Fact-checked by Grok 2 weeks ago

Small-signal model

The small-signal model is a technique used in to analyze the behavior of nonlinear devices, such as transistors, under small alternating-current () signal variations superimposed on a direct-current ( point, enabling the prediction of circuit performance metrics like gain and impedance. This method linearizes the device's nonlinear current-voltage characteristics around the quiescent (Q-point) by considering only first-order perturbations, which simplifies analysis using standard linear circuit theory. The derivation of the small-signal model typically involves a expansion of the device's governing equations around the Q-point, retaining the constant () term and the linear () term while neglecting higher-order nonlinearities for small signals. For a (BJT) biased in the forward-active region, the exponential relationship i_C = I_S e^{v_{BE}/V_T} expands to yield the small-signal collector current i_c \approx g_m v_{be}, where g_m = I_C / V_T is the and V_T is the thermal voltage. This leads to equivalent circuits like the , which includes g_m, input resistance r_\pi = \beta / g_m, and output resistance r_o = V_A / I_C, where \beta is the current gain and V_A is the Early voltage. For metal-oxide-semiconductor field-effect transistors (MOSFETs) operating in , the small-signal model similarly uses a Taylor expansion of the drain current equation, resulting in i_d = g_m v_{gs} + g_o v_{ds} + g_{mb} v_{bs}, with g_m proportional to the square root of the drain current I_D and output conductance g_o accounting for channel-length . Key parameters include infinite gate input resistance due to the insulated gate and capacitances like C_{gs} and C_{gd} for high-frequency . These models are essential for designing amplifiers, where the signal amplitude must remain small (e.g., v_{be} \leq 5 mV for BJTs) to maintain . In practice, small-signal analysis proceeds in two steps: first, DC biasing to establish the Q-point by treating capacitors as open circuits and AC sources as grounds; second, AC analysis using the small-signal equivalent circuit with capacitors as short circuits and DC sources grounded. This approach is widely applied in analog integrated circuits, RF systems, and , providing insights into , , and noise performance without solving full nonlinear differential equations.

Fundamentals

Definition and Purpose

The small-signal model is a technique used in to analyze the behavior of nonlinear devices, such as diodes and transistors, under small (AC) variations superimposed on a steady (DC) bias point. This approach replaces the nonlinear characteristics of the device with an equivalent linear , consisting of resistors, capacitors, and controlled sources, valid only for signal amplitudes small enough that higher-order nonlinear effects are negligible. By focusing on incremental changes around the quiescent , the model simplifies the study of dynamics without requiring the full nonlinear simulation. The primary purpose of the small-signal model is to enable the application of established linear analysis methods—such as the , analysis, and frequency-domain techniques like Bode plots—to nonlinear circuits, particularly amplifiers and active filters. These tools allow engineers to efficiently compute critical performance metrics, including voltage or current gain, input and output impedances, , and , which are essential for designing reliable analog systems. Without this approximation, analyzing AC responses in nonlinear devices would involve computationally intensive nonlinear differential equations, making practical design infeasible. This modeling approach developed in the mid-20th century, initially for amplifiers in radio and communication systems, and later adapted for early circuits following the 1947 invention of the at Bell Laboratories, as supplanted vacuum tubes. Conceptually, the small-signal model is rooted in , specifically through first-order linearization of the device's nonlinear transfer functions around the DC , capturing only the dominant linear response to small perturbations while ignoring higher-order terms.

