Fact-checked by Grok 2 weeks ago

Output impedance

Output impedance, also known as source impedance, is the impedance of an electrical or device as seen looking into its output terminals when no load is connected and all independent voltage sources are short-circuited while all independent sources are opened. This is equivalent to the Thevenin equivalent (or impedance) at the output port, representing the internal opposition to flow from the to an external load. In practical terms, output impedance determines the extent to which the output voltage of a drops when a load is applied, as the load impedance forms a with the output impedance. For effective signal transfer in amplifiers and other active devices, a low output impedance is desirable—ideally approaching zero—to minimize loading effects and ensure that the full intended voltage is delivered to the next stage without significant . This is particularly critical in applications like audio amplifiers, where high output impedance can lead to reduced power delivery and distorted signals, or in , where it influences under varying loads. Output impedance is typically measured using the Thevenin theorem by deactivating sources and calculating the equivalent impedance, or experimentally by applying a test signal at the output and observing the ratio of voltage to current. In frequency-dependent contexts, such as circuits, it may include reactive components ( or ), but it is often predominantly resistive in the midband frequency range for many amplifiers. Understanding and controlling output impedance is essential for , which maximizes power transfer between stages and prevents reflections in high-frequency systems like RF circuits.

Fundamentals

Definition

Output impedance is the equivalent impedance presented by a device or circuit at its output terminals, equivalent to the Thevenin impedance looking back into those terminals with all internal independent sources deactivated (voltage sources short-circuited and current sources open-circuited). This concept, formalized in late 19th-century circuit theory, builds directly on , proposed by Léon Charles Thévenin in 1883 as a method to simplify linear electrical networks into an equivalent voltage source in series with an impedance. It serves as the dual to , which is similarly defined looking into the input terminals. In alternating current (AC) circuits, output impedance is typically frequency-dependent due to reactive components like inductors and capacitors, and it is represented as a complex quantity Z_\text{out} = R + jX, where R is the real (resistive) part and X is the imaginary (reactive) part. For direct current (DC) analysis, it reduces to a purely resistive value. A low output impedance is desirable for devices functioning as voltage sources, as it ensures the output voltage remains stable regardless of load variations by minimizing voltage division effects. An ideal voltage source has zero output impedance, delivering constant voltage without drop; in real sources with non-zero Z_\text{out}, connecting a load Z_\text{load} results in a reduced load voltage given by the voltage divider formula V_\text{load} = V_\text{source} \cdot \frac{Z_\text{load}}{Z_\text{out} + Z_\text{load}}.

Theoretical Models

The Thevenin equivalent circuit represents the output port of a linear as an ideal V_{th} in series with an output impedance Z_{out}, where Z_{out} is the equivalent impedance seen at the output terminals when all independent sources are deactivated—voltage sources are short-circuited and current sources are open-circuited. This model simplifies analysis by reducing complex circuits to a basic form that accurately predicts voltage and current behavior under load. To compute Z_{out}, the deactivated circuit's impedance is calculated looking back into the output port, often involving series-parallel combinations of resistors, capacitors, and inductors. The equivalent provides a of the same network, consisting of an ideal I_n in parallel with the identical output impedance Z_{out}, derived from the same deactivation process as the Thevenin model. The equivalence holds because source transformations between voltage-current pairs preserve the port's I-V characteristics, with Z_{out} remaining unchanged in both models. In modeling output impedance, small-signal analysis employs linear approximations around a operating point to characterize behavior, treating the circuit as a linear network where Z_{out} is frequency-dependent but constant for small perturbations. This approach is suitable for audio or RF amplifiers, using -\pi or T-models to derive impedances like output r_o. In contrast, large-signal models account for nonlinear effects in power applications, where Z_{out} varies with signal amplitude due to or regions, requiring time-domain simulations rather than linear equivalents. For a simple formed by two resistors R_1 (from input to output) and R_2 (from output to ), the output impedance Z_{out} is the combination of R_1 and R_2 when the input is shorted to . The derivation begins by deactivating the source, leaving R_1 connected from the output node to and R_2 directly to , forming a parallel network. The equivalent impedance is then: Z_{out} = \frac{R_1 R_2}{R_1 + R_2} This follows from the parallel resistance formula $1/Z_{out} = 1/R_1 + 1/R_2, confirming the unloaded Thevenin resistance at the output port. In reactive circuits, output impedance exhibits frequency dependence, as modeled by Z_{out}(j\omega), where \omega is the angular frequency. For instance, in a circuit with resistive R shunted by a capacitor C, the output impedance is the parallel combination: Z_{out}(j\omega) = R \parallel \frac{1}{j\omega C} = \frac{R \cdot \frac{1}{j\omega C}}{R + \frac{1}{j\omega C}} = \frac{R}{1 + j\omega R C} At low frequencies, the capacitive reactance $1/(j\omega C) dominates, making Z_{out} \approx R; at high frequencies, the capacitor shorts, reducing Z_{out} toward zero. This variation influences bandwidth and stability in AC-coupled systems.

