

The governing equations for RLC circuits are derived from Kirchhoff's laws, providing the mathematical foundation for analyzing both transient and steady-state behaviors in series and parallel configurations. In a series RLC circuit, Kirchhoff's voltage law (KVL) is applied around the loop, summing the voltage drops across the inductor, resistor, and capacitor to equal the applied voltage v(t). The inductor voltage is L \frac{di}{dt}, the resistor voltage is Ri, and the capacitor voltage is \frac{1}{C} \int i \, dt, yielding the integro-differential equation: L \frac{di}{dt} + Ri + \frac{1}{C} \int i \, dt = v(t). Differentiating both sides with respect to time eliminates the integral, resulting in the second-order linear differential equation for the current i(t): L \frac{d^2 i}{dt^2} + R \frac{di}{dt} + \frac{1}{C} i = \frac{dv}{dt}. For the source-free case where v(t) = 0, this simplifies to the homogeneous form: \frac{d^2 i}{dt^2} + \frac{R}{L} \frac{di}{dt} + \frac{1}{LC} i = 0. This equation governs the natural response of the circuit. For a parallel RLC circuit, Kirchhoff's current law (KCL) is used at the node, summing the currents through the capacitor, resistor, and inductor to equal the applied current i(t). The capacitor current is C \frac{dv}{dt}, the resistor current is \frac{v}{R}, and the inductor current is \frac{1}{L} \int v \, dt, giving the integro-differential equation: C \frac{dv}{dt} + \frac{v}{R} + \frac{1}{L} \int v \, dt = i(t). Differentiating with respect to time removes the integral, producing the second-order linear differential equation for the voltage v(t): C \frac{d^2 v}{dt^2} + \frac{1}{R} \frac{dv}{dt} + \frac{1}{L} v = \frac{di}{dt}. In the source-free case with i(t) = 0, the homogeneous equation is: \frac{d^2 v}{dt^2} + \frac{1}{RC} \frac{dv}{dt} + \frac{1}{LC} v = 0. This describes the circuit's natural dynamics. To solve these differential equations, initial conditions must be specified. For the series circuit, these typically include the initial current i(0) through the inductor (which cannot change instantaneously) and the initial capacitor voltage v_C(0), from which \frac{di}{dt}(0) can be determined using the differentiated equation. Equivalently, for the parallel circuit, the initial voltage v(0) across the capacitor and the initial inductor current i_L(0) are used, allowing computation of \frac{dv}{dt}(0). These conditions reflect the energy stored in the reactive elements at t=0. The coefficients in the RLC equations extend time constant concepts from simpler RC and RL circuits. In an RC circuit, the time constant \tau = RC appears in the first-order term; in an RL circuit, \tau = L/R. For RLC, the damping term \frac{R}{L} (series) or \frac{1}{RC} (parallel) generalizes this decay rate, while \frac{1}{LC} introduces oscillatory behavior analogous to inertia in mechanical systems.

Characteristic Parameters

The characteristic equation for a series RLC circuit arises from the second-order linear homogeneous differential equation governing the transient response of the current or voltage. For the inductor current i(t), the equation is \frac{d^2 i}{dt^2} + \frac{R}{L} \frac{di}{dt} + \frac{1}{LC} i = 0, leading to the characteristic equation s^2 + \frac{R}{L} s + \frac{1}{LC} = 0. s^2 + \frac{R}{L} s + \frac{1}{LC} = 0 The roots of this quadratic equation are obtained using the quadratic formula: s = -\frac{R}{2L} \pm \sqrt{\left(\frac{R}{2L}\right)^2 - \frac{1}{LC}} These roots determine the form of the natural response of the circuit. The natural undamped angular frequency \omega_0 is defined as \omega_0 = \frac{1}{\sqrt{LC}}, representing the frequency of oscillation in the absence of resistance. The damping ratio \zeta quantifies the degree of damping relative to the natural frequency and is given by \zeta = \frac{R}{2} \sqrt{\frac{C}{L}}, or equivalently \zeta = \frac{R}{2L \omega_0}. The nature of the circuit's response is classified based on the value of \zeta: if \zeta > 1, the response is overdamped with two distinct real roots; if \zeta = 1, it is critically damped with a repeated real root; and if \zeta < 1, it is underdamped with complex conjugate roots involving oscillation. For a parallel RLC circuit, the characteristic equation is derived from the differential equation for the capacitor voltage v(t): \frac{d^2 v}{dt^2} + \frac{1}{RC} \frac{dv}{dt} + \frac{1}{LC} v = 0, yielding s^2 + \frac{1}{RC} s + \frac{1}{LC} = 0. The roots are s = -\frac{1}{2RC} \pm \sqrt{\left(\frac{1}{2RC}\right)^2 - \frac{1}{LC}}. Here, the natural undamped frequency remains \omega_0 = \frac{1}{\sqrt{LC}}, but the damping ratio is \zeta = \frac{1}{2R} \sqrt{\frac{L}{C}}, or \zeta = \frac{1}{2RC \omega_0}, with the same classification criteria for \zeta applying to the response types. These parameters, particularly the roots and \zeta, govern the qualitative behavior of the transient response in both configurations.

Resonance Phenomena

Resonance Condition

In a driven RLC circuit, resonance is defined as the condition where the angular frequency of the driving voltage source, denoted as ω, matches the natural angular frequency of the circuit, ω₀ = 1/√(LC). This equality arises because the inductive reactance X_L = ωL and capacitive reactance X_C = 1/(ωC) become equal in magnitude but opposite in sign, leading to their cancellation in the total reactance. In a series RLC configuration, resonance results in minimum circuit impedance Z = R, as the reactive components nullify each other, allowing maximum current amplitude for a given driving voltage. Conversely, in a parallel RLC circuit, resonance produces maximum impedance at ω = ω₀, resulting in minimum current drawn from the source, a phenomenon sometimes termed anti-resonance. At resonance, the phase shift between the applied voltage and the circuit current is zero, meaning the current is in phase with the voltage, and the circuit behaves purely resistively. This condition facilitates peak energy oscillation between the inductor's magnetic field and the capacitor's electric field, with the resistor limiting the amplitude by dissipating energy as heat.

Natural Resonant Frequency

The natural resonant frequency, denoted as \omega_0, of an RLC circuit characterizes the inherent oscillation rate in the absence of driving forces and is given by the formula \omega_0 = \frac{1}{\sqrt{LC}}, where L is the inductance in henries and C is the capacitance in farads; this expression yields the angular frequency in radians per second. This frequency arises from the balance between the energy storage in the inductor's magnetic field and the capacitor's electric field during free oscillations. Notably, \omega_0 is independent of the resistance R, as the resistive element primarily influences energy dissipation rather than the core oscillatory dynamics. Physically, \omega_0 represents the oscillation frequency of an ideal LC circuit with zero resistance (R = 0), where energy shuttles indefinitely between the inductor and capacitor without loss, producing sustained sinusoidal behavior at this rate. In practical RLC circuits with nonzero R, the undamped natural frequency \omega_0 still serves as the reference point for the system's oscillatory tendency, though actual motion is modified by damping. For underdamped cases (where the damping ratio \zeta < 1), the observed frequency of free oscillations shifts to the damped natural frequency \omega_d = \omega_0 \sqrt{1 - \zeta^2}, with \zeta = \frac{R}{2} \sqrt{\frac{C}{L}} quantifying the relative damping strength; this adjustment accounts for the slight reduction in oscillation speed due to energy losses, but \omega_0 remains the foundational undamped value. To measure \omega_0 experimentally, one excites the circuit with an initial charge or current and observes the period T of the resulting free oscillations using an oscilloscope, then computes \omega_0 \approx \frac{2\pi}{T} (with higher accuracy for low damping where \omega_d \approx \omega_0); this method directly captures the circuit's intrinsic response by timing intervals between zero crossings or peaks in voltage or current waveforms. The natural resonant frequency also aligns with the driving frequency that maximizes response amplitude in driven RLC circuits, as detailed in the resonance condition section. Variations in \omega_0 arise primarily from manufacturing tolerances in L and C, as the frequency scales inversely with the square root of their product; for instance, a 5% tolerance in either component can shift \omega_0 by approximately 2.5%, necessitating precise component selection or trimming in applications like filters or oscillators to maintain accuracy.

