Semiconductor Devices Theory and Application

1

Chapter Objectives and Semiconductor Fundamentals Overview

1.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

Semiconductors revolutionized electronics by replacing vacuum tubes with smaller, more reliable solid-state devices, and understanding their operation requires mastering both the atomic structure of doped silicon and a consistent naming convention for circuit analysis.

📌 Key points (3–5)

  • Historical shift: Semiconductors displaced vacuum tubes mid-20th century, enabling modern electronics from radios to smartphones.
  • What semiconductors are: Materials whose electrical properties lie between conductors and insulators, modified through doping to create P-type and N-type materials.
  • Core learning goals: Define semiconductors, describe atomic energy levels, explain silicon crystal structure and doping effects, and distinguish P-type from N-type materials.
  • Naming convention matters: Consistent notation (uppercase for DC/components, lowercase for AC/model parameters) prevents confusion when analyzing circuits.
  • Common confusion: Don't confuse external circuit components (e.g., R_E) with internal device model parameters (e.g., r_d)—subscript case distinguishes them.

📜 Historical context and scope

📜 The electronic age transition

  • 20th century shift: First half dominated by vacuum tubes (radio, TV, radar, long-distance telephone); mid-century saw solid-state semiconductors take over.
  • 1947 breakthrough: Bell Labs invented the first working transistor (point contact type), quickly replaced by the bipolar junction transistor.
  • Advantages over vacuum tubes: Semiconductors proved smaller, lighter, more reliable, and less expensive to manufacture.

🔬 Integrated circuits and modern applications

  • Evolution: Early integrated circuits contained dozens of transistors; today's devices contain billions.
  • Manufacturing note: ICs don't assemble individual transistors—they build all transistors simultaneously in layers, "rather like a layer cake."
  • Modern applications: Cell phones, GPS devices, laptops, tablets, and global communications infrastructure all depend on semiconductor technology.

🎯 Text scope and audience

"Any sufficiently advanced technology is indistinguishable from magic." —Arthur C. Clarke

  • Focus: Operation and application of semiconductor devices, not the design of semiconductors themselves.
  • Rationale: More people need to design, manufacture, and maintain devices using semiconductors than need to design the semiconductors.
  • Example: Many more people use cell phones than design them—this text targets the larger group who work with semiconductor devices.

🎓 Learning objectives breakdown

🎓 Core semiconductor concepts to master

After completing the chapter, students should be able to:

ObjectiveWhat it covers
Define semiconductorThe term itself and its meaning
Energy level differencesHow conductors, semiconductors, and insulators differ at the atomic level
Silicon structureAtomic structure of mono-crystalline silicon
Doping effectsHow doping changes silicon crystal properties
P vs N materialsDifferences between P-type and N-type materials
Energy diagramsDraw energy level diagrams for both P- and N-type materials

🧪 Material types and doping

  • Doping: The process of modifying a silicon crystal's electrical properties (details in later sections).
  • Two material types: P material and N material—understanding their differences is fundamental.
  • Visual representation: Students must learn to draw energy level diagrams for each type.

🔤 Variable naming convention

🔤 Why naming matters

  • Problem: Nomenclature often confuses beginning students in any subject.
  • Solution: Consistent naming convention throughout the text to minimize confusion.
  • Context: Circuits contain multiple passive and active components with various parameters and signals.

🔤 Component notation rules

Passive components:

  • R: Resistor (DC or actual circuit component)
  • r: Resistor (AC equivalent, where phase is 0 or ignored)
  • C: Capacitor
  • L: Inductor

Active components:

  • Q: Transistor (Bipolar or FET)
  • D: Diode

Electrical quantities:

  • V: Voltage (DC)
  • v: Voltage (AC)
  • I: Current (DC)
  • i: Current (AC)

🔤 Subscript conventions

Device-related subscripts (uppercase):

  • Differentiate components via subscript referring to the connected active device.
  • Example: R_E is a DC bias resistor connected to a transistor's emitter.
  • Example: r_C is the AC equivalent resistance at a transistor's collector.
  • Example: C_E is a capacitor connected to a transistor's emitter lead.

