Fact-checked by Grok 2 weeks ago

Loop gain

In control systems and electronics, loop gain refers to the product of the forward-path gain (often denoted as A or G(s)) and the feedback-path gain (denoted as \beta or H(s)) in a closed-loop feedback configuration, representing the total amplification encountered by a signal as it travels around the feedback loop. This measure, typically expressed as T(s) = A(s)\beta(s) in the frequency domain, quantifies the strength of the feedback mechanism and is fundamental to understanding system behavior. Loop gain plays a critical role in determining the performance and of feedback systems, such as amplifiers and automatic control loops, by influencing key metrics like closed-loop accuracy, , and . Higher loop gain enhances output accuracy by minimizing errors due to variations in input or component tolerances—for instance, in operational amplifiers, it ensures the closed-loop gain closely approximates the $1/\beta under the assumption of high . Conversely, excessive loop gain at certain frequencies can lead to , where the system oscillates if the loop gain magnitude reaches unity (0 dB) while the phase shift approaches -180 degrees. To assess stability, engineers analyze loop gain through frequency-response tools like Bode plots, which plot the magnitude and phase of T(j\omega) versus frequency \omega. These plots reveal gain margin (the factor by which loop gain can increase before ) and phase margin (the additional phase lag tolerable before ), with typical design targets of at least 10 dB gain margin and 45–60 degrees for robust operation. Loop gain is measured practically by techniques such as voltage injection, where a test signal is introduced into the loop to compute the ratio of returned to injected voltages, accounting for real-world parasitics. In broader applications, from analog regulators to systems, optimizing loop gain balances trade-offs between speed, precision, and reliability, underpinning designs in fields like and .

Fundamentals

Definition

Loop gain refers to the product of the gains, or more generally the functions, of all components comprising a closed loop in a dynamic , quantifying the or experienced by a signal after traversing the entire loop once. This measure captures the cumulative effect of the forward and paths, playing a pivotal role in determining the overall performance and characteristics of systems. In the context of negative feedback, a common convention incorporates a negative sign into the loop gain definition to reflect the subtractive mixing at the input summing junction, expressed as T = -A\beta, where A is the forward path gain and \beta is the feedback factor. This notation accounts for the phase inversion in the feedback signal, ensuring that for stable operation, the magnitude |A\beta| > 0 while the overall feedback opposes changes in the output; the resulting closed-loop gain approximates $1/\beta when |T| \gg 1. Unlike , which describes the amplification provided solely by the forward path (e.g., A or G(s)) without considering , loop gain explicitly includes the contributions from the path but excludes the dynamics of the summing itself, focusing on the signal's circulation around the . A typical representation appears in the standard of a , featuring an input signal entering a summing where it is subtracted by the signal from the output, followed by a forward path with G(s) leading to the output, and a path with H(s) returning to the ; here, the loop gain is defined as L(s) = G(s) H(s).

Basic Principles

One of the primary benefits of high loop gain in systems is gain desensitization, which minimizes the closed-loop gain's dependence on variations in the open-loop gain due to factors like or tolerances. When the loop gain magnitude |T| is much greater than 1, the closed-loop gain approximates 1/β, where β is the factor, making the overall response primarily determined by stable passive components rather than the potentially variable active elements. Loop gain also enables bandwidth extension by trading off some low-frequency gain for a wider frequency response, as the feedback loop compensates for the amplifier's inherent roll-off, allowing the closed-loop system to maintain flat gain over a broader range. This effect arises because the loop gain provides additional gain at higher frequencies, effectively extending the usable bandwidth until the point where |T| approaches unity. In addition, high loop gain suppresses distortions from nonlinearities and reduces the impact of external by a factor of approximately (1 + T), as the feedback mechanism corrects errors introduced in the forward path. This attenuation applies to both harmonic distortion generated within the and additive sources, improving overall signal . Feedback influenced by loop gain further modifies impedances, such as increasing the in non-inverting configurations by roughly (1 + T) to better isolate the source, while decreasing the to enhance load driving capability. These alterations make the system less sensitive to loading effects. For these principles to hold effectively, the loop gain magnitude |T| must be much greater than 1 at the frequencies of interest, ensuring the dominates and the approximations remain valid.