Damping Effects

In free RLC circuits, damping arises from the resistance, which dissipates the energy stored in the electric and magnetic fields as heat, causing the amplitude of any oscillations to decay over time. This energy transfer between the capacitor and inductor is gradually reduced by the resistor, analogous to friction in a mechanical oscillator, preventing perpetual motion and ensuring the system returns to equilibrium. The damping ratio \zeta, a dimensionless parameter defined as \zeta = \frac{R}{2} \sqrt{\frac{C}{L}} for a series configuration, quantifies the damping level relative to the natural oscillation tendency and determines both the decay rate and the presence of oscillations. When \zeta > 1, the circuit is overdamped, resulting in a non-oscillatory exponential decay where the response slowly approaches equilibrium without crossing it. In contrast, for \zeta < 1, the underdamped case features decaying oscillations enveloped by e^{-\alpha t}, where \alpha = \zeta \omega_0 and \omega_0 = 1/\sqrt{LC} is the undamped natural frequency, leading to a gradual amplitude reduction while the circuit rings at a frequency slightly below \omega_0. At the boundary \zeta = 1, the circuit is critically damped, providing the fastest possible return to equilibrium without overshooting or oscillating, as the response follows a form that avoids any ringing. These damping regimes highlight how increasing resistance enhances energy loss per cycle, suppressing oscillations more effectively in overdamped and critically damped scenarios compared to underdamped ones. The specific transient behaviors for series RLC circuits, including detailed response forms, are covered in subsequent sections on overdamped, underdamped, and critically damped cases.

Bandwidth Calculation

In RLC circuits, the bandwidth refers to the range of angular frequencies over which the circuit's response remains significant near resonance, specifically defined as the difference between the upper and lower half-power frequencies, Δω = ω₂ - ω₁, where the power dissipated is half the maximum value at resonance (corresponding to the 3 dB points in the magnitude response). This definition arises from the frequency-dependent impedance or admittance, where the half-power condition occurs when the magnitude of the current or voltage across the resistive element is 1/√2 times its resonant value. For a series RLC circuit, the bandwidth is given by Δω = R/L, which equals the resonant angular frequency ω₀ divided by the quality factor Q. This result is derived by solving the impedance Z(ω) = R + j(ωL - 1/(ωC)) for the frequencies where |Z(ω)| = R√2, leading to a quadratic equation whose roots separate by R/L. In a parallel RLC circuit, the bandwidth is Δω = 1/(RC), also equal to ω₀ / Q. The derivation follows from the admittance Y(ω) = 1/R + j(ωC - 1/(ωL)), where the half-power points occur when |Y(ω)| = (1/R)√2, yielding the separation 1/(RC) between the roots. The half-power frequencies can be approximated for circuits with high Q as ω_{1,2} ≈ ω₀ ± (Δω)/2, providing a symmetric interval around the resonant frequency; this approximation holds well when damping is light, as the exact solutions involve square roots but simplify near resonance. A narrower bandwidth implies greater selectivity, meaning the circuit more sharply distinguishes the resonant frequency from others, which is crucial for applications requiring precise frequency response.

Quality Factor

The quality factor, denoted as Q, quantifies the sharpness of resonance in an RLC circuit and its efficiency in storing energy relative to dissipation. It is defined as the ratio of the resonant angular frequency \omega_0 to the bandwidth \Delta \omega, expressed as Q = \frac{\omega_0}{\Delta \omega}. Equivalently, in terms of energy, Q = 2\pi \times \frac{\text{maximum energy stored}}{\text{energy dissipated per cycle}}, highlighting how effectively the circuit maintains oscillatory energy. In a series RLC circuit, the quality factor is given by Q = \frac{\omega_0 L}{R} = \frac{1}{R} \sqrt{\frac{L}{C}}, where L is the inductance, C is the capacitance, and R is the resistance. For a parallel RLC circuit, it is Q = \frac{R}{\omega_0 L} = R \sqrt{\frac{C}{L}}. These expressions demonstrate that higher resistance reduces Q in series configurations while enhancing it in parallel ones, reflecting the circuit's topology. A high Q value implies sustained oscillations with minimal decay and a narrow bandwidth, enabling precise frequency selection in applications like filters. Conversely, low Q results in broader resonance and faster energy loss. The unloaded quality factor Q_U represents the intrinsic performance without external influences, determined by component losses such as Q_U = \omega_0 \frac{L}{R} for inductors or Q_U = \frac{1}{\omega_0 R C} for capacitors. External loading, such as added source or load resistances, introduces the loaded quality factor Q_L, which is always lower than Q_U and given by Q_L = \frac{Q_U Q_{\text{ext}}}{Q_U + Q_{\text{ext}}}, where Q_{\text{ext}} accounts for external dissipation; this reduction broadens the bandwidth and increases insertion loss.

Series RLC Circuit

Impedance in Series Configuration

In a series RLC circuit driven by a sinusoidal voltage source, the total impedance Z(\omega) is the phasor sum of the resistance R and the reactive components from the inductor and capacitor. The inductive reactance is j\omega L, while the capacitive reactance is -j/( \omega C ), leading to the complex impedance Z(\omega) = R + j\left( \omega L - \frac{1}{\omega C} \right), where \omega is the angular frequency. The magnitude of the impedance is given by |Z| = \sqrt{ R^2 + \left( \omega L - \frac{1}{\omega C} \right)^2 }, which represents the effective opposition to the alternating current flow. This expression shows that |Z| varies with frequency, reaching a minimum value at the resonant frequency where the reactive terms cancel. The phase angle \phi between the voltage and current is \phi = \tan^{-1} \left[ \frac{\omega L - 1/(\omega C)}{R} \right], indicating the circuit's behavior as inductive (positive \phi) for \omega > 1/\sqrt{LC} and capacitive (negative \phi) for \omega < 1/\sqrt{LC}. At resonance, when \omega L = 1/(\omega C), the imaginary part of Z vanishes, resulting in Z = R, a purely resistive impedance with \phi = 0^\circ and |Z| = R. This condition maximizes the current for a given voltage amplitude, as the circuit presents the least opposition. For completeness, the admittance Y(\omega) = 1/Z(\omega) describes the circuit's ability to conduct alternating current, though it is more commonly analyzed in parallel configurations.

Transient Response Overview

The transient response in a series RLC circuit characterizes the temporary behavior of the current i(t) or voltage across components following the application of a step or DC input, before settling to steady-state conditions. The general solution for the current is expressed as i(t) = i_h(t) + i_p(t), where i_h(t) is the homogeneous solution representing the natural response driven by initial stored energy, and i_p(t) is the particular solution capturing the forced response due to the input. The homogeneous solution i_h(t) arises from solving the characteristic equation of the second-order differential equation governing the circuit, \frac{d^2 i}{dt^2} + \frac{R}{L} \frac{di}{dt} + \frac{1}{LC} i = 0, with roots s_{1,2} = -\alpha \pm \sqrt{\alpha^2 - \omega_0^2}, where \alpha = \frac{R}{2L} is the damping factor and \omega_0 = \frac{1}{\sqrt{LC}} is the natural frequency. Depending on the discriminant \alpha^2 - \omega_0^2, the form of i_h(t) varies: for overdamped and critically damped cases (real and equal roots), it is i_h(t) = A e^{s_1 t} + B e^{s_2 t}; for the underdamped case (complex roots), it becomes i_h(t) = e^{-\alpha t} (A \cos \omega_d t + B \sin \omega_d t), with damped frequency \omega_d = \sqrt{\omega_0^2 - \alpha^2}. The damping factor \alpha sets the decay rate, influencing how quickly the transient dies out. For a DC step input of voltage V, the particular solution i_p(t) is the steady-state value \frac{V}{R}, as the inductor acts as a short and the capacitor as an open in the long term. The constants A and B are found by applying initial conditions to the full solution: the initial inductor current i(0) = I_L(0) and the initial capacitor voltage v_C(0), which relates to the derivative \frac{di(0)}{dt} = \frac{V - v_C(0) - R I_L(0)}{L}. These conditions ensure the solution matches the physical state at t = 0^+. The nature of the damping—overdamped, critically damped, or underdamped—is determined by the relative magnitude of \alpha and \omega_0, affecting the oscillatory or monotonic approach to steady state.

Overdamped Behavior

In the overdamped case of a series RLC circuit, the damping ratio ζ exceeds 1, resulting in real and distinct roots for the characteristic equation, leading to a non-oscillatory transient response. The roots are given by s_{1,2} = -\alpha \pm \sqrt{\alpha^2 - \omega_0^2}, where \alpha = \frac{R}{2L} is the damping factor and \omega_0 = \frac{1}{\sqrt{LC}} is the natural resonant frequency. Since \alpha > \omega_0, both roots are real and negative, with s_1 > s_2 (i.e., |s_1| < |s_2|), ensuring exponential decay without crossing zero. For a DC step input, the total current is i(t) = \frac{V}{R} + A e^{s_1 t} + B e^{s_2 t}, t \geq 0, where A and B are determined from initial conditions such as the initial current i(0) and initial capacitor voltage v_C(0) (via \frac{di(0)}{dt}). In a typical step voltage input scenario with zero initial conditions, such as applying a unit step across the series combination, the current rises smoothly and monotonically to the steady-state value without overshoot or oscillation, as illustrated in simulations where R = 100 \, \Omega, L = 1 \, \mathrm{H}, and C = 0.01 \, \mathrm{F} yield roots approximately at -1 and -99, resulting in a dominant time constant over about 1 second. This response exhibits a monotonic approach to steady state, characterized by two distinct time constants \tau_1 = \frac{1}{|s_1|} and \tau_2 = \frac{1}{|s_2|}, with \tau_1 > \tau_2. The longer time constant \tau_1 governs the eventual slow approach to equilibrium, making the overall settling time longer than in the critically damped case, where a single time constant \tau = \frac{1}{\alpha} applies. Physically, the overdamped behavior arises when the resistance R is sufficiently large to dominate the circuit dynamics, suppressing any oscillatory tendency from the L and C interaction, much like high friction in a mechanical mass-spring-damper system prevents bouncing and enforces a sluggish return to rest. This regime is common in applications requiring stable, non-oscillatory settling, such as certain filter designs or protective circuits.