Model parameter subscripts (lowercase):

  • Exception to uppercase rule: If resistance or capacitance is part of the device model itself, use lowercase subscript.
  • Example: r_d is the AC dynamic resistance of a diode (internal model parameter, not external component).
  • Don't confuse: R_E (external emitter resistor) vs r_e (internal emitter resistance in the model).

Simple numbering:

  • When no active devices are present or multiple similar items exist: R_1, R_2, etc.
  • For particularly important components in complex circuits: specific names like R_source.

🔤 Voltage notation rules

Two-letter subscripts (node-to-node):

  • V_XY: DC potential from node X to node Y.
  • v_XY: AC signal across node X to node Y.

Single-letter subscripts (relative to ground):

  • V_X: DC potential from node X to ground.
  • v_X: AC signal at node X relative to ground.

Power supply exceptions:

  • Double-letter subscript: Indicates connection point.
  • Example: V_CC is the collector power supply.
  • Particularly important potentials: May receive special notation (text cuts off here, but implies exceptions for critical voltages).
2

Introduction to Semiconductor Devices

1.1 Introduction

🧭 Overview

🧠 One-sentence thesis

Semiconductor devices, particularly transistors and integrated circuits, have transformed modern technology by replacing vacuum tubes with smaller, lighter, more reliable, and less expensive components that now number in the billions per device.

📌 Key points (3–5)

  • Historical evolution: The transistor (invented mid-20th century) superseded vacuum tubes and enabled modern electronics; integrated circuits now contain billions of transistors built simultaneously.
  • Scope of this text: Focuses on the operation and application of semiconductor devices, not the design of semiconductors themselves—more people need to use and apply devices than design the underlying materials.
  • Atomic structure fundamentals: Atoms are mostly empty space; electrons occupy probability regions (orbitals) at discrete energy levels, not planet-like orbits; understanding electron energy levels is key to semiconductor behavior.
  • Common confusion: The popular planetary model of the atom is incorrect—electrons do not orbit in neat planar paths but occupy 3D probability contours; the Bohr model is an energy-level diagram, not a physical picture.
  • Variable naming convention: Uppercase letters (V, I, R) denote DC quantities or actual components; lowercase (v, i, r) denote AC equivalents or model parameters; subscripts indicate connection points or device associations.

📜 Historical context and device evolution

📜 From vacuum tubes to transistors

  • The point contact transistor was invented by John Bardeen, Walter Brattain, and William Shockley.
  • It was quickly superseded by the bipolar junction transistor (a major topic of the text).
  • Semiconductors proved superior to vacuum tubes in every practical dimension:
    • Smaller
    • Lighter
    • More reliable
    • Less expensive

🔬 The integrated circuit revolution

  • The last ~30 years of the 20th century saw rapid expansion of integrated circuits (ICs).
  • Early ICs contained the equivalent of a dozen or so individual devices; today they contain billions.
  • Manufacturing does not involve creating and connecting billions of discrete transistors; instead, all transistors are built simultaneously in layers (like a layer cake).
  • This extreme density enables common applications: cell phones, GPS devices, laptops, tablets, and global communications infrastructure.

🎯 Scope and audience of this text

The text focuses on the operation and application of semiconductor devices rather than the design of the semiconductors themselves.

  • Arthur C. Clarke observed: "Any sufficiently advanced technology is indistinguishable from magic."
  • Most citizens use numerous electronic devices daily without knowing how they work.
  • Key distinction: Many more people can use a cell phone than design one.
  • There is greater need for people who can design, manufacture, and maintain devices based on semiconductors than for people who design the semiconductors themselves.