Mathematical Representation

Time-Domain Formulation

In the time domain, loop gain characterizes the propagation of an signal through the loop and back to the summing , representing the system's response after one complete traversal. For linear time-invariant systems, this is expressed as the integral f(t) = \int_{0}^{t} h_T(\tau) e(t - \tau) \, d\tau, where e(t) is the signal, h_T(t) is the of the open-loop transfer function around the loop, and f(t) is the returned signal. This arises from the superposition of responses of all components in the loop, such as amplifiers, sensors, and actuators, convolved in series. For linear time-invariant systems, the time-domain loop gain manifests in differential equations that describe the closed-loop dynamics, incorporating the loop traversal to relate input to output. Consider a basic first-order system with plant dynamics \dot{y}(t) = - \frac{1}{\tau} y(t) + \frac{K}{\tau} u(t) and unity feedback u(t) = r(t) - y(t), where K is the forward gain. Substituting yields the closed-loop equation \dot{y}(t) + \frac{1 + K}{\tau} y(t) = \frac{K}{\tau} r(t), with the loop gain influencing the effective damping term (1 + K)/\tau. Here, higher loop gain K accelerates the transient response by reducing the time constant to \tau / (1 + K). A representative example is the of this loop to a unit step input r(t) = 1 for t \geq 0. The is y(t) = 1 - e^{-(1 + [K](/page/K)) t / [\tau](/page/Tau)}, where the (to within 2% of the final value) approximates $4 \tau / (1 + [K](/page/K)). Thus, increasing the loop gain K shortens the , enhancing transient performance without invoking . In nonlinear systems, loop gain deviates from a fixed , varying with signal and due to elements like or . Time-domain simulations capture this variability qualitatively, revealing how nonlinearities alter error propagation and potentially introduce or limit cycles during transients. Direct measurement of time-domain loop gain poses challenges, as it requires isolating the loop traversal response amid transients, unlike sinusoidal injection in the frequency domain; practical assessment often depends on simulation tools like SPICE for injecting perturbations and observing convolutions via transient analysis.

Frequency-Domain Formulation

In the frequency domain, the loop gain is characterized by its response to sinusoidal inputs, represented as T(j\omega) = |T(j\omega)| \angle \phi(\omega), where \omega denotes the angular frequency, |T(j\omega)| is the magnitude, and \phi(\omega) is the phase angle. This polar form facilitates analysis of steady-state behavior in linear time-invariant systems, particularly for assessing amplification and phase shifts across frequencies. The formulation derives from the transfer functions of the forward path and feedback path. For a negative feedback system, the loop gain is given by T(j\omega) = A(j\omega) \beta(j\omega), where A(j\omega) is the forward gain and \beta(j\omega) is the feedback factor. This expression arises by substituting s = j\omega into the Laplace-domain loop gain T(s) = A(s) \beta(s), transforming the system into the for analysis. For practical plotting and design, the magnitude is often expressed on a logarithmic scale as T_{\text{dB}}(\omega) = 20 \log_{10} |T(j\omega)|, enabling Bode plots that visualize and phase contributions versus \log \omega. This decibel representation highlights regions of high loop gain for noise rejection and low gain for stability at higher frequencies. The unity gain frequency, denoted \omega_c, is defined as the angular frequency where |T(j\omega_c)| = 1 (equivalently, 0 dB). This crossover point serves as a reference for stability metrics, such as , by indicating where the loop gain transitions from to . As an illustrative example, consider an in an configuration with through a , at frequencies below the corner \omega \ll 1/(RC) where the factor \beta(j\omega) \approx 1. Under the single-pole for the op-amp's A(j\omega) \approx \frac{\omega_c}{j\omega}, the loop gain simplifies to T(j\omega) = \frac{\omega_c}{j\omega}. The magnitude |T(j\omega)| = \frac{\omega_c}{\omega} exhibits a -20 / roll-off, while the phase remains constant at -90°, reflecting the dominant pole's influence on the system's .