Underdamped Behavior

In the underdamped case of a series RLC circuit, the damping ratio ζ is less than 1, leading to a transient response characterized by decaying oscillations around the steady-state value. The characteristic equation for the circuit's differential equation yields complex conjugate roots s = -\alpha \pm j \omega_d, where \alpha = \zeta \omega_0 is the damping factor, \omega_d = \omega_0 \sqrt{1 - \zeta^2} is the damped angular frequency, and \omega_0 is the natural resonant frequency. The general solution for the current in the underdamped regime is given by
i(t) = e^{-\alpha t} (A \cos \omega_d t + B \sin \omega_d t),
where the constants A and B are determined by initial conditions such as the capacitor voltage and inductor current at t = 0. This form reveals an oscillatory component modulated by an exponential decay envelope e^{-\alpha t}, with the decay rate governed by \alpha = \zeta \omega_0.
The persistence of these oscillations before significant decay can be quantified using the quality factor Q = 1/(2\zeta), where the amplitude typically reduces to $1/e of its initial value after approximately Q/\pi cycles. A practical example is the ring-down response observed when a charged capacitor is suddenly connected to the inductor and resistor, producing a damped sinusoidal waveform that gradually diminishes, as seen in laboratory demonstrations of transient behavior.

Critically Damped Behavior

In the critically damped case of a series RLC circuit, the characteristic equation has repeated real roots given by s = -\alpha, where \alpha = \frac{R}{2L} and this equals the natural resonant frequency \omega_0 = \frac{1}{\sqrt{LC}}. This condition occurs when the damping ratio \zeta = 1. The general solution for the current i(t) in the transient response takes the form i(t) = (A + B t) e^{-\alpha t}, where A and B are constants determined by initial conditions. This form arises because the repeated root requires a linear term in time multiplied by the exponential decay to satisfy the differential equation. The behavior of a critically damped series RLC circuit is characterized by a monotonic decay to equilibrium without oscillation or overshoot, representing the fastest possible return to steady state among non-oscillatory responses. The response approaches zero asymptotically, with the time-multiplied exponential ensuring a smooth transition that avoids the slower decay seen in overdamped cases. Critically damped behavior is considered optimal in many control systems applications, as it minimizes settling time while preventing overshoot and oscillations. This property makes it ideal for systems requiring rapid stabilization, such as in mechanical suspensions or electronic filters where quick response without ringing is essential. For example, consider a series RLC circuit with initial capacitor charge and zero initial current; the critically damped current response decays to zero faster than in the overdamped case, reaching near-equilibrium in the shortest time without crossing zero. In such a scenario, with L = 1 H, C = 1 F, and R = 2 Ω to achieve critical damping (\alpha = 1), the current follows i(t) = (I_0 + Q_0 t) e^{-t} (normalized units), settling effectively within a few time constants.

Parallel RLC Circuit

Admittance in Parallel Configuration

In a parallel RLC circuit driven by a sinusoidal voltage source, the total admittance Y(\omega) is the sum of the conductances and susceptances of the individual components, as admittances add directly in parallel configurations. The resistor contributes a real part G = 1/R, the capacitor a susceptance B_C = \omega C, and the inductor a susceptance B_L = -1/(\omega L). Thus, the frequency-dependent admittance is expressed as Y(\omega) = \frac{1}{R} + j \left( \omega C - \frac{1}{\omega L} \right), where \omega is the angular frequency, R is the resistance, L the inductance, and C the capacitance. The magnitude of the admittance is |Y(\omega)| = \sqrt{ \left( \frac{1}{R} \right)^2 + \left( \omega C - \frac{1}{\omega L} \right)^2 }, which quantifies the total opposition to current flow, combining conductive and reactive effects. The corresponding impedance Z(\omega) = 1/Y(\omega) reaches its maximum value at resonance, where the imaginary part of Y(\omega) vanishes, yielding Z = R; this condition occurs when \omega C = 1/(\omega L), or \omega = 1/\sqrt{LC}. The phase angle \phi of the admittance, representing the shift between voltage and total current, is given by \phi = \tan^{-1} \left[ R \left( \omega C - \frac{1}{\omega L} \right) \right]. At resonance, the admittance becomes purely real and conductive (\phi = 0), resulting in the minimum total current for a given voltage input, as the reactive currents in the inductor and capacitor cancel each other. This contrasts with the series RLC configuration, where resonance leads to maximum current.

Transient Response in Parallel

In a parallel RLC circuit, the transient response refers to the time-domain evolution of the voltage across the parallel-connected resistor, inductor, and capacitor following the removal of an excitation source, driven solely by the initial stored energy in the reactive elements. This free response is governed by a second-order linear homogeneous differential equation derived from Kirchhoff's current law, which states that the sum of the currents through each branch equals zero: \frac{d^2 v}{dt^2} + \frac{1}{RC} \frac{dv}{dt} + \frac{1}{LC} v = 0, where v(t) is the voltage across the parallel combination, R is the resistance, L the inductance, and C the capacitance. The characteristic equation associated with this differential equation is s^2 + \frac{1}{RC} s + \frac{1}{LC} = 0, with roots determining the form of the solution. The nature of the transient response—overdamped, critically damped, or underdamped—depends on the damping ratio \zeta = \frac{1}{2R} \sqrt{\frac{C}{L}}, where the natural resonant frequency is \omega_0 = \frac{1}{\sqrt{LC}} and the damping factor is \alpha = \frac{1}{2RC}. When \zeta > 1, the roots are real and distinct, yielding an overdamped response with exponential decay terms; for \zeta = 1, critically damped behavior occurs with repeated roots leading to a fastest non-oscillatory decay; and for \zeta < 1, underdamped oscillation results with complex roots, producing a damped sinusoidal voltage v(t) = e^{-\alpha t} (A_1 \cos \omega_d t + A_2 \sin \omega_d t), where \omega_d = \omega_0 \sqrt{1 - \zeta^2} is the damped frequency and the envelope decays at rate \alpha. These solution forms are analogous to those in the series RLC circuit, though here the focus is on voltage rather than current. To solve for the specific constants in the response, initial conditions must be applied: the initial voltage v(0) equals the initial capacitor voltage, reflecting stored electrostatic energy, and the initial derivative \frac{dv}{dt}(0) = -\frac{1}{C} \left( \frac{v(0)}{R} + i_L(0) \right), where i_L(0) is the initial inductor current, accounting for the magnetic energy and resistor current that influence the rate of voltage change at t=0. These conditions arise directly from the constitutive relations of the capacitor (i_C = C \frac{dv}{dt}) and inductor (v = L \frac{di_L}{dt}) combined with KCL at the instant of switching. A key distinction from the series RLC configuration lies in the topology: the voltage is the shared variable across all elements, resulting in oscillatory transients in voltage that drive branching currents through each component, whereas in series circuits, the current is shared and voltages across elements sum. This parallel arrangement emphasizes voltage-centric analysis, with the resistor providing a discharge path that influences the overall damping through the term \frac{1}{RC}.

Frequency Domain Response

In the frequency domain, a parallel RLC circuit driven by a sinusoidal voltage source exhibits resonance at the angular frequency \omega_0 = \frac{1}{\sqrt{LC}}, where the susceptances of the inductor and capacitor cancel each other out, making the total admittance purely real and equal to \frac{1}{R}. At this condition, the total circuit current I is \frac{V}{R} and flows entirely through the resistor branch, as the currents through the inductor (I_L = \frac{V}{j \omega_0 L}) and capacitor (I_C = \frac{V}{-j / (\omega_0 C)}) are equal in magnitude but opposite in phase, resulting in zero net reactive current. This current division highlights the circuit's behavior as a purely resistive load at resonance, with the magnitude of the branch currents |I_L| = |I_C| = V \sqrt{\frac{C}{L}}, which is Q times the total current, where Q is the quality factor. The power factor of the parallel RLC circuit reaches unity at resonance because the total current is in phase with the applied voltage, maximizing real power transfer to the resistor with no reactive power component. This unity power factor condition arises directly from the admittance being real-valued at \omega_0, as derived in the parallel configuration analysis. The selectivity of the circuit is characterized by the curve of the branch current magnitude |I| (through L or C) versus angular frequency \omega, which exhibits a peak at \omega_0 under voltage drive, reflecting the circuit's ability to preferentially respond to the resonant frequency. The sharpness of this peak is governed by the quality factor Q = R \sqrt{\frac{C}{L}} = \omega_0 R C, expressed in terms of admittance parameters, where higher Q values indicate narrower bandwidth \Delta \omega = \frac{\omega_0}{Q} and greater frequency discrimination; this Q formulation mirrors the series case but applies to the parallel tank's impedance magnification. As a tank circuit, the parallel RLC configuration provides high impedance at resonance (Z = R), effectively isolating signals by minimizing current draw from the source while allowing large circulating currents between the inductor and capacitor branches, a property essential for applications requiring frequency-selective energy storage.