🔤 Variable naming convention

🔤 Component and signal notation

The text establishes a consistent naming scheme to minimize confusion:

SymbolMeaningNotes
RResistor (DC or actual component)Uppercase for DC/actual
rResistor (AC equivalent, phase 0 or ignored)Lowercase for AC equivalent
CCapacitor
LInductor
QTransistor (Bipolar or FET)
DDiode
VVoltage (DC)Uppercase for DC
vVoltage (AC)Lowercase for AC
ICurrent (DC)Uppercase for DC
iCurrent (AC)Lowercase for AC

🔤 Subscript conventions

  • Device-related subscripts (always uppercase): e.g., R<sub>E</sub> is a DC bias resistor connected to a transistor's emitter; r<sub>C</sub> is the AC equivalent resistance at the collector; C<sub>E</sub> is a capacitor connected to the emitter.
  • Exception—model parameters (lowercase subscript): If the resistance or capacitance is part of the device model itself (not an external component), use lowercase. Example: r<sub>d</sub> is the AC dynamic resistance of a diode.
  • Simple numbering: If no active devices are present or several items exist, use R<sub>1</sub>, R<sub>2</sub>, etc.
  • Specific names: Particularly important components get descriptive names, e.g., R<sub>source</sub>.

🔤 Voltage and current subscripts

  • Two-letter subscripts indicate measurement nodes: V<sub>XY</sub> is the DC potential from node X to node Y; v<sub>XY</sub> is the AC signal across X to Y.
  • Single-letter subscript indicates potential relative to ground: V<sub>X</sub> is the DC voltage from node X to ground.
  • Exceptions:
    • Power supplies use double letters indicating the connection point: V<sub>CC</sub> is the collector power supply.
    • Directly named potentials: v<sub>in</sub> (AC input voltage), V<sub>R2</sub> (DC voltage across R<sub>2</sub>).
  • Currents generally use a single subscript referring to the measurement node: I<sub>X</sub> is the DC current flowing into or out of node X.
  • Mixed AC/DC equations: If an equation is valid for both AC and DC equivalent circuits, the uppercase form is preferred (consistent with directly coupled circuits that can amplify both AC and DC signals).

🔤 Why this matters

  • By following this scheme, you can always determine:
    • Whether the item is DC or AC equivalent
    • Its approximate circuit location
    • Other factors about it (e.g., external component vs. model parameter)

⚛️ Atomic structure fundamentals

⚛️ Why we need a model

  • Fundamental question: What is the internal structure of an atom?
  • It is nonsensical to ask what an atom "looks like" because its components are smaller than the shortest wavelengths of visible light.
  • Instead, we need a model to explain observed behavior.

⚛️ The planetary model (incorrect but popular)

The planetary model: nucleus at the center (protons and neutrons); electrons revolving in nice, regular, planar paths like planets around the sun.

  • This model is starkly incorrect.
  • It has been used as a symbol for nuclear regulatory agencies and a DEVO album cover from the 1970s, but it does not reflect reality.
  • Don't confuse: The planetary model's neat orbits with the actual behavior of electrons.

⚛️ Subatomic components and scale

  • Protons and neutrons: Similar masses, about 1.67×10<sup>−24</sup> grams each; most of an atom's mass comes from them.
  • Electrons: Mass roughly 2000 times smaller than a proton.
  • Scale: Proton radius ≈ 0.87×10<sup>−15</sup> meters; mean distance to nearest electron ≈ 5.3×10<sup>−11</sup> meters.
  • The electron is about 60,000 times farther from the proton than the proton's own size.
  • Analogy: This ratio is roughly the same as a golf ball to a sphere with a radius of 3/4 mile (1200 meters).
  • This ratio holds for other substances, including hard solids like diamond and quartz.

⚛️ The illusion of solidity

  • The vast majority of what we call "something" is really just empty space.
  • Example: You feel your buttocks pressed against a chair; both are considered solid, yet at the atomic level the vast majority of both is nothingness.
  • The feeling of solidity is just the result of the interaction of atomic forces between the two.