Linearization Technique

The technique for small-signal models involves approximating the behavior of nonlinear devices around a using a first-order expansion, enabling the analysis of small perturbations as linear responses. This method decomposes the total signal into a component and a small variation, where the nonlinear function describing the device characteristic is locally linearized to facilitate circuit analysis under superposition principles. Consider a nonlinear f(x) representing the device , such as current as a function of voltage. The expansion around the point x = X (the quiescent value) is given by f(x) = f(X) + f'(X)(x - X) + \frac{1}{2} f''(X)(x - X)^2 + \ higher-order\ terms. For small-signal analysis, higher-order terms are neglected, yielding the first-order approximation f(x) \approx f(X) + f'(X)(x - X). Here, f(X) is the component at the bias point, and the term f'(X)(x - X) captures the linear response to small deviations x - X. This approximation holds because the nonlinearity is locally linear over a sufficiently small range around X. In terms of variables, let the total be i(t) = I + i_s(t) and the total voltage be v(t) = V + v_s(t), where capitals denote values and lowercase with subscript s denote small-signal variations. The linearized relation simplifies to the general small-signal equation i_s(t) \approx g \cdot v_s(t), where g = f'(X) evaluated at the quiescent point represents the small-signal (for current-controlled devices) or conductance (for voltage-controlled devices). This parameter g quantifies the slope of the characteristic curve at the bias point and is constant within the small-signal regime. The validity of this linearization requires that the amplitude of the small-signal variation |v_s(t)| be much smaller than the characteristic scale of the nonlinearity, ensuring higher-order terms remain negligible. For semiconductor devices, this scale is typically the thermal voltage V_t \approx 26 mV at room temperature (300 K), derived from V_t = kT/q, where k is Boltzmann's constant, T is absolute temperature, and q is the electron charge; thus, small signals are generally limited to amplitudes below about 5-10 mV to maintain accuracy. Outside this regime, distortion from nonlinear terms becomes significant, invalidating the model. From this approximation, the small-signal equivalent is derived by replacing the nonlinear device with a linear : the is modeled as a constant independent source f(X), while the small-signal part g \cdot v_s(t) is represented as a linear dependent source or (with $1/g for two-terminal devices). For multi-terminal devices, the multivariate Taylor expansion extends this to a of conductances between terminals, forming a Thevenin- or Norton-equivalent linear that preserves the device's incremental behavior without the DC sources during analysis. This substitution allows standard linear techniques, such as or impedance calculations, to be applied directly.

Notation and Analysis

Variable Notation

In small-signal analysis of electronic circuits, standardized variable notation is essential to distinguish between constant (DC) bias components, time-varying (AC) small-signal components, and their superposition, facilitating clear separation of quiescent operating conditions from dynamic perturbations. This convention, widely adopted in , ensures consistency across analyses of devices like diodes and transistors, where nonlinear behavior is approximated linearly around a bias point. DC bias quantities, representing steady-state values at the quiescent , are denoted using uppercase letters, often with uppercase subscripts to specify terminals or nodes (e.g., V_{BE} for the base-emitter voltage of a or I_C for collector current). These symbols indicate time-independent constants, such as V_{BE} = 0.7 \, \text{V}, derived from the circuit's analysis. Small-signal AC quantities, which capture the incremental time-varying deviations from the bias point, are represented by lowercase letters with lowercase subscripts (e.g., v_{be}(t) for the time-domain base-emitter voltage or i_c(t) for collector variation). These are assumed to have zero time average, emphasizing their role in linear approximations where amplitudes are small enough to neglect higher-order nonlinearities (e.g., v_{be}(t) = 1 \, \text{mV} \cos(\omega t)). The total instantaneous quantity combines the and small-signal components additively, denoted by lowercase letters with uppercase subscripts (e.g., v_{BE}(t) = V_{BE} + v_{be}(t) or i_C(t) = I_C + i_c(t)). This decomposition allows the total signal to be expressed as the sum of a constant and a superimposed , preserving the full nonlinear description while enabling small-signal . In the , small-signal analysis employs notation to represent sinusoidal components as complex quantities, using lowercase symbols with the argument j\omega (e.g., v_{be}(j\omega) for the of the base-emitter voltage, where j = \sqrt{-1} and \omega is the ). This convention extends to network parameters, such as impedance Z(j\omega) and Y(j\omega), which characterize linear circuit elements under sinusoidal excitation by incorporating and in a single complex value. simplify differential equations into algebraic forms, with the time-domain signal recoverable via v_{be}(t) = \Re \{ v_{be}(j\omega) e^{j\omega t} \}, where \Re denotes the real part.