Advanced Configurations

Ladder and Pi Networks

Ladder networks in RLC circuits consist of alternating series and shunt branches composed of resistors, inductors, and capacitors, forming a cascaded structure that extends beyond simple series or parallel configurations. This topology is commonly employed in the design of filters and delay lines, where the series elements typically carry the signal current and the shunt elements provide grounding paths. The basic building block repeats a pattern of a series impedance followed by a parallel (shunt) impedance, allowing for multi-stage implementations that approximate distributed systems like transmission lines. Analysis of ladder networks proceeds by cascading the impedances of individual sections, computing the overall input impedance or transfer function iteratively from the load end toward the source. For an RLC ladder, the series arms may include inductors or resistors in combination with capacitors, while shunt arms often feature parallel LC combinations for resonance effects. This modular approach facilitates synthesis using methods like signal-flow graphs to realize specified voltage transfer functions without requiring mutual coupling. Key advantages include a broadband frequency response due to the distributed element nature, which reduces sensitivity to individual component tolerances, and the ability to achieve controlled phase shifts through frequency transformations in the design process. Pi networks, named for their resemblance to the Greek letter π, incorporate two shunt RLC elements connected by a central series element, serving primarily as impedance matching circuits in applications requiring transformation between source and load resistances. The shunt branches are typically capacitors or inductors providing reactive loading, while the series arm adjusts the overall reactance for conjugation. This configuration allows for greater flexibility in matching real impedances with differing magnitudes, often outperforming single-stage L networks by enabling adjustment of the circuit's quality factor (Q) to optimize bandwidth. For instance, higher Q values narrow the bandwidth but enhance harmonic attenuation, making pi networks suitable for RF amplifier output stages. The dual of the pi network is the T network, featuring two series elements bridged by a central shunt element, which is particularly effective for impedance transformations where the intermediate virtual resistance must exceed the source and load values. In RLC realizations, the series arms might combine inductors and resistors, with the shunt providing capacitive or inductive termination. Analysis treats both pi and T networks as back-to-back L sections, solving for element values to achieve the desired impedance at a specific frequency while maintaining passivity. These networks offer advantages in broadband operation and phase control, as the additional degree of freedom allows tuning for wider frequency ranges without excessive insertion loss.

Coupled RLC Systems

Coupled RLC systems extend the basic RLC circuit model by incorporating magnetic coupling between inductors, enabling interactions similar to those in transformers. In such configurations, two or more RLC loops share mutual inductance, where the changing current in one inductor induces a voltage in the adjacent inductor through their overlapping magnetic fields. The mutual inductance M quantifies this interaction and is defined for two inductors with self-inductances L_1 and L_2 by the coupling coefficient k = \frac{M}{\sqrt{L_1 L_2}}, where $0 \leq k \leq 1; perfect coupling occurs at k = 1, while k = 0 indicates no interaction. The governing equations for two series RLC circuits with mutual coupling derive from Kirchhoff's voltage law, accounting for both self- and mutually induced voltages. For the primary circuit, the voltage balance is V_1 = R_1 I_1 + L_1 \frac{dI_1}{dt} + M \frac{dI_2}{dt} + \frac{1}{C_1} \int I_1 \, dt, and for the secondary circuit, V_2 = R_2 I_2 + L_2 \frac{dI_2}{dt} + M \frac{dI_1}{dt} + \frac{1}{C_2} \int I_2 \, dt. These coupled differential equations describe how currents in one loop influence the other, leading to interdependent transient and steady-state behaviors. In the frequency domain, particularly for lightly damped systems near resonance, the natural frequencies split due to coupling. For identical uncoupled resonant frequencies \omega_0 = \frac{1}{\sqrt{LC}}, weak coupling (k \ll 1) results in two approximate modes at \omega \approx \omega_0 \pm \frac{k \omega_0}{2}, manifesting as symmetric splitting around \omega_0 in the response spectrum. This frequency detuning enables selective energy transfer between circuits and is observable as dual peaks in the impedance or admittance curves. These systems find applications in electrical isolation, where mutual coupling allows signal transfer without direct electrical connection, as in transformers, and in wireless power transfer through resonant inductive coupling, enabling efficient mid-range energy delivery (detailed in specialized applications sections). Damping in coupled RLC systems influences both circuits collectively, as resistance in one loop dissipates energy that affects the shared magnetic field, broadening the resonance peaks and reducing the quality factor of the coupled modes. This interdependency means that asymmetric damping can lead to uneven mode decay rates, impacting overall system stability and efficiency.

Historical Development

Early Discoveries

The foundational concept of inductance, essential to RLC circuits, emerged from Michael Faraday's experimental discovery of electromagnetic induction in 1831. Faraday demonstrated that a changing magnetic field could induce an electric current in a nearby conductor, laying the groundwork for understanding how coils store energy in magnetic fields. This breakthrough, achieved through simple apparatus like rotating copper disks between magnets, revealed the inductive properties that would later define the "L" component in circuit theory. Building on empirical observations, James Clerk Maxwell provided a theoretical framework in 1865 with his equations unifying electricity, magnetism, and light propagation. Maxwell's dynamical theory predicted that electromagnetic disturbances could propagate as waves, offering a mathematical basis for oscillatory behaviors in inductive-capacitive systems, though without explicit circuit models at the time. This work inspired subsequent analyses of electrical oscillations. In 1853, William Thomson (later Lord Kelvin) advanced the understanding of LC oscillations by mathematically analyzing the discharge of a Leyden jar through an inductance, demonstrating that the process should exhibit oscillatory charge flow rather than monotonic decay, assuming negligible resistance. Thomson's derivation introduced the concept of inductance explicitly and calculated the frequency of these ideal undamped oscillations, marking a key step toward modeling resonant circuits. Heinrich Hertz's experiments in the late 1880s provided the first practical demonstration of electromagnetic waves using spark-gap devices that functioned as prototype LC circuits. By charging capacitors to produce sparks across gaps in inductive loops, Hertz generated and detected radio waves, confirming Maxwell's predictions through observable resonances at frequencies around 50 MHz. In these setups, resistance inherently caused damping, as the oscillatory sparks visibly decayed over time, highlighting the dissipative role of "R" in real-world inductive-capacitive systems during early wireless telegraphy trials.

Theoretical Advancements

In the late 19th century (around 1893), Oliver Heaviside developed operational calculus as a formal method for solving linear differential equations arising in electrical engineering, particularly for analyzing transients in circuits involving resistors, inductors, and capacitors. This approach treated differentiation and integration as algebraic operations on functions, enabling efficient computation of responses to sudden changes, such as step inputs in RLC configurations, without solving differential equations explicitly. Heaviside's methods, building on experimental foundations from the late 19th century, provided a precursor to Laplace transforms and were instrumental in formalizing transient behavior in lumped-parameter circuits. George Ashley Campbell extended Heaviside's operational calculus to practical electrical problems, applying it to transient phenomena in transmission lines and related RLC networks in the early 20th century. Campbell's work at Bell Laboratories demonstrated how operational methods could model attenuation and distortion in loaded lines, treating inductive and capacitive elements as integral to transient propagation, thus bridging theoretical calculus with circuit design. In the 1910s and early 1920s, Balthasar van der Pol advanced the theory by modeling nonlinear effects in RLC circuits, particularly for self-sustaining oscillations in vacuum-tube amplifiers. His seminal equation described a series RLC circuit with a nonlinear resistor, capturing relaxation oscillations where energy dissipation varies with amplitude, providing a mathematical framework for analyzing limit cycles in nonlinear systems. This contribution shifted focus from linear transients to stability and bifurcation in oscillatory RLC behaviors, influencing subsequent studies in nonlinear dynamics. During the 1920s, Ronald M. Foster formulated the reactance theorem, which characterized the driving-point impedance of lossless LC networks (subsets of RLC systems) as a sum of partial fractions with positive residues at poles on the imaginary axis. This theorem enabled systematic synthesis of reactive networks by ensuring realizability through monotonic reactance increase with frequency, laying groundwork for filter design in RLC configurations. Concurrently, Otto Zobel at Bell Laboratories developed network synthesis techniques for RLC filters, introducing m-derived sections to achieve sharp cutoffs while maintaining constant image impedance. Zobel's methods allowed precise realization of transfer functions using ladder networks of resistors, inductors, and capacitors, optimizing performance for telephone multiplexing applications. In the 1940s, Ernst A. Guillemin pioneered comprehensive synthesis theory for passive RLC networks, emphasizing realizability conditions for arbitrary transfer functions in lumped-element systems. His work, initiated during MIT's Radiation Laboratory efforts, provided analytical procedures to decompose network functions into canonical forms, ensuring stability and passivity through pole-zero configurations. Guillemin's approach unified approximation and realization, enabling the design of broadband filters and equalizers from specified frequency responses. Post-World War II developments introduced precursors to digital simulation for RLC circuits, addressing the limitations of analytical methods for complex networks. Early programs like ECAP (1965) at IBM used numerical integration to solve differential equations for transient and steady-state responses in RLC systems, facilitating parameter sweeps and optimization on vacuum-tube computers. These tools, though rudimentary, marked the transition from hand calculations to computational verification, underemphasized in earlier theoretical literature but essential for validating synthesis techniques.