Historical Development

Early Recognition

In the early , the development of circuits for radio receivers and audio amplifiers highlighted the role of in causing system instabilities, such as unwanted oscillations and in public address systems, prompting engineers to explore the underlying mechanisms in closed loops. These observations built on informal references to feedback loops in and radio design during the and early , where regenerative circuits demonstrated how partial could amplify signals but risked instability if not controlled. A key milestone occurred in 1927 when Harold S. Black, an engineer at Bell Laboratories, invented the to address distortion and gain instability in long-distance telephone repeaters. Black's approach involved feeding a portion of the output back to the input in antiphase, relying on high loop gain (the product of amplifier gain A and factor β) to make the closed-loop gain approximately 1/β, independent of variations in A. This innovation provided a practical framework for using loop gain to enhance system performance and stability, though initial patent delays until 1937 limited immediate adoption. A pivotal advancement came in 1921 when German physicist Heinrich Barkhausen introduced the concept of loop gain explicitly through his oscillation criterion, stating that sustained oscillations occur when the loop gain has a magnitude of 1 and a shift of 360° (or equivalently 0° incorporating the negative for ). However, this early perspective viewed loop gain primarily as a constant value, neglecting its dependence on frequency, which often led to practical design challenges like unanticipated oscillations in broadband applications.

Key Theoretical Advances

In the 1930s, engineers at Bell Laboratories formalized the concept of loop gain through frequency-domain analysis, providing tools for precise prediction of system stability in feedback amplifiers. This era marked a shift from qualitative assessments to quantitative methods, enabling engineers to evaluate how loop gain influences closed-loop performance without extensive time-domain simulations. A pivotal advance came from in 1932, who introduced the . This method assesses closed-loop stability by examining the encirclement of the critical point (-1, 0) in the by the open-loop T(jω) as ω varies from 0 to infinity. The criterion states that the closed-loop system is stable if the Nyquist plot of T(jω) encircles the critical point a number of times equal to the number of right-half-plane poles of the open-loop system, in the counterclockwise direction. This graphical technique revolutionized stability analysis by directly linking loop gain characteristics to potential oscillations. Building on Nyquist's work, Hendrik Bode advanced loop gain theory in the late by developing Bode plots, which graph the magnitude and phase of the loop gain in decibels and degrees against the logarithm of frequency. These plots reveal gain-phase relationships critical for stability, such as gain and phase margins, allowing engineers to design feedback systems that avoid instability margins below 6 dB for gain and 45 degrees for phase. Bode's approach emphasized the integral relationship between gain and phase, providing a practical framework for optimizing bandwidth and distortion reduction. In 1943, R. B. Blackman extended loop gain concepts to impedance networks with his impedance formula, which calculates the change in port impedance due to as Z = Z_0 (1 + T) / (1 + T_s), where Z_0 is the impedance without feedback, T is the return ratio around the port, and T_s is the return ratio with the port shorted. This formula facilitated the design of feedback amplifiers by accounting for bilateral effects in multi-port systems, influencing in high-frequency circuits. These theoretical advances proved instrumental during , enabling the development of reliable systems for tracking and secure communications at Bell Laboratories and affiliated projects. Frequency-domain tools like Nyquist and Bode methods were applied to servo-mechanisms in systems, ensuring stable automatic control under noisy battlefield conditions and supporting over 100 variants deployed by Allied forces. Bode further refined terminology in his 1945 work by shifting from "loop gain" to "return ratio" to encompass more general configurations beyond simple unilateral loops, promoting broader applicability in network analysis.