Quiescent Operating Point

The quiescent operating point, commonly referred to as the Q-point, represents the steady-state DC bias condition of an active device such as a , where operates without any applied signal. It is defined as the intersection between the DC load line—plotted on 's characteristic curve—and the nonlinear I-V relationship of itself, ensuring remains in its for linear . To determine the Q-point, two primary methods are employed. The graphical approach involves constructing the load line on the device's output characteristic curves by identifying the endpoints: one where the collector current I_C = 0 (corresponding to the supply voltage V_{CC}) and the other where the collector-emitter voltage V_{CE} = 0 (corresponding to the maximum current through the load R_C, given by I_{C(\max)} = V_{CC}/R_C); the intersection with the appropriate family of curves yields the Q-point coordinates (V_{CEQ}, I_{CQ}). Alternatively, the solves the system of nonlinear equations derived from Kirchhoff's voltage and current laws applied to the DC-equivalent circuit, where capacitors are treated as open circuits and inductors as short circuits. This often requires iterative techniques or equations like the Ebers-Moll relations for bipolar junction transistors (BJTs). Stability and sensitivity analysis of the Q-point are crucial due to potential shifts caused by temperature variations or device parameter fluctuations, such as changes in the current gain \beta or base-emitter voltage V_{BE}. Temperature increases can elevate leakage currents and \beta, leading to Q-point drift toward saturation or cutoff, which degrades amplifier performance; stability factors, like the factor S = \partial I_C / \partial I_{CO}, quantify this sensitivity, with lower values indicating better thermal runaway resistance through techniques such as emitter degeneration or feedback biasing. For instance, in a common-emitter biased with a network using resistors R_1 and R_2 across V_{CC}, the Q-point is selected at approximately V_{CEQ} = V_{CC}/2 and I_{CQ} = V_{CC}/(2R_C) to maximize undistorted output swing, allowing the AC signal to vary symmetrically around the without clipping, as verified by load line analysis showing the centered between saturation and cutoff.

Device Models

PN Junction Diodes

The small-signal model for a diode approximates the device's behavior around a quiescent for small perturbations in voltage and current, enabling linear circuit analysis. This model is derived from the diode's nonlinear current-voltage characteristics by applying techniques, such as expansion, at the point. The foundational nonlinear relationship governing the diode current I_D is the : I_D = I_S \left( e^{V_D / (n V_T)} - 1 \right) where I_S is the , V_D is the diode voltage, n is the ideality factor (typically between 1 and 2), and V_T = kT/q is the thermal voltage with k as Boltzmann's constant, T as absolute temperature, and q as the . This equation, derived from the physics of carrier diffusion and drift across the junction, captures the exponential forward-bias behavior and negligible reverse current. For small-signal analysis, the is biased at a quiescent point (V_Q, I_Q), and small variations v_d and i_d are superimposed. The small-signal conductance g_d, representing the incremental inverse, is the derivative of the diode current with respect to voltage evaluated at the quiescent point: g_d = \left. \frac{dI_D}{dV_D} \right|_{Q} = \frac{I_Q}{n V_T} This conductance models the as a linear r_d = 1/g_d for signals, with lower resistance at higher forward bias due to increased carrier injection. In addition to conductance, the small-signal model incorporates capacitances arising from charge storage effects. The diffusion capacitance C_d accounts for the storage and recombination of excess minority carriers in the neutral regions under forward bias: C_d = \tau g_d where \tau is the effective carrier lifetime, reflecting the time scale for carrier diffusion and recombination. This capacitance increases with forward bias and is dominant in forward-biased operation. The junction capacitance C_j, or depletion capacitance, arises from the variation of charge in the space-charge region and is significant in reverse or lightly forward bias: C_j = \frac{C_{j0}}{\sqrt{1 - V_R / V_{bi}}} where C_{j0} is the zero-bias junction capacitance, V_R is the reverse voltage (negative for forward bias), and V_{bi} is the built-in potential. For an abrupt junction, this follows a $1/\sqrt{V} dependence, modeling the widening/narrowing of the depletion layer with voltage. The complete small-signal equivalent circuit consists of the parallel combination of g_d, C_d, and C_j, representing the intrinsic junction behavior, all in series with the diode's bulk series resistance r_s to account for ohmic losses in the neutral regions and contacts. This circuit is valid for frequencies where the signal amplitude remains small compared to the thermal voltage, ensuring linearity.