Practical Applications

Tuned Circuits

Tuned circuits employing RLC configurations play a central role in radio receivers and communication systems by enabling frequency-selective tuning, where the circuit resonates at a specific frequency to amplify desired signals while attenuating others. In these applications, an RLC circuit—typically configured in parallel for high impedance at resonance—achieves maximum response when the inductive reactance equals the capacitive reactance, occurring at the resonant angular frequency \omega_0 = \frac{1}{\sqrt{LC}}. This resonance condition allows the circuit to act as a bandpass filter centered on the desired carrier frequency, essential for isolating broadcast stations in the AM or FM bands. Variable capacitor tuning is a foundational method in analog radio receivers, where the capacitance C is mechanically adjusted to shift \omega_0 and select the target frequency. By rotating a variable capacitor—often a multi-plate air-dielectric design—the user alters the effective capacitance in the RLC tank circuit, enabling precise alignment with incoming radio signals across a wide range, such as 530–1710 kHz for AM broadcasts. This approach provides smooth, continuous tuning and has been standard since early radio designs due to its reliability and low loss at radio frequencies. Gang tuning extends this principle by mechanically linking multiple variable capacitors on a common shaft, synchronizing the resonant frequencies across several RLC stages in the receiver, such as the RF amplifier and local oscillator sections. This simultaneous adjustment ensures consistent tracking between the antenna input circuit and the mixer stage, simplifying operation and maintaining image frequency rejection without complex individual controls. In multi-stage receivers, gang tuning typically involves two or more capacitor gangs, enhancing overall circuit alignment for better signal capture. In superheterodyne receivers, which dominate modern radio architecture, the intermediate frequency (IF) stage features a fixed RLC tuned circuit optimized for a constant IF, usually 455 kHz for AM or 10.7 MHz for FM. After the incoming RF signal is mixed with a local oscillator to produce this fixed IF, the RLC circuit in the IF amplifier provides sharp selectivity and gain at that frequency, independent of the tuned RF input. This fixed tuning simplifies design, allowing high-Q transformers or ceramic resonators in the IF stages for consistent performance across the receiver's band. The quality factor Q of an RLC tuned circuit critically influences its performance, defined as Q = \frac{\omega_0 L}{R} = \frac{1}{\omega_0 R C} for series or parallel equivalents, quantifying the ratio of stored to dissipated energy. A higher Q yields narrower bandwidth \Delta \omega = \frac{\omega_0}{Q}, enhancing selectivity by rejecting adjacent-channel interference, but it introduces a trade-off with sensitivity—the circuit's ability to detect weak signals—as excessively high Q can overly restrict the passband and complicate precise tuning. Optimal Q values, often 50–200 in radio IF stages, balance these factors for practical reception quality. While traditional mechanical tuning dominates analog discussions, modern implementations increasingly incorporate varactor diodes—semiconductor devices whose junction capacitance varies with applied reverse bias voltage—for electronic tuning in RLC circuits. These varactors replace or augment mechanical capacitors in portable radios and integrated receivers, enabling voltage-controlled \omega_0 adjustment with tuning ratios exceeding 10:1 and Q > 100, though they introduce some nonlinearity that requires compensation circuits. This shift supports compact, automated tuning in digital communication systems while preserving the core RLC resonance principles. Metal detectors employ RLC resonant circuits to identify metallic objects by exploiting changes in resonance caused by nearby conductors. In very low frequency (VLF) metal detectors, a transmitter coil connected to an LC circuit generates an alternating magnetic field at the resonant frequency. When a metal object enters the field, it induces eddy currents that alter the effective inductance or mutual coupling, detuning the circuit and producing a detectable phase shift or amplitude variation in the receiver coil's induced voltage. This perturbation, often processed through a balanced RLC configuration, triggers an audio or visual alert, enabling applications in security screening, archaeology, and treasure hunting.

Filter Designs

Passive RLC filters form the cornerstone of passive signal processing circuits, employing resistors (R), inductors (L), and capacitors (C) to selectively attenuate frequencies outside a desired passband through resonance and impedance variation. Unlike active filters, these designs rely solely on passive elements, making them suitable for high-frequency applications where amplification is unnecessary or undesirable, such as in RF front-ends and power supplies. The resonance in LC components creates sharp transitions at cutoff frequencies, enabling effective broadband filtering without external power. The order of an RLC filter corresponds to the number of energy-storage elements (L and C pairs), with a single RLC stage yielding a second-order response characterized by a quadratic transfer function. Higher-order filters are constructed by cascading multiple second-order sections or incorporating additional reactive elements, which steepen the frequency response and improve selectivity; for instance, a fourth-order filter doubles the asymptotic roll-off rate compared to second-order. In the stopband, a second-order RLC filter exhibits an attenuation slope of 12 dB per octave (or 40 dB per decade), providing moderate rejection that scales with order for more demanding applications. The transfer function H(s) = \frac{V_\text{out}(s)}{V_\text{in}(s)} for RLC filters is obtained by analyzing the voltage division across impedances in the Laplace domain, where Z_R = R, Z_L = sL, and Z_C = 1/(sC). For a prototypical series RLC low-pass configuration—with input across R in series with L, and output across a shunt C—the transfer function simplifies to H(s) = \frac{1}{s^2 LC + s RC + 1}, with the natural frequency \omega_0 = 1/\sqrt{LC} setting the cutoff and damping factor \zeta = R/2 \sqrt{C/L} influencing the quality factor Q and potential peaking. Attenuation is quantified in decibels as $20 \log_{10} |H(j\omega)|, revealing passband flatness and stopband rejection versus frequency. Impedance matching plays a critical role in RLC filter performance, ensuring maximum power transfer from source to load while minimizing signal reflections and losses, especially in systems with mismatched characteristic impedances like 50 Ω RF lines. Design principles involve selecting component values to equate the filter's input impedance to the source conjugate and output to the load, often using L-sections or pi-networks augmented with RLC elements; mismatches can degrade insertion loss by several dB and alter the frequency response. For example, in a low-pass RLC filter interfacing a 50 Ω source and load, iterative adjustment of L and C values achieves near-optimal matching across the passband.

Oscillator Circuits

RLC circuits form the basis of feedback oscillators, where an amplifier provides positive feedback to a resonant tank circuit, sustaining oscillations at the resonant angular frequency \omega_0. The principle relies on injecting a portion of the output signal back into the input with the correct phase and amplitude to reinforce the signal, requiring an initial loop gain greater than unity to overcome losses and build up the oscillation amplitude. For sustained oscillations, the Barkhausen criterion must be satisfied: the magnitude of the loop gain must equal 1, and the total phase shift around the feedback loop must be an integer multiple of 360° (or 0°), ensuring the fed-back signal is in phase with the input at \omega_0. High-quality factor (Q) components are essential in the RLC tank to minimize damping and achieve stable sinusoidal output with low distortion. Common configurations include the Colpitts oscillator, which employs a parallel RLC tank with a tapped capacitor divider for feedback; the two capacitors in series act as a voltage divider across the inductor, providing the necessary phase shift for positive feedback. In contrast, the Hartley oscillator uses a series RLC arrangement with a tapped inductor; the inductor is split into two sections, allowing feedback from the tap point to the amplifier input while the total inductance resonates with the capacitor. Oscillations initiate from thermal noise or other perturbations in the circuit, which are selectively amplified at the resonant frequency \omega_0 due to the high gain at resonance, gradually building to a steady-state amplitude limited by nonlinear effects in the amplifier. To enhance frequency stability in practical RLC oscillators, especially against temperature variations and aging, crystal control is often integrated; a quartz crystal replaces or augments the LC tank, exploiting the piezoelectric effect for a much higher effective Q and stability on the order of parts per million.