🌀 Electron orbitals and energy levels

🌀 The Heisenberg Uncertainty Principle and probability contours

  • Electrons do not whirl around the nucleus in stable, planet-like orbits.
  • First, the electron inhabits a region of 3D space, not simply a plane.
  • Second, due to the Heisenberg Uncertainty Principle, we cannot precisely plot the position and trajectory of a given electron.
  • The best we can do is make a plot of where the electron is likely to be: a probability contour.
  • Imagine recording the position of an electron thousands of times; plotting them all yields a cloud of dots around the nucleus.
  • This cloud is called an orbital (not "orbit"—they are different).

🌀 Shells, subshells, and orbitals

Orbitals indicate the electron energy level: a higher orbital implies a higher energy level.

  • Permissible electron energy levels are grouped into shells, then subshells, then orbitals.
  • Shells are denoted by their principal quantum number, n: 1, 2, 3, etc. Higher numbers can contain more subshells.
  • Subshells are organized by orbital shape and designated by letters: s, p, d, f.
    • Shell 1 contains only subshell s (1s).
    • Shell 2 contains subshells s and p (2s, 2p).
    • Shell 3 contains subshells s, p, and d (3s, 3p, 3d).
  • Subshells have variations within them (these are the orbitals):
    • One variation on s
    • Three variations on p
    • Five variations on d
    • Each orbital can hold a maximum of two electrons.

🌀 Maximum electron capacity

  • First shell: maximum 2 electrons (two in 1s).
  • Second shell: maximum 8 electrons (two in 2s, six in 2p—two in each of three p orbitals).
  • Third shell: maximum 18 electrons (two in 3s, six in 3p, ten in 3d—two in each of five d orbitals).
  • Formula: Maximum electrons in shell n = 2n².
  • Important: Orbitals fill from lowest energy level to highest energy level.

🌀 Orbital shapes and probability contours

  • 1s orbital (Figure 1.2): Spherical in shape; nucleus at the center. All s orbitals are spherically shaped (though internals change). 1s is the lowest energy orbital.
  • 2p orbitals (Figure 1.3): Three variations, one oriented along each of the X, Y, and Z axes. The nucleus is in the small void between two lobes. This is nothing like elliptical planetary orbits.
  • Higher orbitals can be very complex; combined contours can become intricate.
  • These graphics are cumbersome to work with, so a more functional model is needed.

🎨 The Bohr model (energy-level diagram)

🎨 What the Bohr model represents

The Bohr model (named after Niels Bohr) is an energy description of the atom, not an attempt to mimic its physical appearance or structure.

  • The nucleus is placed at the center.
  • Concentric rings represent the electron shells.
  • Higher ring number → larger ring → greater energy level.
  • Don't confuse: The Bohr model is not a picture of electrons orbiting in lanes; it is an energy-level depiction.

🎨 Energy transitions

  • If an electron moves from a higher level to a lower level, the energy difference is radiated out (as heat or light).
  • Example: This transition is what makes light emitting diodes (LEDs) function.
  • The inverse is also possible: by absorbing energy, an electron can move into a higher orbital.
  • These concepts are powerful for understanding semiconductor behavior.

🎨 Bohr model examples

  • Copper (atomic number 29): 29 protons, 29 electrons; electron shell configuration 2-8-18-1.
    • First three shells completely filled; single electron in the fourth shell.
    • This single outer electron is only loosely bound → copper is a very good conductor.
    • Bohr model: four rings, first three filled, one electron in the fourth ring.
  • Silicon (atomic number 14, Figure 1.5): Electron shell configuration 2-8-4.
    • Individual electrons drawn in each shell; atomic number indicated at nucleus.

🎨 Simplified Bohr model

  • Often useful to omit the filled inner shells (Figure 1.6).
  • Replace atomic number with the number of electrons in the outermost (valence) shell.
  • The valence shell is particularly important as it gives insight into the general behavior of the material.
  • Example: Silicon's simplified Bohr model shows +4 at the center (four valence electrons).