Bipolar Junction Transistors

The small-signal hybrid-π model is a widely used for analyzing the linear behavior of bipolar junction transistors (BJTs) operating in the forward-active region under small perturbations around a quiescent . This model is obtained by linearizing the nonlinear Ebers-Moll large-signal model at the Q-point, where the base-emitter junction is forward-biased and the collector-base junction is reverse-biased. The topology consists of a voltage-controlled representing the , input and output resistances, and capacitances to capture frequency-dependent effects. It facilitates circuit analysis in common configurations like common-emitter amplifiers by replacing the nonlinear device with linear elements. Key parameters in the hybrid-π model include the transconductance g_m = \frac{I_C}{V_T}, where I_C is the quiescent collector current and V_T is the thermal voltage (approximately 26 mV at ). The input resistance at the base-emitter is r_\pi = \frac{\beta V_T}{I_C}, with \beta denoting the low-frequency current gain. The output is r_o = \frac{V_A}{I_C}, where V_A is the Early voltage accounting for base-width effects. The current gain relates as \beta = g_m r_\pi, directly derived from the linearized Ebers-Moll equations expressing collector current variations in terms of base-emitter voltage changes. For high-frequency operation, the model incorporates the base-emitter C_\pi, which includes and components, and the collector-base C_\mu, primarily the reverse-biased . These capacitances model charge storage and effects, limiting the ; for instance, C_\pi arises from the forward-biased base-emitter C_{de} = g_m \tau_F, where \tau_F is the forward transit time. In the common-emitter configuration, the small-signal voltage is approximately A_v \approx -g_m R_C, where R_C is the collector load , assuming high r_o and negligible capacitances at low frequencies. This expression the BJT's current-amplifying , with the negative indicating inversion.

Metal-Oxide-Semiconductor Field-Effect Transistors

The small-signal model for metal-oxide-semiconductor field-effect transistors (MOSFETs) is derived by linearizing the device's nonlinear characteristics around a quiescent , particularly in the region where typically occurs. In , the drain current I_D is approximated by the square-law expression I_D = \frac{1}{2} \mu C_{ox} \frac{W}{L} (V_{GS} - V_{TH})^2 (1 + \lambda V_{DS}), where \mu is the carrier mobility, C_{ox} is the per unit area, W/L is the , V_{GS} is the gate-source voltage, V_{TH} is the , V_{DS} is the -source voltage, and \lambda accounts for channel-length modulation. This model, originally proposed by Shichman and Hodges, forms the basis for small-signal analysis by enabling the computation of incremental parameters that describe the device's response to small perturbations. Key small-signal parameters include the transconductance g_m, which quantifies the change in drain current with respect to gate-source voltage, given by g_m = \frac{\partial I_D}{\partial V_{GS}} \big|_{Q} = \mu C_{ox} \frac{W}{L} (V_{GS} - V_{TH}), where the subscript Q denotes evaluation at the quiescent point. The output conductance g_{ds}, representing the change in drain current with drain-source voltage, is g_{ds} = \frac{\partial I_D}{\partial V_{DS}} \big|_{Q} \approx \lambda I_D, which introduces a finite output resistance r_o = 1/g_{ds} in the model. Additionally, the body effect parameter g_{mb} = \frac{\gamma g_m}{2 \sqrt{2\phi_F + V_{SB}}}, where \gamma is the body effect coefficient, \phi_F is the Fermi potential, and V_{SB} is the source-body voltage, captures the influence of the body (substrate) bias on the channel. For high-frequency operation, parasitic capacitances are included: the gate-source capacitance C_{gs} \approx \frac{2}{3} W L C_{ox} + W C_{ov} and the gate-drain capacitance C_{gd} \approx W C_{ov}, where C_{ov} is the gate overlap capacitance per unit width. The equivalent small-signal consists of a voltage-controlled g_m v_{gs} in parallel with g_{ds} between and , augmented by g_{mb} v_{bs} to account for effects, and the capacitances C_{gs} and C_{gd} connected between gate-source and gate-drain terminals, respectively. This emphasizes the MOSFET's voltage-controlled nature and high , making it suitable for analog amplification stages.