Applications

In Electronic Circuits

In electronic circuits, loop gain plays a pivotal role in the of feedback amplifiers, where it quantifies the product of the and the feedback factor, enabling high closed-loop gain while maintaining . In configurations such as the non-inverting (op-amp), the feedback factor β is given by β = R₁ / (R₁ + R₂), where R₁ and R₂ form the , and the loop gain T approximates -A β, with A being the op-amp's ; this setup ensures the closed-loop gain closely follows 1/β = 1 + R₂/R₁ for large |T|, minimizing sensitivity to variations in A. is achieved when the magnitude of T is sufficiently greater than unity across the of interest, preventing phase shifts that could lead to , as analyzed through frequency-domain plots of T(jω). For multi-stage amplifiers, loop gain must account for loading effects between stages, which can reduce the effective factor β and alter the overall . In a two-stage common-emitter with shunt-series , for instance, the input decreases significantly (e.g., from 1.7 kΩ to 0.035 kΩ) due to feedback loading, while the output increases toward , stabilizing the gain against stage interactions; the loop gain Aβ is computed as -β_f × A_Io, where β_f is derived from emitter and resistors, ensuring dominates despite interstage capacitances and . Practical measurement of loop gain in such circuits involves breaking the loop at a low-impedance point, such as the controlled , injecting a test voltage or current signal, and measuring the return ratio; this method, often using Middlebrook's double-injection technique, yields T = (T_v T_i - 1) / (T_v + T_i + 2), where T_v and T_i are voltage and current return ratios, avoiding disruption to . In oscillator design, loop gain is intentionally set to unity magnitude with zero shift at the desired to sustain oscillations. The Wien bridge oscillator exemplifies this, employing a non-inverting op-amp with a frequency-selective RC network in the feedback path; at the oscillation ω₀ = 1/(RC) for equal R and C values, the loop gain T(jω₀) = 1 ∠0° when the amplifier K = 3, achieved via R_F = 2R_G, with a nonlinear element like a or regulating to prevent amplitude growth or . A specific application of loop gain derivation appears in shunt-shunt feedback transimpedance amplifiers, which convert input current to output voltage with high bandwidth. Using two-port y-parameters for the , the loop gain T is derived based on return ratio analysis, ensuring by verifying |T| < 1 outside the while maximizing transimpedance gain Z_T ≈ -Z_F for large |T|.

In Control Systems

In control systems, loop gain plays a central role in the design and performance of closed-loop configurations, particularly in unity feedback systems where the feedback transfer function H(s) = 1. The loop gain is defined as L(s) = G(s)C(s), with G(s) representing the plant or process and C(s) the controller , such as a PID controller. This formulation determines the overall system behavior, where the closed-loop becomes T(s) = \frac{L(s)}{1 + L(s)}, influencing regulation accuracy and response characteristics. High loop gain at low frequencies, often achieved through integral in C(s), minimizes steady-state errors by driving the tracking error to zero for step . The placement of poles and zeros in the loop gain L(s) significantly shapes the system's transient response and steady-state error via techniques like root locus analysis. As the gain parameter varies, the root locus plots the trajectories of closed-loop poles, revealing how increases in loop gain can shift poles toward the imaginary axis, improving steady-state accuracy through higher velocity or position error constants but potentially reducing damping and causing overshoot. For instance, adding zeros via derivative control in C(s) can pull poles leftward, enhancing stability margins and faster settling times, while integral terms introduce poles at the origin to eliminate offset at the cost of possible slower transients. This analysis guides controller design to balance error reduction with acceptable dynamic performance. In process control applications, such as regulation in chemical reactors or HVAC systems, loop gain tuning directly affects and dynamic response. A high DC loop gain reduces steady-state for setpoint changes or disturbances, ensuring precise maintenance, but excessive gain can lead to overshoot and oscillations, potentially damaging equipment. For example, in a controller, proportional-integral settings that yield a loop gain crossover well below the process dynamics minimize overshoot while achieving near-zero . For digital implementations in sampled-data systems, the loop gain is expressed in the z-domain as L(z) = G(z)C(z)H(z), where discretizations of continuous models must account for sampling effects like . occurs when high-frequency components fold into the due to the sampling rate, distorting the perceived loop gain and potentially destabilizing the system if the is too low relative to process . Design practices recommend sampling rates at least 10-20 times the closed-loop to mitigate these issues, ensuring the discrete loop gain approximates its continuous counterpart. Tuning methods like leverage loop characteristics for practical controller parameterization. The closed-loop variant involves increasing the proportional until sustained oscillations occur at the ultimate frequency \omega_u, where the loop magnitude satisfies |L(j\omega_u)| = 1 for , yielding the ultimate K_u. Controller parameters are then set as fractions of K_u and the ultimate period P_u = 2\pi / \omega_u, such as K_p = 0.6 K_u for , providing a starting point for robust performance in processes like flow or level . This empirical approach, developed in 1942, remains widely used for its simplicity despite assumptions of linear, second-order .