Measurement

Direct Techniques

Direct techniques for measuring output impedance involve physical electrical probing to directly assess the impedance at the output terminals of a device or circuit, typically modeled using the Thévenin equivalent where the output appears as a in series with the impedance. These methods rely on applying known test conditions, such as loads or shorts, and measuring resulting voltages or currents with calibrated instruments to compute the impedance value. The method, suitable for voltage sources like amplifiers or signal generators, determines output impedance by observing the across a known load . First, the open-circuit voltage V_{oc} is measured across the unloaded output terminals using a high-impedance or . Then, a precision load resistor Z_{load} (often purely resistive for low-frequency cases) is connected, and the loaded voltage V_{load} is measured. The output impedance is calculated as Z_{out} = Z_{load} \times \frac{V_{oc} - V_{load}}{V_{load}}, derived from the principle in the Thévenin model. This approach assumes a linear response and is effective when Z_{load} is comparable to the expected Z_{out} to produce a measurable , typically 10-50% of V_{oc}. For current sources, such as certain power supplies or transducers, the short-circuit current method is preferred, where the output impedance is found as Z_{out} = \frac{V_{oc}}{I_{sc}}, with I_{sc} being the current measured when the output terminals are shorted using a low-resistance or shunt. Safety precautions are essential for high-power devices, including using current-limited supplies, protective fuses rated below the device's maximum output, insulated probes, and like gloves and safety glasses to prevent arcing or thermal damage during the short. This method avoids high voltages but requires careful handling to limit fault currents. Common equipment includes oscilloscopes for AC voltage measurements, function generators to drive the device under test, precision resistors (e.g., 0.1% tolerance) as loads, and multimeters for DC or low-frequency readings; calibration involves verifying instrument accuracy against known standards, such as shorting inputs for zero offset or using precision voltage sources, to minimize errors from probe capacitance or resistance mismatches. For DC output impedance, steady-state measurements are used after allowing transients to settle (e.g., seconds to minutes), focusing on resistive components with multimeters. In contrast, AC measurements capture frequency-dependent impedance using network analyzers for RF sweeps (e.g., 1 MHz to 3 GHz), where vector measurements of voltage and current phases enable complex Z_{out} calculation via the RF I-V method. An example procedure for low-frequency circuits (e.g., audio amplifiers at 1 kHz) begins by powering the device and connecting its output to an channel set to AC coupling with 1 MΩ to measure V_{oc} (ensure no load, verify signal stability). Next, select a precision R_{load} (e.g., 50 Ω for expected low Z_{out}) and connect it across the output, measuring V_{load} on the same channel while monitoring for . Compute Z_{out} = R_{load} \times \left( \frac{V_{oc}}{V_{load}} - 1 \right); repeat with multiple loads (e.g., 10 Ω, 100 Ω) for averaging. Error analysis shows typical accuracy limits of ±5%, arising from loading (mitigated by high-impedance probes), tolerance (±0.1%), and signal noise (reduced via averaging over 10 cycles), with systematic errors like cable contributing <1% below 10 kHz.