Specialized Uses

RLC circuits find specialized applications in voltage multiplication for high-voltage generation, where the resonant behavior amplifies voltage across components. In series RLC configurations driven near resonance, the voltage across the inductor or capacitor can exceed the input supply significantly, enabling compact high-voltage sources without transformers. This principle is employed in automotive ignition systems, where a series RLC circuit, triggered by a spark coil, generates pulses up to several kilovolts from a 12 V battery to ignite fuel in engines. In pulse discharge applications, RLC circuits shape high-energy pulses for precise timing in radar and medical systems. Pulse forming networks (PFNs), often comprising cascaded LC sections with inherent or added resistance forming RLC elements, store and release energy to produce flat-topped pulses with controlled duration. The characteristic discharge time in such underdamped RLC networks is approximately \sqrt{LC}, determining pulse width for applications like radar transmitters, where pulses must match antenna bandwidths for optimal range resolution. In medicine, similar RLC-based PFNs drive pulsed lasers for tissue ablation or generate defibrillation shocks, ensuring safe energy delivery with minimal distortion. Snubber circuits incorporating RLC elements suppress voltage and current transients in power electronics, protecting switches from overvoltages during commutation. Unlike simple RC snubbers, RLC configurations provide tuned damping to eliminate ringing at specific frequencies, absorbing inductive kickback energy through resonant oscillation that decays rapidly. For instance, in silicon carbide (SiC) MOSFET-based converters, an inductively coupled RLC snubber reduces switching losses by over 20% while clamping peak voltages below device ratings, enhancing reliability in high-frequency DC-DC converters. In sensor damping, RLC shunt circuits integrated with piezoelectric transducers control vibrations, improving measurement accuracy in dynamic environments. The piezoelectric element converts mechanical strain to electrical charge, which is shunted through a series RLC network tuned to the sensor's resonant frequency; this creates a virtual mass-spring-damper analogy, dissipating vibrational energy as heat in the resistor. Applications include aerospace sensors on vibrating structures, where RLC shunts achieve up to 40 dB attenuation at target modes, and industrial accelerometers requiring stable response amid mechanical noise. Post-2000 advancements in microelectromechanical systems (MEMS) leverage RLC equivalent circuit models for compact sensors, enabling precise detection of physical parameters through resonant frequency shifts. MEMS resonators, such as capacitive or piezoelectric designs, are electrically modeled as series or parallel RLC circuits, where motional resistance R, inductance L, and capacitance C represent mechanical damping, mass, and compliance, respectively. This modeling facilitates integration into RF filters and inertial sensors; for example, in gyroscopes, electromechanical amplification via RLC tuning boosts sensitivity by factors of 10-100, supporting applications in navigation and biomedical implants since the early 2000s. Recent experimental advances as of 2025 have introduced topolectrical circuits, which use networks of interconnected RLC components to simulate topological phases of matter and other complex quantum phenomena in classical electrical systems. These circuits enable the study of non-Hermitian dynamics, Floquet engineering, and nonlinear effects, offering practical platforms for physics research and potential applications in analog quantum simulation and advanced signal processing.