Analysis and Stability

Role in System Stability

In negative feedback systems, the stability of the closed-loop response is determined by the loop gain T(j\omega). For systems with no open-loop right-half-plane poles, the Nyquist plot of T(j\omega) must not encircle the critical point -1 on the to ensure . In general, the number of closed-loop right-half-plane poles is given by Z = P + N, where P is the number of open-loop right-half-plane poles and N is the number of clockwise encirclements of -1; requires Z = 0 (thus N = -P). This , originally formulated for regeneration in systems, assesses whether introduces unstable poles by relating encirclements to open-loop poles. The margin, defined as the of the of the loop at the crossover where \angle T(j\omega) = -180^\circ, quantifies how much the can increase before instability occurs; values greater than 1 (or 0 ) indicate , with larger margins providing robustness against variations. The phase margin further assesses stability by measuring the additional phase lag tolerable at the gain crossover frequency \omega_c, where |T(j\omega_c)| = 1, calculated as \phi_m = 180^\circ + \angle T(j\omega_c). Typical phase margins exceeding 45° ensure adequate and minimal overshoot in the , as lower values lead to increased approaching . At the oscillation threshold, known as the Barkhausen condition, marginal stability arises when |T(j\omega)| = 1 and \angle T(j\omega) = -180^\circ (or equivalently 180° considering the sign), causing the feedback signal to reinforce itself indefinitely without decay or growth. Increasing the magnitude of the loop gain |T(j\omega)| extends the closed-loop , enabling faster response times, but it simultaneously reduces margins if phase lag accumulates beyond the gain crossover, potentially driving the toward zero and inducing oscillations. In such cases, high loop gain amplifies sensitivity to component tolerances or environmental changes that alter characteristics. Certain feedback systems exhibit conditional , where holds for low or high loop gain values but fails at intermediate gains due to the Nyquist of T(j\omega) encircling the - point only in that range, often from multiple shifts or right-half-plane zeros. This phenomenon requires careful gain selection to avoid instability during operation or perturbations, as the 's trajectory may loop around the critical point for specific gain levels.

Graphical Analysis Methods

Graphical analysis methods provide visual tools for evaluating the loop gain T(j\omega) in the frequency domain, enabling engineers to assess stability and performance without solving complex equations directly. These techniques plot the magnitude and phase of the open-loop transfer function across frequencies, revealing how the system behaves near critical points that could lead to instability. The Bode plot consists of two separate graphs: one for the magnitude of T(j\omega) in decibels (dB), plotted against the logarithm of angular frequency \omega, and another for the phase angle in degrees versus \log \omega. The magnitude plot uses straight-line asymptotes to approximate the response; for instance, each pole contributes a slope of -20 dB per decade beyond its corner frequency, while each zero adds +20 dB per decade. These asymptotes simplify sketching for design purposes, with actual curves deviating slightly near corner frequencies due to the $1/(1 + j\omega/\omega_p) term for poles. The Nyquist plot represents T(j\omega) as a polar diagram in the complex plane, tracing the real part versus the imaginary part as \omega varies from 0 to \infty. For complete analysis, the plot includes a below the real for negative frequencies, forming a closed when combined with a at . is determined by the number of clockwise encirclements of the critical ; the system is stable if the number of encirclements equals the negative of the number of open-loop right-half-plane poles. The Nichols chart plots the open-loop magnitude in dB against the phase angle in degrees, effectively rotating the Nyquist plot by 180 degrees for easier readability. It overlays constant-M (closed-loop magnitude) and constant-N (closed-loop phase shift) contours, allowing direct prediction of closed-loop frequency response from the open-loop curve's position relative to these loci. For example, the peak closed-loop magnitude M_r is read from the highest M contour intersected by the plot, indicating potential resonance. Gain and phase margins are extracted from these plots to quantify stability margins. In the Bode plot, the gain margin is the difference in dB from 0 dB to the magnitude at the phase crossover frequency where the phase is -180°, while the is 180° plus the phase at the gain crossover frequency where the magnitude is 0 dB. On the Nyquist plot, the gain margin corresponds to the reciprocal of the distance from the origin to the intersection with the negative real axis, and the to the angle from that point to -1. The Nichols chart facilitates margin reading by marking the -180° line for gain margin (distance above 0 dB) and the 0 dB line for (deviation from -180°). Desirable values include gain margins greater than 6 dB and phase margins between 30° and 60° for robust performance. Modern software tools automate these graphical analyses from system models. and , for instance, generate Bode, Nyquist, and Nichols plots using commands like bode, nyquist, and nichols, while the margin function computes and margins directly. These tools integrate with SISOTOOL for interactive shaping, enhancing design efficiency beyond manual sketching.