Indirect Techniques

Indirect techniques for measuring output impedance infer the value from observed responses or models without directly injecting signals at the output terminals, often useful when physical access is limited or non-invasive assessment is preferred. These methods rely on varying external conditions, transient behaviors, or computational models to estimate , providing insights into circuit performance under operational constraints. The load variation method involves applying a series of known loads to the output and measuring the resulting changes in output voltage or current. By plotting the output voltage against load resistance (or current against load conductance), the output impedance is extrapolated from the slope of the linear region, typically using a least-squares fit to minimize measurement errors. For instance, with two distinct resistive loads R_L1 and R_L2, the output impedance can be calculated as Z_out = (V1 * R_L2 - V2 * R_L1) / (V2 - V1), where V1 and V2 are the corresponding output voltages for a fixed input. This approach is particularly effective for DC or low-frequency measurements in amplifiers and power circuits, as it leverages the . Simulation-based inference employs circuit simulation tools like SPICE to model the device and extract Z_out through small-signal analysis. In LTspice or similar software, an AC analysis is performed by injecting a 1 A AC current source at the output node with the input source set to zero, then computing Z_out as the ratio of output voltage to injected current (Z_out = |V_out / I_inj|) across a frequency sweep. Convergence can be improved by adding small resistors (e.g., 1 mΩ) in series with voltage sources, enabling initial conditions, or using .UIC directives for transient-assisted AC simulations; for complex models, the .NET directive computes network parameters including output impedance directly. This method allows exploration of parametric variations and parasitic effects without hardware, though results depend on the accuracy of component models. In RF applications, non-invasive methods use a vector network analyzer (VNA) to measure S-parameters and convert S_{22} to output impedance via Z_out = Z_0 \frac{1 + S_{22}}{1 - S_{22}}, where Z_0 is the reference impedance (typically 50 Ω). The device under test is connected to port 2 of the VNA with port 1 terminated, and S_{22} is measured over frequency; this reflection coefficient-based approach avoids direct probing by treating the output as a one-port network. It excels for high-frequency characterization, providing magnitude and phase data for complex Z_out. These indirect techniques share limitations tied to underlying assumptions, such as ideal load linearity or accurate model parasitics; errors arise from unmodeled effects like cable inductance, fixture residuals, or non-stationary transients, potentially degrading accuracy by 10-20% at high frequencies. Validation against direct methods is recommended for critical applications.

Applications in Electronics

Amplifiers

In voltage amplifiers, a low output impedance (Z_{out}) is essential to ensure efficient power transfer to the load while minimizing voltage drop and gain reduction. Ideally, Z_{out} should be much less than 1 Ω to drive varying loads without significant attenuation of the output signal. When a load impedance Z_{load} is connected, the effective voltage gain becomes A_{v,loaded} = A_v \frac{Z_{load}}{Z_{load} + Z_{out}}, where A_v is the unloaded gain; this voltage divider effect highlights how even modest Z_{out} can degrade performance if not sufficiently low. Negative feedback plays a crucial role in optimizing Z_{out} for amplifiers, particularly operational amplifiers (op-amps) in configurations like noninverting or voltage-follower setups. By sampling the output and feeding a portion back to the input, negative feedback reduces the effective Z_{out} by the loop gain factor (1 + A\beta), where A is the open-loop gain and \beta is the feedback fraction; thus, Z_{out,closed} \approx \frac{Z_{out,open}}{1 + A\beta}. This derivation stems from the feedback stabilizing the output against load variations, enabling op-amps to achieve Z_{out} values as low as milliohms in closed-loop operation. Amplifier classes influence Z_{out} through their output stage topologies, with variations arising from biasing and conduction patterns. In Class A amplifiers, such as the emitter follower (common-collector) configuration, the output stage provides linear operation but at lower efficiency; here, Z_{out} \approx r_e + \frac{R_E}{\beta}, where r_e is the emitter resistance (approximately kT/qI_E \approx 26 mV/I_E), R_E is the emitter resistor, and \beta is the current gain, yielding low Z_{out} suitable for buffering. Class B and AB push-pull stages, using complementary pairs, improve efficiency (up to 78.5% theoretically for Class B) but introduce crossover distortion if not biased properly; their Z_{out} is similarly low due to the follower arrangement, though it rises slightly at low signal levels from reduced transconductance. Class C amplifiers, biased for tuned applications, exhibit higher Z_{out} variations due to nonlinear conduction, making them less ideal for broadband voltage amplification. High Z_{out} in amplifiers can lead to distortion and stability issues, particularly in audio applications where load impedance varies with frequency. Frequency-dependent loading from elevated Z_{out} causes uneven frequency response, manifesting as harmonic distortion (e.g., up to 2 dB deviations at bass frequencies with Z_{out} \approx 0.8 Ω for an 8 Ω load). The damping factor (DF), defined as DF = \frac{Z_{load}}{Z_{out}}, quantifies control over the load; low DF (<100) reduces damping of speaker resonances, potentially exciting oscillations with reactive loads and compromising stability. Amplifier damping factors greater than 300 are typically targeted to achieve a system-level DF above 150, limiting audible effects. A typical design example for a power amplifier output stage employs complementary MOSFETs in a source-follower configuration within a Class AB amplifier, biased at 50-100 mA quiescent current to minimize crossover distortion. Global negative feedback from the output reduces Z_{out} to approximately 0.1 Ω at audio frequencies (20 Hz to 20 kHz), enabling effective drive of 4-8 Ω speakers with minimal gain loss and high damping (DF >80). This setup, common in high-fidelity audio, uses devices like IRFP240/IRFP9240 pairs with emitter degeneration resistors (0.1-0.22 Ω) for thermal stability, achieving below 0.1% at 100 W output.