References

  1. [1]
    What is a RLC Circuit? - Keysight Oscilloscope Glossary
    A RLC circuit (Resistor-Inductor-Capacitor Circuit), also known as a tuned or resonant circuit, is an electric circuit composed of resistors (R), inductors (L), ...Missing: authoritative | Show results with:authoritative
  2. [2]
    What are RLC Circuits? - Ansys
    RLC circuits are electrical circuits in which resistors, inductors, and capacitors are connected either in series or in parallel.Missing: authoritative | Show results with:authoritative
  3. [3]
    Series RLC Circuit Analysis - Electronics Tutorials
    Series RLC circuits are classed as second-order circuits because they contain two energy storage elements, an inductance L and a capacitance C. Consider the RLC ...Missing: authoritative | Show results with:authoritative
  4. [4]
    [PDF] Chapter 21: RLC Circuits
    element, a resistor, a capacitor, or an inductor. The current amplitude is then measured as a function of frequency. Which one of the following curves ...
  5. [5]
    [PDF] Chapter 12 Alternating-Current Circuits - MIT
    1 Parallel RLC circuit. Unlike the series RLC circuit, the instantaneous voltages across all three circuit elements. R, L, and C are the same, and each ...
  6. [6]
    6.3 The RLC Circuit - Ximera - The Ohio State University
    As we'll see, the circuit is an electrical analog of a spring-mass system with damping. Nothing happens while the switch is open (dashed line).
  7. [7]
    14.6 RLC Series Circuits – University Physics Volume 2
    Figure 14.17 (a) An RLC circuit. Electromagnetic oscillations begin when the switch is closed. The capacitor is fully charged initially. (b) Damped oscillations ...
  8. [8]
    [PDF] Circuits You'll Build Skills and Concepts You'll Learn 1 RLC resonators
    They are at the heart of wireless radio transmissions. Resonators are also at the heart of many chemical and biological engineering applications, such as.
  9. [9]
    [PDF] Chapter 3: Electromagnetic Fields in Simple Devices and Circuits
    May 3, 2011 · RLC resonators typically consist of a resistor R, inductor L, and capacitor C connected in series or parallel, as illustrated in Figure 3.5.1.
  10. [10]
    Energy Dissipation/Storage in R, C, and L
    Energy dissipated by resistor $R$. When a voltage $v(t)$ is applied across $R$ , the current through it is · Energy storage in capacitor $C$. Given voltage $v(t) ...Missing: RLC | Show results with:RLC
  11. [11]
    [PDF] 6.200 Notes: Time Domain R-L-C
    Apr 11, 2023 · The typical LRC circuit consists of a resistor, capacitor, and induc- tor either in parallel or in a series loop configuration. These two cases ...
  12. [12]
    [PDF] 18.03 Differential Equations, Supplementary Notes Ch. 8
    The Mathlet Series RLC Circuit exhibits the behavior of this sys tem, when the voltage source provides a sinusoidal signal. Current flows through the circuit; ...
  13. [13]
    4: First and Second Order Circuits - UCF ECE
    A characteristic equation, which is derived from the governing differential equation, is often used to determine the natural response of the circuit. The ...
  14. [14]
    Parallel RLC Circuit Analysis - Electronics Tutorials
    Taking the derivative, dividing through the above equation by C and then re-arranging gives us the following Second-order equation for the circuit current. It ...
  15. [15]
    [PDF] We shall examine three special cases of driven circuits
    the resistor, (b) the inductor, and (c) the capacitor, of a series RLC circuit. ... As in the series RLC circuit, power is dissipated only through the resistor.
  16. [16]
    [PDF] RESONANCE IN SERIES AND PARALLEL RLC NETWORKS
    There are three methods for defining parallel resonance, each resulting in a different resonant frequency. For the parallel resonant circuit the three different ...
  17. [17]
    [PDF] 8B RLC Circuits: Forced Oscillations and Resonance
    Resonance in RLC circuits occurs when the driving frequency matches the natural frequency, minimizing impedance. The natural frequency is 2πf0, and the current ...
  18. [18]
    [PDF] AC RLC Circuit
    ... circuit has a maximum current. This is the condition of resonance. From the above formulas, it can be seen that resonance occurs for ω = ωR (f = fR) where.<|control11|><|separator|>
  19. [19]
    [PDF] Chapter 31: RLC Circuits
    →Conservation of energy: →Energy oscillates between capacitor and inductor. ◇Endless oscillation between electrical and magnetic energy. ◇Just like ...Missing: storage | Show results with:storage
  20. [20]
    [PDF] Transient Response Series RLC circuit
    The equation that describes the response of this circuit is. 2. 2. 1. 0 d vc vc dt. LC. +. = (1.16). Assuming a solution of the form st. Ae the characteristic ...
  21. [21]
    23.12 RLC Series AC Circuits – College Physics - UCF Pressbooks
    The figure describes an R LC series circuit. It shows a resistor R connected in series with an inductor L, connected to a capacitor C in series to an AC source ...
  22. [22]
    [PDF] Experiment 8: Undriven & Driven RLC Circuits
    What is the period of the oscillations (measure the time between distant zeroes of the current and divide by the number of periods between those zeroes)? What ...
  23. [23]
    [PDF] AC Electrical Circuit Analysis - Mohawk Valley Community College
    Apr 22, 2021 · The goal of this text is to introduce the theory and practical application of analysis of AC electrical circuits. It assumes familiarity with DC.Missing: history | Show results with:history
  24. [24]
    [PDF] 8A RLC Circuits: Free Oscillations
    For under- and critically damped oscillations τ depends on R and L, but for overdamped oscillations we expect τ = RC. Fig. 3. Circuit used throughout this ...
  25. [25]
    RLC natural response - intuition - Spinning Numbers
    The addition of the resistor to the RLC is equivalent to adding friction. Friction causes the pendulum to dissipate energy and slowly come to a halt.
  26. [26]
    [PDF] WPI Physics Dept. Intermediate Lab 2651 Free and Damped ...
    Mar 11, 2017 · For under-damped oscillations: a) determine the oscillation period and compare the measured oscillation frequency with the theoretical value;.
  27. [27]
    8. Damping and the Natural Response in RLC Circuits
    We learn in this section about damping in a circuit with a resistor, inductor and capacitor, using differential equations.
  28. [28]
    [PDF] Frequency response: Resonance, Bandwidth, Q factor
    The current flowing in the system is in phase with the source voltage. The power dissipated in the RLC circuit is equal to the power dissipated by the resistor.
  29. [29]
    [PDF] Lecture 4: RLC Circuits and Resonant Circuits
    Depending on L, C, and ω, the phase angle can be positive or negative! In this example, if ωL > 1/ωC, then V. R. (t) lags V in. (t). Page 3. K.K. Gan.
  30. [30]
    None
    ### Summary of Second-Order Filters (ENGR 202 Section 3)
  31. [31]
    Resonant RLC Circuits - HyperPhysics
    For series resonance the condition of resonance is straightforward and it is characterized by minimum impedance and zero phase. Parallel resonance , which ...
  32. [32]
    [PDF] Lecture 4: RLC Circuits and Resonant Circuits
    At resonance (series, parallel etc):. ◇. At the resonant frequency ... Q (quality factor) of a circuit: determines how well the RLC circuit stores energy.
  33. [33]
    [PDF] resonators and Q - UCSB ECE
    It is easy to show that for a parallel resonant circuit, the Q is given by susceptance/conductance: G. B. Q = where B is the susceptance of the capacitor or ...
  34. [34]
    [PDF] Supplemental Notes on Complex Numbers, Complex Impedance ...
    Complex Impedance, RLC Circuits, and Resonance. Complex numbers. Complex ... For the RLC circuit in series, we obtained above Z = R+iωL−i/(ωC), and ...
  35. [35]
    (PDF) Transient Analysis of R-L-C Series Circuit to Step voltage by Engineering Methods
    ### Summary of Transient Analysis of R-L-C Series Circuit to Step Voltage
  36. [36]
    [PDF] Transients and Oscillations in RLC Circuits
    With damping, the frequency of oscillation is shifted to a smaller value. The logarithmic decrement, δ , is the natural log of the ratio of the charge (see ...
  37. [37]
    [PDF] 1.6 Second-Order Transients
    Let's use some component values in the RLC circuit and see what happens. Overdamped Example t = 0. R .90Ω. R. L. 2.
  38. [38]
    Analogous Electrical and Mechanical Systems - Swarthmore College
    To apply this analogy, every node in the electrical circuit becomes a point in the mechanical system. Ground becomes a fixed location, resistor become friction ...Background · Converting Between Systems · Electrical to Mechanical 1...
  39. [39]
    The "Q" factor of an oscillating system
    How many oscillations to lose amplitude or energy? · after Q/π cycles, amplitude will be (1/e) of its original value · after Q/(2 π) cycles, energy will be (1/e) ...
  40. [40]
    RLC natural response - variations - Spinning Numbers
    We begin again at Step 5 of the strategy: Find the roots of the characteristic equation. α = ω o \alpha =\omega_o α=ωo​ is called critically damped, the ...Strategy · Three variations · Under damped · Euler's formulaMissing: ratio | Show results with:ratio
  41. [41]
    RLC natural response - variations (article) - Khan Academy
    The RLC natural response falls into three categories: overdamped, critically damped, and underdamped. Written by Willy McAllister.
  42. [42]
    [PDF] The Effect of Switch Resistance on Pipelined ADC MDAC Settling ...
    Though a more precise settling optimization exists [4], a near optimal goal is critical damping for minimum settling time without overshoot. This occurs ...
  43. [43]
    [PDF] and Second-Order System Response1 1 First-Order Linear ... - MIT
    The maximum velocity in the critically damped case is less than 45% of the maximum velocity when the damping ... critical damping may be found by setting ζ ...
  44. [44]
    [DOC] CS-474
    Critical damping should be thought of as an “optimal” solution that differentiates between over- and under-damping. Figure 6 depicts this result. Figure 6: An ...
  45. [45]
    Second-Order Circuits - Study Guides - CircuitBread
    Step Response of a Parallel RLC Circuit · Substituting v = L di/dt and dividing by LC, · Again, the complete solution consists of the transient response (it(t)) ...
  46. [46]
    Transient Response of Second-Order Circuits - ECE 3310
    The most basic second order circuit is the one that consists of one inductor and one capacitor that are connected in parallel (or series), as shown in Figure 1b ...Missing: configurations | Show results with:configurations
  47. [47]
    [PDF] NETWORK ANALYSIS & SYNTHESIS
    The simple cascaded networks are in the form of a ―ladder topology‖, which can be RC, RL, or. RLC constructs. Z2. Z3. Z6. ZN. Impedance ladder topology. The ...<|separator|>
  48. [48]
    [PDF] HIGHER-ORDER LADDER PASSIVE AND ACTIVE FILTERS In Fig ...
    Ladder filters use a ladder configuration, starting with an RLC passive circuit, and their performance depends on multiple components, not just one.
  49. [49]
  50. [50]
    [PDF] Quality Factor, Bandwidth, and Harmonic Attenuation of Pi Networks
    Figure 1 shows a pi-network circuit configuration often used to match a load impedance, RL, to a source impedance, RS. A common application of pi networks is in ...
  51. [51]
    [PDF] Improving the Bandwidth of Simple Matching Networks
    Impedance matching ... When the intermediate impedance is lower than the source and load, the result of combining the center components is the Pi-network.Missing: RLC | Show results with:RLC
  52. [52]
    [PDF] Lecture 08: Impedance Matching 2 - Harvey Mudd College
    We start by splitting our pi network into two L-matches as in the last video, but we make a surprising decision with the left matching network. Instead of ...
  53. [53]
    [PDF] Matching Networks, Q, Smith Charts Matching networks and Filters ...
    ➤ Consider the T or Pi network to be a pair of back to back L networks. ➤ The virtual resistance in a Pi network must be smaller that those on the source ...Missing: RLC | Show results with:RLC
  54. [54]
    [PDF] Physics 24100 – Electricity & Optics - Purdue Physics department
    A Circuit with Mutual Inductance. V − i1. R1. − L1. di1. dt. + M. di2. dt. = 0. −i2. R2. − L2. di2. dt. + M. di1. dt. = 0. −. +.