References

  1. [1]
    Loop Gain and its Effect on Analog Control Systems
    Jan 26, 2015 · As discussed above, the accuracy of an amplifier's gain is determined by the loop gain of the amplifier: more loop gain in an amplifier means ...
  2. [2]
    Calculate Loop Gain Using the Voltage Injection Method
    Calculating loop gains is a way to determine the stability and transient response of a control system. Loop gain is the product of all gains present in a loop, ...
  3. [3]
    Loop Gain - an overview | ScienceDirect Topics
    Loop gain is defined as the difference between the open-loop gain and the closed-loop gain of an amplifier system, indicating the amount of negative feedback ...
  4. [4]
    None
    Below is a merged summary of all segments related to "Loop Gain" from *Feedback Control of Dynamic Systems, 8th Ed.*, consolidating the information into a comprehensive response. To retain all details efficiently, I will use a table in CSV format for key information (definitions, formulas, context, etc.) and provide a narrative overview for additional context, sign conventions, and URLs. This approach ensures maximum density while maintaining readability and completeness.
  5. [5]
    [PDF] The Bell System Technical Journal January, 1934 Stabilized ...
    Stabilized Feedback Amplifiers*. By H. S. BLACK. This paper describes and explains the theory of the feedback principle and then demonstrates how stability ...Missing: Harold | Show results with:Harold
  6. [6]
    Negative Feedback, Part 2: Improving Gain Sensitivity and Bandwidth
    Nov 11, 2015 · We are now well aware that negative feedback can improve two critical amplifier characteristics—bandwidth and sensitivity to open-loop gain— ...
  7. [7]
    2.3: ADVANTAGES OF FEEDBACK - Engineering LibreTexts
    May 22, 2022 · The desensitivity is identically equal to the ratio of the forward-path gain to closed-loop gain. Feedback connections are unique in their ...
  8. [8]
    5.3: Gain-Bandwidth Product - Engineering LibreTexts
    May 22, 2022 · ... negative feedback lowers the midband gain. To a first approximation, this gain will continue until it reaches the open loop response. At ...<|control11|><|separator|>
  9. [9]
    Negative Feedback Systems - Electronics Tutorials
    Negative feedback also has effects of reducing distortion, noise, sensitivity to external changes as well as improving system bandwidth and input and output ...
  10. [10]
    Benefits of Negative Feedback - HyperPhysics
    Decreasing Distortion with Feedback. The use of negative feedback can discriminate against sources of noise or distortion within an amplifier.
  11. [11]
    2.5: EFFECTS OF FEEDBACK ON INPUT AND OUTPUT IMPEDANCE
    May 22, 2022 · The amount by which feedback scales input and output impedances is directly related to the loop transmission, as shown by the following example.
  12. [12]
    [PDF] Introduction to Feedback
    Feedback defines closed-loop gain, modifies input/output impedances, increases bandwidth, and reduces nonlinearity. It includes a feedforward amp, sensing ...
  13. [13]
    [PDF] Signal Gain and Noise Gain of the Op-Amp
    Nov 28, 2022 · Where A>>1, or 1/A << β, Equation (12) is as follows. VOUT / VNOISE ... So far, we have explained that the feedback circuit, the negative feedback ...
  14. [14]
    [PDF] Feedback Systems
    ... convolution integral y(t) = ∫. ∞. 0 h(t − τ)u(τ) dτ, where h(t) is the impulse response for the system (assumed causal). Taking the. Laplace transform of ...
  15. [15]
    [PDF] Linear Feedback Loops - OTH Regensburg
    Jan 8, 2019 · Expressions like Y=STFꞏX in frequency domain correspond to a convolution (deutsch: Faltung) in time-domain, according to. Page 3. Martin J.W. ...
  16. [16]
    [PDF] Feedback Systems - Electrical Engineering and Computer Science
    high gain: as K → ∞, y → u and e → 0. - “open loop gain”: |KG| 1. - “closed loop gain”: |KG/(1 + KG)| ≈ 1. ⇒ we can make the output track the input even if we ...
  17. [17]
    The Unit Step Response - Swarthmore College