Signal Generators

In signal generators, particularly function generators used for RF applications, the ideal output impedance is standardized at 50 Ω to ensure maximum power transfer and minimize signal reflections in transmission lines. This value aligns with the of common cables, preventing standing waves that could distort the output . For instance, devices like the 33600A series function generators feature an optional fixed output impedance of 50 Ω on their primary output terminal, designed for direct connection to 50 Ω loads. When the load impedance mismatches this 50 Ω source, reflections occur, quantified by the voltage (VSWR), which measures the of the forward to reflected voltage wave along the line. A VSWR of 1 indicates perfect matching, while values greater than 1.5 can lead to significant signal or in RF testing setups. For oscillators within signal generators, such as those based on parallel LC tanks, the output impedance exhibits a pronounced peak at the resonance frequency f_0, where the inductive and capacitive reactances cancel, leaving a predominantly resistive response amplified by the circuit's quality factor. The peak impedance magnitude is given by Z_{\text{out}} = Q \cdot R at f_0, with Q representing the Q-factor (typically 50–200 for high-performance RF oscillators) and R the effective series resistance of the tank components. Higher Q-factors sharpen this peak, enhancing frequency selectivity but also increasing sensitivity to loading effects that can pull the oscillation frequency. This behavior is critical in voltage-controlled oscillators (VCOs) used for sweeping signals, as it determines the drive capability into subsequent stages without excessive damping. Calibration of output impedance relies on NIST-traceable 50 Ω terminations to verify performance across operational bandwidths, such as from 1 Hz to 1 GHz in versatile function generators. These terminations, often precision loads with tolerances below 0.5%, serve as reference standards during vector analyzer (VNA) measurements, ensuring the generator's impedance remains within specifications like ±5% variation. Measurements reveal that output impedance can deviate slightly with due to parasitic elements in the output , requiring periodic recalibration to maintain for applications. For example, Keysight's N4690D electronic calibration modules provide automated NIST-traceable verification for 50 Ω ports up to 18 GHz. High output impedance in signal generators can compromise when driving capacitive loads, such as probes or long cables, leading to droop at higher frequencies. The capacitive forms a with the source impedance, attenuating the signal as |Z_C| = 1/(2\pi f C) decreases, resulting in up to 50% loss above 1 MHz for unbuffered outputs into 100 pF loads. To mitigate this, output buffers—typically unity-gain op-amps like those in ' high-speed amplifiers—are employed to provide low-impedance driving (e.g., <1 Ω), isolating the generator's core from load variations while preserving waveform fidelity. A representative example is a sine wave generator configured for RF output, where Z_{\text{out}} = 50 + j0 \, \Omega ensures a purely resistive match into coaxial systems. In such setups, impedance transformation occurs via the cable's characteristic impedance, maintaining signal amplitude if terminated properly; for instance, generators deliver flat response up to 100 MHz into 50 Ω coax when buffered against reactive mismatches. This configuration is standard for precision testing, emphasizing the need for matching to avoid phase shifts or harmonic distortion.