🎨 Energy level diagram (Figure 1.7)

  • Alternative representation: "straighten out" the Bohr model.
  • Show energy levels graphically as lines or bands, without counting specific electrons.
  • Vertical axis represents energy; horizontal lines represent shells (n=1, n=2, n=3, n=4, etc.).

Note: The excerpt ends mid-sentence ("We used silicon in the preceding example on purpose. The fact that it...") and transitions to a new section (1.3 Crystals) without substantive content in the provided text.

3

Atomic Structure

1.2 Atomic Structure

🧭 Overview

🧠 One-sentence thesis

Electron shells organize into energy bands in crystalline semiconductors, and the gap between valence and conduction bands determines whether electrons can move through the material.

📌 Key points (3–5)

  • Electron shell capacity: Each shell holds a maximum of 2n² electrons, distributed across s, p, and d subshells.
  • Bohr model purpose: The Bohr model is an energy-level diagram, not a physical picture of electron orbits.
  • Covalent bonding in crystals: Silicon atoms share valence electrons with four neighbors, forming a stable face-centered cubic structure.
  • Energy bands vs discrete levels: In crystals, individual atomic energy levels blur into continuous bands separated by forbidden gaps.
  • Common confusion: The Bohr model shows energy levels, not actual electron paths—don't imagine electrons orbiting in lanes like planets.

🔬 Electron shells and orbitals

🔢 Shell capacity formula

  • Each shell can hold a maximum of 2n² electrons, where n is the shell number.
  • Example: Shell 3 holds up to 2×3² = 18 electrons (two in 3s, six in 3p, ten in 3d).

🌐 Orbital shapes

Orbital: A region describing the probability of finding an electron at a given energy level.

  • 1s orbital: Spherical shape, lowest energy, nucleus at center.
  • 2p orbitals: Two-lobe shape with nucleus in the small void between lobes; three variations oriented along X, Y, and Z axes.
  • Higher orbitals become very complex; probability contours can be cumbersome to visualize.

Don't confuse: Orbitals are probability regions, not fixed paths—nothing like planetary orbits around the sun.

⚡ The Bohr model

🎯 What the Bohr model represents

Bohr model: An energy description of the atom, not a depiction of physical structure.

  • Nucleus at center, surrounded by concentric rings representing electron shells.
  • Higher ring number = larger radius = greater energy level.
  • Example: Copper (atomic number 29) has configuration 2-8-18-1; the Bohr model shows four rings, the first three filled and one electron in the fourth.

🔄 Energy transitions

  • Electron drops to lower level: Energy difference radiates out as heat or light.
    • Example: This transition makes light-emitting diodes (LEDs) function.
  • Electron absorbs energy: Moves into a higher orbital.

Key reminder: The Bohr model is an energy-level depiction—do not imagine electrons orbiting the nucleus in lanes.

🧩 Simplified Bohr model

  • Often omit filled inner shells.
  • Replace atomic number with the number of electrons in the valence shell (outermost shell).
  • Example: Silicon (atomic number 14, configuration 2-8-4) simplifies to show only the valence shell with four electrons.
  • Alternative: "Straighten out" the model into energy-level lines or bands without counting specific electrons.

🔗 Silicon crystals and covalent bonding

💎 Why silicon is special

  • Silicon has four electrons in its valence shell—half-filled.
  • As-is, it's neither a great conductor nor a superior insulator.
  • With proper structure, it becomes a semiconductor.

🤝 Covalent bonding in crystals

Covalent bond: Sharing of electrons "with or among the valence" to achieve stability.

  • Pure silicon can form a monocrystalline structure: all atoms align in a well-ordered pattern without voids.
  • Each silicon atom shares one electron from its four closest neighbors.
  • Result: Each atom effectively has eight electrons in its outer shell (four of its own + four shared), achieving stability.
  • All atoms are tightly bound to neighbors.