    Key Concept: Step response of a first order system. The unit step response of a first order system is: yγ(t)=yγ(∞)+(y(0+)−yγ(∞))e−t/τ=H(0)+(H(∞)−H(0))e−t/τ ...
  18. [18]
    [PDF] Time Domain Stability Margin Assessment
    The gain and phase margins derived from the nonlinear time domain simulations demonstrate good overall agreement to the frequency domain LTI-derived margins.
  19. [19]
    Estimating by time domain simulation the LoopGain transfer function
    Mar 2, 2024 · Relevant video How to get equally spaced data in LTspice? https://youtu.be/TGWlvHw2_e0.
  20. [20]
    [PDF] Feedback Systems
    Jul 24, 2020 · Feedback systems use frequency domain techniques, focusing on the open loop transfer function to design controllers and shape the closed loop ...
  21. [21]
    Integrator Limitations: The Op-Amp's Gain Bandwidth Product
    Aug 28, 2019 · ... T = aβ is called the loop gain. Moreover, β is called the feedback factor. We find it by setting the input to zero, breaking the loop at the ...
  22. [22]
    [PDF] Introduction to “Stabilized Feed-Back Amplifiers” | Cascade Tubes
    Black recognized that by using an amplifier comprising several vacuum-tube stages in cascade to yield a very high open- loop gain, and then “killing” most of ...<|control11|><|separator|>
  23. [23]
    [PDF] H Op Amp History - Analog Devices
    Working as a young Western Electric Company engineer on telephone channel amplifiers, Harold S. Black first developed feedback amplifier principles. Note that ...
  24. [24]
    Barkhausen Stability Criterion - MIT
    Nov 14, 2002 · The Barkhausen Stability Criterion is simple, intuitive, and wrong. During the study of the phase margin of linear systems, this criterion is often suggested ...
  25. [25]
    [PDF] Regeneration Theory - By H. NYQUIST
    Regeneration Theory. By H. NYQUIST. Regeneration or feed-back is of considerable importance in many appli- cations of vacuum tubes. The most obvious example ...
  26. [26]
    [PDF] Effect of Feedback on Impedance - World Radio History
    By R. B. BLACKMAN. THE impedance of a network is defined as the complex ratio of the alter- nating potential difference maintained across its terminals by ...
  27. [27]
    Brief History of Feedback Control - F.L. Lewis
    Unfortunately, Airy discovered that by improper design of the feedback control loop, wild oscillations were introduced into the system. He was the first to ...Missing: telephony | Show results with:telephony
  28. [28]
    [PDF] Feedback | ECEN326: Electronic Circuits Spring 2022
    CH 12 Feedback. 13. Gain Desensitization. ➢ A large loop gain is needed to create a precise gain, one that does not depend on A. 1. , which can vary by ±20 ...
  29. [29]
    [PDF] EECS 240 Analog Integrated Circuits Topic 11: Feedback
    The loop gain can be calculated directly by breaking the feedback loop and injecting a test signal and observing the “return” ratio. The return signal should ...
  30. [30]
    [PDF] Feedback
    Figure 8.40 (a) A gain stage in a multistage amplifier ... Figure 8.44 (a) Magnitude and (b) phase of the loop gain Aβ of the feedback amplifier circuit in Fig.
  31. [31]
    [PDF] Oscillators - Marshall Leach
    This resistance change varied the gain of the amplifier to maintain a loop gain of unity. Wien Bridge Oscillator. Figure 3 shows an op-amp Wien Bridge ...
  32. [32]
    None
    ### Summary of Shunt-Shunt Feedback in Transimpedance Amplifiers
  33. [33]
    [PDF] Feedback Control Systems Loop Shaping Design With Practical ...
    Loop shaping mainly consists of designing the frequency domain response of L(s) to satisfy control system requirements and then given the transfer function of ...
  34. [34]
    Loop Gain Function - an overview | ScienceDirect Topics
    The loop gain function is defined as the transfer function that represents the amplitude and phase modifications introduced to an input drive by an ...
  35. [35]
    [PDF] Chapter Eight Root Locus Control Design 8.3 Common Dynamic ...