Applications in Power Systems

Batteries

In electrochemical batteries, the output impedance, often referred to as internal impedance, represents the opposition to current flow within the cell and manifests as a combination of ohmic resistance (R_\Omega), charge transfer resistance (R_{ct}), and diffusion-related components. The ohmic resistance arises from the electrolyte, separators, and current collectors, typically measured at high frequencies (>10 kHz) in electrochemical impedance spectroscopy (EIS). Charge transfer resistance, associated with electrochemical reactions at the electrode-electrolyte interface, dominates around 10 Hz, while diffusion impedance accounts for transport limitations in the electrodes at low frequencies (<1 Hz). For lithium- (Li-) batteries, these components contribute to typical internal impedance values ranging from 10 to 100 mΩ for fresh cells, depending on chemistry, size, and state of charge. The effective output impedance increases with discharge rate, influencing battery performance through Peukert's law, which describes the reduction in available capacity at higher currents. This law is expressed as C_\text{eff} = C \left( \frac{I}{I_0} \right)^{1-n}, where C is the nominal capacity at reference current I_0, I is the discharge current, and n (Peukert exponent, typically 1.1–1.3 for Li-ion) reflects losses from elevated internal resistance and incomplete ion recovery between reactions. Higher discharge rates amplify ohmic and diffusion losses, leading to greater effective impedance and reduced usable capacity, particularly in high-power applications. Battery aging elevates output impedance over charge-discharge cycles, primarily due to solid electrolyte interphase (SEI) layer growth on the anode, which consumes lithium and electrolyte while increasing resistance to ion transport. This growth, driven by side reactions and mechanical cracking, raises charge transfer and diffusion impedances, contributing to power fade and capacity loss. AC impedance spectroscopy at 1 kHz is commonly used to quantify this, as it approximates the ohmic resistance and provides a standard metric for state-of-health assessment, with values rising non-monotonically—often decreasing slightly early in life before increasing toward end-of-life. Different battery types exhibit distinct output impedance profiles. Alkaline batteries, such as zinc-manganese dioxide cells, typically have higher internal resistance (around 200–500 mΩ per cell) due to their non-rechargeable chemistry and thicker separators, leading to more pronounced voltage drops under load compared to rechargeable alternatives. Lithium-polymer (LiPo) batteries, a variant of Li-ion with polymer electrolytes, maintain lower impedances (5–50 mΩ fresh) for better high-rate performance but are susceptible to similar aging mechanisms. For lead-acid batteries, fresh cells show low impedance (~5 mΩ), but degradation from sulfation and electrolyte stratification can elevate it to >50 mΩ, significantly impairing cranking power. Practically, elevated output impedance causes voltage sag under load, where the terminal voltage drops as V = E - I Z_\text{out} (with E as open-circuit voltage and I as current), reducing efficiency and triggering low-voltage cutoffs prematurely. In electric vehicle (EV) battery packs, for instance, if each Li-ion cell has ~1 mΩ internal resistance and the pack comprises 100 cells in series (with parallel strings scaling inversely), the total series impedance could reach 100 mΩ, resulting in a 100 V sag at 1000 A draw—highlighting the need for low-impedance cells to minimize range loss and heat generation.