Comparisons and Applications

Differences from Large-Signal Models

Large-signal models describe the complete nonlinear behavior of devices across their full operating range, capturing effects such as , , and large voltage or swings during transient events. These models rely on the intrinsic nonlinear I-V characteristics of devices; for instance, diodes are typically represented by the Shockley equation, I_D = I_S \left( e^{V_D / (n V_T)} - 1 \right), where I_S is the reverse , n is the emission coefficient, V_D is the diode voltage, and V_T is the thermal voltage. Similarly, transistors use the Ebers-Moll equations to model forward and reverse s in active, , and regions, enabling accurate simulation of switching and amplification under varying biases. Such models are essential in circuit simulators like for transient analysis of nonlinear phenomena, including power dissipation and dynamic responses to large excursions. In contrast, small-signal models linearize device behavior around a quiescent , providing a valid only for incremental signals where the perturbation amplitude |v| is much smaller than characteristic device scales, such as the thermal voltage V_T \approx 25 mV at . Beyond this limit—typically voltage variations exceeding 10-25 mV—the linear assumption breaks down, leading to harmonic distortion and inaccurate predictions of or impedance as nonlinear effects like current dependence dominate. This restriction confines small-signal analysis to low-level , where is assumed, unlike large-signal models that handle the full nonlinear regime without such constraints. Computationally, small-signal analysis benefits from the of the , allowing efficient solution via matrix-based nodal or methods to compute functions, or Laplace transforms for frequency-domain responses including and . These algebraic techniques enable rapid hand calculations or simulations, often orders of magnitude faster than large-signal approaches. Large-signal simulations, however, require iterative (e.g., trapezoidal or Gear methods) to solve the coupled nonlinear differential equations over time, increasing complexity and runtime for transient behaviors in tools like . Hybrid approaches integrate both paradigms to leverage their strengths, particularly in distortion analysis for power amplifiers where large-signal models simulate overall envelope dynamics and efficiency, while small-signal linearizations at varying bias points evaluate intermodulation distortion and linearity metrics like IP3. This combination is common in RF design, using large-signal transient sweeps to identify operating loci, followed by small-signal assessments for , though it demands careful validation to ensure the linear approximation holds amid nonlinear .

Applications in Circuit Analysis

Small-signal models are essential for analyzing the performance of linear amplifiers, where they enable the calculation of key metrics such as voltage gain, current gain, input impedance, and output impedance. In transistor-based amplifiers, these models represent the device as a two-port network, facilitating the use of parameters like h-parameters (hybrid parameters) for common-emitter or common-source configurations, which relate input voltage and current to output counterparts under small-signal conditions. For instance, the forward current gain h_{21} approximates the low-frequency current gain of a bipolar junction transistor amplifier, while the input impedance h_{11} helps determine the loading on preceding stages. Similarly, y-parameters (short-circuit admittance parameters) are particularly useful for high-frequency amplifier analysis, as they incorporate admittances that account for capacitive effects at the ports. The of amplifiers is evaluated using small-signal models by incorporating parasitic s, which introduce poles and zeros that shape the versus frequency characteristic. Bode plots, constructed from the model's , visualize the magnitude and , revealing the low-frequency , midband flat region, and high-frequency roll-off due to capacitances like the base-emitter capacitance C_\pi in BJTs or gate-source capacitance in MOSFETs. A critical is the - frequency f_T, defined as the frequency where the short-circuit current extrapolates to , given by f_T = \frac{[g_m](/page/Transconductance)}{2\pi (C_\pi + C_\mu)}, where g_m is the and C_\mu is the base-collector capacitance; this parameter indicates the transistor's speed limit for and is vital for designing circuits. In multi-stage designs, small-signal models account for loading effects between stages, where the of one stage interacts with the of the next, potentially reducing overall gain; connections are analyzed by multiplying individual stage transfer functions while adjusting for these interactions. configurations, such as in operational amplifiers, use small-signal models to assess by examining phase margins and gain margins from the , ensuring oscillations are avoided. The , arising from the multiplication of an interstage by the gain of the inverting stage, significantly increases effective input capacitance and limits in multi-stage circuits, necessitating compensation techniques like Miller capacitors for . Practical analysis of small-signal circuits often combines hand calculations with tools, where manual methods using two-port parameters provide initial insights into and impedance. SPICE-based simulations linearize the circuit around the operating point and compute frequency-domain responses, enabling efficient evaluation of Bode plots and parameter sweeps without physical prototyping. In modern radio-frequency integrated circuits (RF ICs), small-signal models extend to (S-parameters), which describe power waves and are preferred for their compatibility with high-frequency measurements and matching networks, supporting designs up to millimeter-wave frequencies.