  55. [55]
    [PDF] Coupled and Characteristic Modes of a Wideband Slot Antenna ...
    With K ~ 𝜅𝜔0/2 [2] and 𝜅 = 𝜅e + 𝜅m= 0.45, (1) predicts coupled resonant frequencies of 670 MHz and 1061 MHz, within 1% of the CMA-calculated resonant ...
  56. [56]
    Wireless Power Transfer via Strongly Coupled Magnetic Resonances
    Using self-resonant coils in a strongly coupled regime, we experimentally demonstrated efficient nonradiative power transfer over distances up to 8 times ...
  57. [57]
    September 4, 1821 and August 29, 1831: Faraday and ...
    Aug 1, 2001 · Faraday then proceeded to demonstrate that the lines of magnetic force could be cut, and a current induced, simply by rotating a copper disc by ...Missing: primary source
  58. [58]
    VIII. A dynamical theory of the electromagnetic field - Journals
    A dynamical theory of the electromagnetic field. James Clerk Maxwell ... A (2025) The Foundation of Electromagnetism: A Comprehensive Study of Maxwell's Equations ...
  59. [59]
    WILLIAM THOMSON, LORD KELVIN (1824–1907) - ScienceDirect
    Kelvin's paper of 1853 gave the full modern theory of this effect. From his definition of inductance L, he showed that the charge Q on the condenser must ...
  60. [60]
    [PDF] Mechanical resonance: 300 years from discovery to the full ... - arXiv
    Nov 18, 2018 · In 1853 William Thomson, later famously known as Lord Kelvin (1824–1907), one of the leading physicists of the entire 19th century,. Page 25 ...
  61. [61]
    Revisiting the 1888 Hertz experiment | American Journal of Physics
    We revisit the original experiment performed by Hertz in 1888 and use a simple setup to produce electromagnetic waves with a frequency in the range of 3MHz.
  62. [62]
    Van der Pol and the history of relaxation oscillations - AIP Publishing
    Apr 30, 2012 · Browsing the literature from the late 19th century, we identified four self-oscillating systems in which relaxation oscillations have been ...
  63. [63]
    [PDF] The Heaviside Operational Calculus - vtda.org
    formula. The present paper, by attacking the problem from a different standpoint, shows that the Heaviside operational formula is a shorthand equivalent of.<|separator|>
  64. [64]
  65. [65]
    [PDF] Van der Pol and the history of relaxation oscillations - arXiv
    Aug 21, 2014 · Our aim is to investigate how re- laxation oscillations were actually discovered. Browsing the literature from the late 19th century, we ...
  66. [66]
    [PDF] The theorem gives the most general form of the driving-point - vtda.org
    A Reactance Theorem. By RONALD M. FOSTER. SYNOPSIS: The theorem gives the most general form of the driving-point impedance of any network composed of a finite ...
  67. [67]
    A reactance theorem - Nokia Bell Labs - IEEE Xplore
    The theorem gives the most general form of the driving-point impedance of any network composed of a finite number of self-inductances, mutual inductances, ...
  68. [68]
    m-derived filter - Wikipedia
    They were invented by Otto Zobel in the early 1920s. This filter type was originally intended for use with telephone multiplexing and was an improvement ...
  69. [69]
    Synthesis of RC‐Networks - Guillemin - 1949 - Wiley Online Library
    This study was begun in the Spring of 1944, being suggested by certain phases of the M. I. T. Radiation Laboratory program. Although the theoretical ...Missing: RLC 1940s
  70. [70]
  71. [71]
    [PDF] A Historical Review of Circuit Simulation
    The main focus of this paper is the history of the computer-aided analysis of electronic circuits. The principal sections of this paper constitute the author's ...
  72. [72]
    The tuning circuit in an FM radio receiver is a series RLC circui...
    It is given by the formula f₀ = 1 / (2π√(LC)), where f₀ is the resonant frequency, L is the inductance, and C is the capacitance. At this frequency, the circuit ...
  73. [73]
    HOW DO TUNING CAPACITORS WORK?
    A variable capacitor used in radio tuning circuits works by changing the capacitance in an LC (inductor-capacitor) circuit to set the resonant frequency, ...Missing: RLC | Show results with:RLC
  74. [74]
    Ganged Tuning definition - Farnell® UK
    Ganged Tuning. AM radios have a tunable LC circuit at their front end, whose resonant frequency can be adjusted to tune into different radio stations.Missing: RLC | Show results with:RLC
  75. [75]
    Superheterodyne Radio IF Amplifier & Filter - Electronics Notes
    Very basic superhet radio receiver circuit with IF stages marked in yellow The circuit of the simple superheterodyne radio shows how this is achieved.Missing: RLC | Show results with:RLC
  76. [76]
    Superheterodyne AM Receiver - Working with Block Diagram and ...
    Mar 12, 2020 · First IF Filter: The fourth block is the first IF filter. In most AM receivers, it is a resonant circuit placed in the collector of the mixer ...
  77. [77]
    Superheterodyne Receiver - Radartutorial.eu
    The superheterodyne receiver changes the rf frequency into an easier to process lower IF- frequency. This IF- frequency will be amplified and demodulated to ...
  78. [78]
    Q Factor and Bandwidth of a Resonant Circuit | Electronics Textbook
    The Q factor measures a circuit's 'goodness'. Bandwidth is calculated as fc/Q, where fc is the resonant frequency. A higher Q means a narrower bandwidth.
  79. [79]
    Quality factor and bandwidth | Electrical Circuits and Systems II ...
    Quality factor (Q) measures the efficiency of energy storage in a resonant circuit · Higher Q values indicate lower energy losses and sharper resonance peaks ...
  80. [80]
    The tuning circuit in an FM radio receiver is a series RLC circui...
    Nov 11, 2023 · The tuning circuit is a series RLC circuit, and the goal is to determine the maximum resistance such that the peak current at 103.9 MHz is no more than 0.10% ...
  81. [81]
    Using Varactor Diodes for FM Signal Generation - Technical Articles
    Aug 31, 2025 · Learn how a varactor diode's variable capacitance, together with an LC tank circuit, can drive a voltage-controlled oscillator (VCO) to create ...
  82. [82]
    Varactor - an overview | ScienceDirect Topics
    A varactor (which is a shortened form of variable reactor) is simply a pn-junction diode that has been engineered to maximize the value and range of junction ...Missing: RLC | Show results with:RLC
  83. [83]
    [PDF] PREFERRED-CIRCUIT TECHNIQUES FOR VARACTOR-TUNED ...
    Jul 9, 1975 · This report describes an investigation of circuit techniques to minimize spurious reponses in varactor-tuned filters. Desi&n equations for the ...
  84. [84]
    [PDF] CHAPTER 8 ANALOG FILTERS
    This is especially true if the single-pole is implemented as a passive filter. For the passive design, we will choose the zero input impedance configuration.
  85. [85]
    Second Order Filters and the 2nd-order low pass filter circuit
    Second Order Filters tutorial and how 2nd Order Active Low Pass Filters have a roll-off rate of -40 dB/decade in the stop band.
  86. [86]
    [PDF] RLC Filter Design for ADC Interface Applications - Texas Instruments
    Consider a typical filter design target and apply the design equations developed here to implement a 2nd- order passive RLC filter, using the circuit of Figure ...
  87. [87]
    Impedance Matching Filter Circuit Design – LC, L and PI Filters
    Apr 24, 2020 · There are many types of filter circuits that can be used in for Impedance matching, the most common ones are discussed in this article.Missing: passive | Show results with:passive
  88. [88]
    [PDF] Oscillator Circuits - Oregon State University
    For self-sustaining oscillations: • the feedback signal must positive. • the overall gain must be equal to one (unity gain). Page 3. 3. If the feedback signal ...Missing: principle | Show results with:principle
  89. [89]
    Barkhausen's Stability Criteria for Stability in Oscillators
    Jan 5, 2024 · This condition is ideal for sustained oscillations in an oscillator circuit and it marks the borderline between stable/unstable behavior.
  90. [90]
    Colpitts Oscillators
    Here are the requirements for these circuits to oscillate: an LC tank tuning circuit that generates sinusoidal oscillation at its resonant frequency.Missing: RLC | Show results with:RLC
  91. [91]
    Colpitts Oscillator Tutorial and Colpitts Design - Electronics Tutorials
    The Colpitts Oscillator design uses two centre-tapped capacitors in series with a parallel inductor to form its resonance tank circuit producing sinusoidal ...Missing: RLC | Show results with:RLC
  92. [92]
    Hartley Oscillator - Electronics Tutorials
    The Colpitts oscillator uses two capacitors in series to form a centre tapped capacitance in parallel with a single inductance within its resonant tank circuit.Missing: RLC | Show results with:RLC
  93. [93]
    [PDF] Analyzing the Behavior of an Oscillator and Ensuring Good Start-up
    Noise: Thanks to the noise an oscillator is able to startup. This noise has different origins: – thermal noise due to the transistor junctions and resistors, – ...
  94. [94]
    [PDF] Microcontroller Oscillator Circuit Design Considerations
    The physical properties of quartz make it very stable over both time and temperature. The usual model of a crystal is a network of two capacitors, an inductor ...<|control11|><|separator|>
  95. [95]
    Is the voltage in a series RLC circuit always greater than zero? - Quora
    Jul 8, 2023 · Besides above applications , RLC circuits are used in automobile ignition system to generate very high voltage with a mere 12V. Inductor is ...
  96. [96]
    Series RLC Resonant Circuit Used as Frequency Multiplier - MDPI
    Dec 9, 2022 · Due to the higher increase of the input supply voltage by increasing the multiplier “n”, there are increases in the instantaneous input ...
  97. [97]
    [PDF] Pulsed power - People@UTM
    • In lieu of the capacitor and inductor in the RLC circuit, a transmission line is employed as a pulse forming network. • The transmission line is initially ...<|separator|>
  98. [98]
    Pulse Forming Networks - General Atomics
    PFNs are usually comprised of a number of capacitors and inductors arranged so that the discharge pulses from the capacitors are spaced in time.Missing: RLC | Show results with:RLC
  99. [99]
    (PDF) Compact low‐cost high‐voltage pulse generator for biological ...
    May 24, 2025 · A new architecture of a pulse generator is proposed. It provides a quasi‐rectangular output with nanosecond rise time and fully controlled repetition rate.
  100. [100]
    A novel inductively coupled RLC damping scheme for eliminating ...
    Mar 25, 2023 · Without the RLC snubber, the measured power loss of the whole system is equal to 1.42 mJ. The power loss of the total system with the RLC is ...
  101. [101]
    Switching waveforms with RL snubber and RLC damping.
    Nevertheless, the switching oscillations of SiC MOSFETs could be eliminated with an RLC snubber [44] . However, the examination of the effectiveness of the ...
  102. [102]
    Shunt Damping Vibration Control Technology: A Review - MDPI
    Shunt damping utilizes an external circuit across the terminals of smart structure based transducers to realize vibration control.
  103. [103]
    (PDF) Shunt Damping Vibration Control Technology: A Review
    May 11, 2017 · RLC shunted piezoelectric transducers for increasing the transmission loss of a plate,. reproduced with permission from [128], Academic Press ...
  104. [104]
    [PDF] Microelectromechanical Resonators for Radio Frequency ... - arXiv
    The most common transduction type used in MEMS resonators is electrostatic actuation and sensing, as it can produce small components that are robust, relatively ...
  105. [105]
    An analytical design methodology for microelectromechanical (MEM ...
    This paper starts with a complete treatment of series RLC equivalent circuit in terms of derivation of frequency response in Section 2 followed by similar ...
  106. [106]
    Resonance
    Boston University physics notes on resonance in LRC circuits, including applications to metal detectors.
  107. [107]
    Series Resonance in a Series RLC Circuit
    Electronics Tutorials explanation of series resonance and its use in devices like metal detectors.
  108. [108]
    Resonance in LRC Circuits
    LibreTexts physics resource on LRC resonance, with examples including metal detection principles.