Power Supplies

In electronic power supplies, output impedance plays a critical role in maintaining stable voltage delivery to loads, particularly under varying current demands. Linear power supplies achieve very low output impedance, typically on the order of 1–10 mΩ, through the use of pass transistors that provide negative feedback to minimize voltage variations. This low impedance ensures excellent load regulation, where the change in output voltage per unit change in load current, ΔV/ΔI, approximates the output impedance Z_out. Switching-mode power supplies (SMPS) also achieve low output impedance at low frequencies, typically on the order of 10–50 mΩ, though their output impedance increases at higher frequencies due to the switching operation and output filter characteristics. For capacitive output filters common in both linear and switching designs, the impedance is dominated by the capacitive reactance, given by Z_out ≈ 1/(2πfC), where f is the of interest and C is the . This formula highlights how larger capacitors or lower reduce Z_out, improving but potentially increasing size and cost. Load specifications directly tie to this, as poor regulation manifests as voltage droop proportional to load current and Z_out. Transient response to load steps is heavily influenced by output impedance, with lower Z_out enabling faster recovery times from sudden changes. For instance, supplies with Z_out below 0.1 Ω can recover from a full-load step in under 1 μs, minimizing voltage undershoot and supporting dynamic loads like processors. In AC generators such as alternators, output impedance is characterized by the synchronous X_s, typically 0.5–5 Ω per for medium-sized machines, which causes voltage drops under varying loads and impacts overall system stability. To mitigate voltage drops due to output impedance in distribution lines, techniques are employed, where sense wires connect directly to the load to adjust the supply's output voltage accordingly. For example, in a 5 V USB , maintaining Z_out below 0.5 Ω—often achieved through low-resistance pass elements and filtering—ensures compliance with USB standards for minimal voltage deviation at the load, even with lengths up to 3 m.