References

  1. [1]
    Lecture 10: Amplifiers - Small Signal Model | Circuits and Electronics
    Lecture 10: Amplifiers - Small Signal Model. Topics covered: Amplifiers - small signal model. Instructor: Prof. Anant Agarwal.
  2. [2]
    None
    ### Summary of Introduction to Small-Signal Analysis for BJT
  3. [3]
    None
    ### Definition and Key Aspects of Small-Signal Modeling for Linear Amplification
  4. [4]
    [PDF] Small signal model of MOSFET, MOSFET in digital circuits
    On Tuesday we talked about small signal models of MOSFETs. Why do we need small signal modeling? To linearize circuits. Linear circuits are much easier to work ...
  5. [5]
    Small-signal model - (Intro to Electrical Engineering) - Fiveable
    Definition. A small-signal model is a linear approximation of a nonlinear device's behavior, focusing on small fluctuations around a bias point.
  6. [6]
    Small-Signal Analysis - MATLAB & Simulink - MathWorks
    Small-signal analysis approximates the behavior of a nonlinear power electronics system, such as a switched-mode power supply, with a linear time-invariant ...
  7. [7]
    Understanding Small Signal Analysis in Electronics - AI-FutureSchool
    Sep 19, 2025 · Small signal models account for these effects, allowing for accurate predictions of gain, bandwidth, and stability. High-frequency analysis ...
  8. [8]
    Vacuum Tubes: The World Before Transistors - Engineering.com
    Jan 19, 2018 · In 1947, the trio of physicists William Shockley, Walter Brattain and John Bardeen created the world's first transistor, and marked the ...
  9. [9]
    [PDF] Diodes in Circuits - Harvey Mudd College
    This technique is called small signal modeling, and it relies on linearizing circuit elements using a Taylor series. ... small signal model circuits because ...
  10. [10]
    None
    ### Summary of Small-Signal Model as Taylor Series Expansion
  11. [11]
    None
    ### Summary of Notation for DC and Small-Signal Components
  12. [12]
    [PDF] Units & Symbols for Electrical & Electronic Engineers - IET
    Upper-case letters (e.g. V, I) are used for steady, mean and r.m.s values; lower-case letters for instantaneous values which vary with time (e.g. V, i).
  13. [13]
    [PDF] Basic BJT Amplifiers
    The conventional phasor notation for the small-signal base-emitter voltage is Vπ , called the control voltage. The dependent current source is then given by.
  14. [14]
    Transistor Biasing and the Biasing of Transistors - Electronics Tutorials
    This central operating point is called the “Quiescent Operating Point”, or Q-point for short. When a bipolar transistor is biased so that the Q-point is ...
  15. [15]
    What is DC Load Line Analysis and Its Significance - ElProCus
    The DC load line analysis main intention is to find the Quiescent Point (Q – point); The Quiescent Point obtain by the dc load line at which the parameters ...
  16. [16]
    Q Point or Load Line Analysis of JFET - Electrical4U
    Nov 17, 2018 · Q Point Definition: The Q point of a JFET is the intersection of the DC load line with the characteristic curve, crucial for stable operation.
  17. [17]
    [PDF] Bias Circuit Design for Microwave Amplifiers - UCSB ECE
    Sep 2, 2004 · Never fix IB: IC = β IB β varies tremendously from device to device and increases with temperature as well (0.7%/degree C). Field Effect ...
  18. [18]
    [PDF] A Comparison of Various Bipolar Transistor Biasing Circuits
    In addition, transistor parameters can vary over temperature causing a drift in IC at temperature. The low power supply voltages typically available for ...
  19. [19]
    [PDF] IV. Transistors (Biasing & Small-Signal Model)
    3) By definition amplifiers should be linear, i.e., they should not alter the shape of the signal. However, transistors are non-linear devices. We will discuss ...Missing: electronics | Show results with:electronics
  20. [20]
    [PDF] Recitation 16 Small Signal Model of pn Diode - MIT OpenCourseWare
    The small signal model of a p-n diode under forward bias includes junction capacitance (Cj) and diffusion capacitance (Cd), with linearized conductance and  ...Missing: derivation | Show results with:derivation
  21. [21]
    [PDF] Lecture 16 - MIT
    The pn Junction Diode (III). Outline. • Small-signal equivalent circuit model. • Carrier charge storage. –Diffusion capacitance. Reading Assignment: Howe and ...
  22. [22]
    [PDF] PN and Metal–Semiconductor Junctions
    SMALL-SIGNAL MODEL OF THE DIODE. 5. In a class of circuits called analog circuits, the diode is biased at a DC current, IDC, and the circuit behavior is ...
  23. [23]
    None