    In the following we present dynamic controller design techniques in three categories: improvement of steady state errors (PI and phase-lag controllers),.
  36. [36]
    Learning PID loop tuning from an expert - Control Engineering
    Dec 4, 2014 · It can help reduce the oscillation time and remove the offset, but improper adjustment can cause an increase in overshoot as well as lead to the ...
  37. [37]
    [PDF] Discrete Time Control Systems
    • Convolution in the time domain is equivalent to multiplication in the ... Notice that even in this limiting case the LQG open-loop gain does not reach the LQR ...
  38. [38]
    [PDF] Control of dynamic sampled-data systems with frequency aliasing
    Jan 1, 2004 · H (z) is a discrete model of the continuous part of the sampled-data system. The response of the analog system at a given frequency f is the sum ...
  39. [39]
    9.3: PID Tuning via Classical Methods - Engineering LibreTexts
    Mar 11, 2023 · The Ziegler-Nichols closed-loop tuning method allows you to use the ultimate gain value, Ku, and the ultimate period of oscillation, Pu, to ...Introduction · Ziegler-Nichols Method · Cohen-Coon Method · Other MethodsMissing: jω_u | Show results with:jω_u
  40. [40]
    Optimum Settings for Automatic Controllers | J. Fluids Eng.
    Dec 20, 2022 · In this paper, the three principal control effects found in present controllers are examined and practical names and units of measurement are proposed for each ...
  41. [41]
    Regeneration Theory - Nyquist - 1932 - Bell System Technical Journal
    Regeneration or feed-back is of considerable importance in many applications of vacuum tubes. The most obvious example is that of vacuum tube oscillators.
  42. [42]
  43. [43]
    Gain margin, phase margin, and crossover frequencies - MathWorks
    In other words, the gain margin is 1/g if g is the gain at the –180° phase frequency. Similarly, the phase margin is the difference between the phase of the ...
  44. [44]
  45. [45]
    Understand Power Supply Loop Stability and Loop Compensation
    Jan 27, 2022 · In the 1930s, another Bell Labs engineer, Hendrik Wade Bode, devised a simple method for graphing gain and phase shift plots. They are known as ...
  46. [46]
    Negative Feedback, Part 5: Gain Margin and Phase Margin
    Nov 23, 2015 · More phase margin means more stability, because higher phase margin indicates that the frequency at which the loop gain magnitude reaches unity ...
  47. [47]
    [PDF] Conditional Stability in Feedback Systems
    If the loop gain is 100, in negative feedback the closed loop gain (Y/X) would be G(s)/101, while in positive feedback the closed loop gain would be –G(s)/99.
  48. [48]
    [PDF] Chapter 7. Loop Analysis of Feedback Systems
    We will illustrate the issue of conditional stability by an example. Notice that Nyquist's theorem does not hold if the loop transfer function has a pole in the ...
  49. [49]
    [PDF] 16.30 Topic 3: Frequency response methods - MIT OpenCourseWare
    Sep 15, 2010 · Nichols Plot – |G(jω)| vs. ¿G(jω), which is very handy for sys tems with lightly damped poles. 3. Bode Plot – Log |G ...
  50. [50]
    [PDF] Loop Analysis - ResearchGate
    To explore this further we will introduce a graphical representation of the loop transfer function. The Nyquist Plot. The frequency response of the loop ...
  51. [51]
    [PDF] Control System Design Based on Frequency Response Analysis
    Plot the open-loop. Bode diagram and determine the gain and phase margins for each controller. Page 31. 31. Chapter 14. Figure 14.11. Comparison of ...<|control11|><|separator|>
  52. [52]
    [PDF] The Nichols Chart - MIT OpenCourseWare
    The Nichols chart was once very useful, since computers were not available to do the kids of calculations that are now done by e.g., Matlab.
  53. [53]
    [PDF] ANALYSIS AND DESIGN OF SPACE VEHICLE FLIGHT CONTROL ...
    Consequently, one can deal with phase and gain margins on a Bode or Nichols plot only when these quantities have been correctly related to a Nyquist locus. 3.1.