References

  1. [1]
    [PDF] Electronics II - NJIT
    Output impedance of a circuit is the impedance looking into the output terminals of the circuit without any load connected to the output and all internal and ...
  2. [2]
    [PDF] 1 DC Linear Circuits
    The output impedance of circuit A is simply its Thevenin equivalent resistance RTH. The output impedance is sometimes called “source impedance”. • The input ...
  3. [3]
    Input and output impedance | Advanced Teaching Labs - Physics
    Oct 13, 2020 · It determine how much current you need to draw from the input (simply Ohm's Law) ... It determine how much voltage will be shared by the black box ...
  4. [4]
    [PDF] EXPERIMENT 7 – BJT AMPLIFIER CONFIGURATIONS & INPUT ...
    Likewise, output impedance is in general a complex quantity, but is predominantly resistive in the midband range. To calculate output impedance one must remove ...
  5. [5]
    [PDF] Input and Output Impedance and Thevenin's Theorem
    If the circuit is viewed as providing current or voltage to a subsequent stage, the impedance is called the output impedance.Missing: history | Show results with:history
  6. [6]
    (PDF) Origins of the equivalent circuit concept: The voltage-source ...
    Aug 5, 2025 · The voltage-source equivalent was first derived by Hermann von Helmholtz (1821-1894) in an 1853 paper. Exactly thirty years later in 1883, Leon ...
  7. [7]
    Impedance and Complex AC Impedance - Electronics Tutorials
    But impedance is also frequency dependant and therefore has a phase angle associated with it. The phase angle of reactance, either inductive or capacitive ...
  8. [8]
    [PDF] 6.200 Notes: Thevenin-Helmholtz and Mayer-Norton Theorems
    Feb 23, 2023 · Output Impedance Thevenin resistance of a circuit when viewed at a port where a signal would normally be output from the circuit. Thevenin ...
  9. [9]
    How to Use Thevenin's Theorem | DC Network Analysis
    Applying Thevenin's theorem allows us to simplify any linear circuit to its Thevenin equivalent circuit with a single voltage source and series resistance. This ...Missing: history | Show results with:history
  10. [10]
    [PDF] Thevenin's and Norton's Equivalent Circuit Tutorial. (by Kim, Eung)
    Thevenin's uses a voltage source in series with impedance; Norton's uses a current source in parallel with impedance. Norton's is a source transformation of  ...Missing: output | Show results with:output
  11. [11]
    [PDF] ECEN326: Electronic Circuits Spring 2022 - Texas A&M University
    For small-signal analysis, we “linearize” the response about the DC operating point. • If the signal is small enough, linearity holds and the.
  12. [12]
    Voltage Divider Rule and Voltage Division - Electronics Tutorials
    Using the voltage divider ratio rule, we can see that the largest resistor produces the largest I*R voltage drop. Thus, R1 = 4V and R2 = 8V.
  13. [13]
    None
    ### Summary of Frequency Response in Reactive Circuits
  14. [14]
    [PDF] Lab 1a Input and Output Resistance
    The equivalent circuit consists of one battery and one resistor, namely the Thévenin equivalent source and the Thévenin equivalent output resistance.
  15. [15]
    [PDF] Measuring Input Impedance - Physics
    In the diagram above, Zout is the output impedance of the circuit to be measured. If you measure the open circuit voltage V (with no load), then add the load ...
  16. [16]
    [PDF] CIRCUITS LABORATORY EXPERIMENT 5
    In this experiment the student should learn: (1) How to measure the output impedance of a signal source,. (2) The circuit representation of an inductor,. (3) ...
  17. [17]
    [PDF] Keysight Technologies Impedance Measurement Handbook
    Many modern impedance measuring instruments measure the real and the imaginary parts of an impedance vector and then convert them into the desired parameters ...
  18. [18]
    Impedance Measurement Handbook PDF Asset Page | Keysight
    Learn the basics of impedance measurements using Keysight LCR meters, impedance analyzers, and network analyzers. Explore impedance measurement techniques ...
  19. [19]
    Practical Implementation of a Novel Output Impedance ... - NIH
    In this technique two known resistive loads are sequentially attached to the output of the current source when applying a fixed current. The difference between ...
  20. [20]
    LTspice: AC Analysis Using The Step Command - Analog Devices
    Jul 4, 2016 · In LTspice, AC analysis involves computing the AC complex node voltages as function of frequency using an independent voltage or current ...
  21. [21]
    [PDF] Modeling the output impedance of an op amp for stability analysis
    Dec 17, 2015 · This allows for an open-loop measure- ment of the output impedance for all frequencies of interest. The CT capacitor is open at DC and then.
  22. [22]
    Amplifier Output Impedance (Damping Factor) and Speakers
    Jun 6, 2021 · First note damping factor, DF, is the load impedance divided by the amplifier's output impedance: DF = Zload / Zamp. The load is usually the ...A more exact understanding of amplifier output impedanceWhat is meant by damping factor and why is it importantMore results from www.audiosciencereview.comMissing: ringing | Show results with:ringing
  23. [23]
    Understanding S-parameters | Rohde & Schwarz
    This parameter indicates how much of the output signal is reflected due to impedance mismatch at the output. S parameters in a two port network.
  24. [24]
    [PDF] Understanding Basic Analog – Ideal Op Amps - Texas Instruments
    The output impedance of most op amps is a fraction of an ohm for low current flows, so this assumption is valid in most cases.Missing: Z_out | Show results with:Z_out
  25. [25]
    Negative Feedback Systems - Electronics Tutorials
    We see that the effect of the negative feedback is to reduce the gain by the factor of: 1 + βG. This factor is called the “feedback factor” or “amount of ...
  26. [26]
  27. [27]