    ### Summary of BJT Small-Signal Model Derivation from Ebers-Moll
  24. [24]
    [PDF] • The dominant model used for small-signal analysis of a BJT in the ...
    This model is known as the hybrid-π model. It is similar to the h-parameter model used previously in this text, but has particular utility when frequency- ...
  25. [25]
  26. [26]
    [PDF] Lecture 11 - MOSFET (III) MOSFET Equivalent Circuit Models
    Oct 18, 2005 · Small-signal is small. ⇒ response of non-linear components becomes linear. • Can separate response of MOSFET to bias and small signal.
  27. [27]
    [PDF] Nonlinear Analysis of Diode Circuits To-do: Go through all 4 steps to ...
    Equation 1. Shockley's ideal diode equation. where is the diode current, is the saturation current (constant), is the voltage across the diode, and is the ...
  28. [28]
    [PDF] SPICE - (Simulation Program with Integrated Circuit Emphasis)
    Apr 12, 1973 · The BJT models are based on either the Ebers-. Moll [10] or the Gummel-Poon [17] formulations; the models for the FET '6 are derived from the ...
  29. [29]
    [PDF] 1. INTRODUCTION SPICE is a general-purpose circuit simulation ...
    SPICE is a general-purpose circuit simulation program for nonlinear dc, nonlinear transient, and linear ac analyses. Circuits may contain resistors, ...
  30. [30]
    [PDF] chapter 3 diodes
    3.52 The small-signal model is said to be valid for voltage variations of about 10 mV. To what percentage current change does this correspond (consider both ...<|separator|>
  31. [31]
    [PDF] Small Signal Circuits - MIT OpenCourseWare
    a circuit, we can replace large signal device models with corresponding small signal device models, and then analyze the resulting small signal circuit.
  32. [32]
    CDS 110/ChE 105: Lecture 7-1 Small Signal (Freq Domain) Analysis
    - Use exponential response (Laplace transforms) to model dynamics and control. - Use “loop shaping” (frequency domain design) for design: performance vs ...
  33. [33]
    [PDF] A STUDY OF NUMERICAL INTEGRATION TECHNIQUES FOR USE ...
    Apr 17, 1992 · This thesis concerns numerical intergration techniques used in the resistive cotnpanion circuit method for calculation of electrical transients.
  34. [34]
    [PDF] Fully Integrated CMOS Power Amplifier - UC Berkeley EECS
    Dec 6, 2006 · 54 It was found that small-signal stability analysis across swept operating points is a very good indicator of the stability under large signal ...
  35. [35]
    [PDF] A Linear RF Power Amplifier with High Efficiency for Wireless ...
    Feb 19, 2014 · The nonlinear operation of the power amplifier causes the generation of unwanted distortion products, which extend out of the intended channel ...
  36. [36]
    [PDF] Two-Port Network Parameters
    Figure EB.1 shows the small-signal equivalent circuit model of a transistor. Calculate the values of the h parameters. $.Missing: analysis | Show results with:analysis
  37. [37]
    [PDF] I. Frequency Response Review A. Phasor Analysis of the Low-Pass ...
    Small Signal Models for Frequency Response. A. Bipolar Transistor. B. MOS Transistor - V. SB. = 0. • Replace C gs for Cπ. • Replace C gd for Cµ. • Let rπ ---> ∞.
  38. [38]
    [PDF] Frequency Response Lesson #12 Small Signal Equivalent Circuits ...
    This capacitance is represented by capacitors in the small signal equivalent circuit for these devices. They are essentially open circuits in the low and mid.
  39. [39]
    [PDF] EE113 Course Notes Electronic Circuits
    3: DC ANALYSIS (LARGE-SIGNAL) ... Amplifiers always increase (or at least maintain) the signal power. The gain of an amplifier is expressed as a voltage ...
  40. [40]
    [PDF] High Gain Low Power Operational Amplifier Design and ...
    As a result, all multistage amplifiers suffer closed loop stability problems. Single Miller compensation is used for the simple two-stage amplifier; while the ...
  41. [41]
    [PDF] Design considerations for fast-settling operational amplifiers
    Sep 11, 1989 · optimum phase margin condition, we have extended the small-signal model to include resistance effects in addition to. Miller capacitance effects ...
  42. [42]
    [PDF] Small-Signal Modeling of RF CMOS - Stanford TCAD
    This dissertation explores the phys- ical origins of small-signal characteristics of RF MOSFETs and discusses many of the challenges facing the CMOS RF circuit ...<|control11|><|separator|>
  43. [43]
    A small-signal MOSFET model for radio frequency IC applications
    In this paper, we present a small-signal model for an integrated MOS transistor which takes into account the distributed nature of the gate structure.