    [PDF] Lecture 8: Output Stages and Power Amplifiers - Texas A&M University
    • The emitter follower output stage only has a peak efficiency of 25% as V ... • Emitter follower is a Class A power amplifier. • Display high linearity ...Missing: ABC | Show results with:ABC
  28. [28]
    [PDF] Fundamentals of Microelectronics Chapter 5 Bipolar Amplifiers
    Oct 20, 2010 · Output Impedance of Emitter Follower. ➢ Emitter follower lowers the source impedance by a factor of β+1 improved driving capability. E m s out.
  29. [29]
    Audio Myth - "Damping Factor Isn't Much of a Factor"
    ### Summary of Damping Factor and Related Effects
  30. [30]
    Audio Power Amplifier Design Guidelines - Elliott Sound Products
    Dec 27, 2006 · The output stage is not a simple impedance, and it varies as the load impedance changes. Bipolar transistors reflect the load impedance back to ...
  31. [31]
    [PDF] Agilent 33120A User's Guide - UNL EE Shop
    Select the output termination for output amplitude and offset voltage. The function generator has a fixed output impedance of 50 ohms on the. OUTPUT terminal.
  32. [32]
    RF Design Basics: VSWR, Return Loss, and Mismatch Loss
    Feb 5, 2023 · Learn about voltage standing wave ratio (VSWR), return loss, and mismatch loss, which helps characterize the wave reflections in a radio frequency (RF) design.
  33. [33]
    Q Factor and Bandwidth of a Resonant Circuit | Electronics Textbook
    The Q, or quality, factor of a resonant circuit is a measure of the “goodness” or quality of a resonant circuit.
  34. [34]
    N4690B Electronic Calibration Module (ECal), 300 kHz to 18 GHz ...
    Free deliveryThe Keysight N4690B MW electronic calibration (ECal) module makes calibration of vector network analyzers fast, easy and accurate.<|separator|>
  35. [35]
    [PDF] NisT Calibration Services Users Guide 1991 - GovInfo
    J. Pulse Waveform Measurements. 184. 65100S. Impulse Generator Spectrum Amplitude. (50 Ohm). 65200S. Fast Repetitive Broadband Pulse Parameters. (50 Ohm).
  36. [36]
    [PDF] AN47 - High Speed Amplifier Techniques - Analog Devices
    The power supply the amplifier sees at its terminals has high impedance at high frequency. This impedance forms a voltage divider with the amplifier and its ...
  37. [37]
    Arbitrary Function Generators - Tektronix
    The Tektronix AFG31000 Series is a high-performance AFG with built-in arbitrary waveform generation, real-time waveform monitoring, and the largest touchscreen ...Missing: terminations | Show results with:terminations
  38. [38]
    [PDF] Predicting battery capacity from impedance at varying temperature ...
    Dec 21, 2022 · SUMMARY. Prediction of battery health from electrochemical impedance spec- troscopy (EIS) data can enable rapid measurement of battery state.
  39. [39]
    [PDF] Discrete-Time Modelling of Li-ion Batteries with Electrochemical ...
    Jul 6, 2021 · The DNRC model characterizes ohmic, charge transfer, and diffusion overpotentials in the time domain with physically-meaningful circuit.
  40. [40]
    Expanding the Peukert equation for battery capacity modeling ...
    Aug 1, 2013 · The intent of this equation is to account for the intrinsic losses associated with discharging batteries at elevated currents. In particular, ...
  41. [41]
    Lithium ion battery degradation: what you need to know
    Mar 22, 2021 · Formation of the pSEI layer and accelerated growth of the negative SEI layer due to TM dissolution will cause an increase in the cell impedance.
  42. [42]
    How does Internal Resistance affect Performance? - Battery University
    ... mOhm for nickel-metal-hydride and a moderate 320 mOhm for lithium-ion. These internal resistance readings are typical of aging batteries with these chemistries.
  43. [43]
    Comparative capacity performance and electrochemical impedance ...
    The study being reported here compared a range of seven different AA alkaline brands in terms of price and performance by discharging them using three different ...<|separator|>
  44. [44]
    Pack Internal Resistance - Battery Design
    The other components are normally connected to the cells in series and typically add up to around 10mΩ to 30mΩ. Series and Parallel. When we look at the cells ...Missing: impedance per
  45. [45]
    Understand LDO Concepts to Achieve Optimal Designs
    At light load currents, typically less than 50 mA, the output impedance of the pass element is high. The LDO's output appears to be an ideal current source due ...
  46. [46]
    SMPS output impedance - Simulation, hardware & system design tools
    Output impedance is determined by Cout and Rload. Zout = sqrt(Rl*Rl + w*w*Cout*Cout). To increase impedance, a series inductor is needed, but it would be large ...
  47. [47]
    AC Capacitor Circuits | Reactance and Impedance—Capacitive
    Capacitive reactance is a capacitor's opposition to AC current, measured in ohms, and is inversely proportional to frequency. It is calculated as XC = 1/(2πfC).
  48. [48]
    Load Regulation | Power Supply terms - Matsusada Precision
    The evaluation formula is expressed as: Δ Vout / Δ Iload (%/A). where Δ ... Lower output impedance indicates superior load regulation performance. Power ...
  49. [49]
    [PDF] Easy Calculation Yields Load Transient Response - TI E2E
    The transient response to a load step can be calculated by multiplying the load step by the filter output impedance at the crossover frequency. In both voltage ...
  50. [50]
    Calculating the short-circuit current across the terminals of a ...
    Aug 21, 2015 · Calculation method for an alternator or a synchronous motor. Alternator 15 MVA; Voltage U = 10 kV; X'd = 20%.Missing: output | Show results with:output
  51. [51]
    [PDF] Remote sensing for power supplies - Texas Instruments
    Remote sensing compensates for resistive voltage drops in power paths, especially in low-voltage, high-current applications, using resistors to connect load ...