Don't confuse: The diagram is an energy representation, not a physical map—valence electrons are not zipping between atoms in figure-eight patterns.

🧊 Face-centered cubic structure

  • Represent each silicon atom as a ball, covalent bond as a connecting tube.
  • Each atom binds to four others in a regular, equal pattern.
  • Overall structure: essentially a cube with a silicon atom at the center of each face.
  • Therefore, the crystal structure is face-centered cubic.

📊 Energy bands and the Fermi level

📉 Discrete levels blur into bands

  • In a single atom: discrete, permissible energy steps.
  • In a crystal: each atom is affected by neighbors, causing slight energy-level changes.
  • Result: discrete levels blur into broader energy bands—a continuum of permissible electron energy levels.
  • Non-permissible zones between bands are called band gaps (forbidden regions).

⚖️ Fermi level

Fermi level: The energy level at which there is a 50% probability that it is filled with electrons.

  • Levels below the Fermi level tend to be filled; levels above tend to be empty.
Fermi level positionMaterial behavior
Within a bandGood conductor
Between widely separated bandsGood insulator
Between relatively close bandsSemiconductor

🔋 Intrinsic semiconductor energy diagram

Intrinsic semiconductor: A semiconductor with no impurities in the crystal (e.g., ideal silicon crystal).

  • Valence band: Lower energy band where electrons normally reside.
  • Conduction band: Higher energy band where electrons can "wander" through the crystal.
  • Band gap: Forbidden region between valence and conduction bands; the amount of energy needed to move an electron from valence to conduction band.
  • The Fermi level lies in the band gap.
  • Band gap value depends on the precise material used.

🌡️ Electron movement and hole flow

⚡ Thermal energy and electron movement

  • Without external energy (isolated at absolute zero): crystal lattice is stable, no electron movement.
  • Add thermal energy: valence electrons can jump to the conduction band.
  • In the conduction band, electrons can wander through the crystal.

🕳️ Holes and recombination

  • When an electron jumps to the conduction band, it leaves behind a hole (a place devoid of an electron).
  • The hole provides a place for another electron to "fall into."
  • Higher temperature → more freed electrons → more corresponding holes.
  • Result: thermally-induced electron movement (or equivalently, opposite-direction "hole flow").

Electron-hole recombination: When an electron moves into a hole, filling it.

Example: Imagine a horizontal bar with four dots (electrons) and one empty space (hole) on the left. When the leftmost electron moves into the hole, it fills the hole but creates a new hole where the electron was. Repeat this process: the hole appears to move from left to right, even though electrons are moving right to left.

Key insight: You can view the process as either electron movement in one direction or hole flow in the opposite direction—both perspectives describe the same phenomenon.

4

Crystals

1.3 Crystals

🧭 Overview

🧠 One-sentence thesis

In semiconductor crystals, discrete atomic energy levels blur into bands, and adding small amounts of impurities (doping) dramatically changes conductivity by creating an excess of either free electrons (N-type) or holes (P-type).

📌 Key points (3–5)

  • Crystal structure effect: atoms in a crystal affect each other, causing discrete energy levels to blur into continuous bands separated by forbidden gaps.
  • Intrinsic vs extrinsic: pure (intrinsic) semiconductors have equal numbers of thermally-produced electrons and holes and are not very useful; doped (extrinsic) semiconductors have added impurities that create excess charge carriers.
  • N-type vs P-type: pentavalent dopants (5 outer electrons) create N-type material with excess electrons as majority carriers; trivalent dopants (3 outer electrons) create P-type material with excess holes as majority carriers.
  • Common confusion: electron flow vs hole flow—electrons move in one direction (negative charge), holes move in the opposite direction (positive charge), but both represent charge movement.
  • Fermi level position: determines whether a material is a conductor (Fermi level within a band), insulator (between widely separated bands), or semiconductor (between relatively close bands).

🔬 Energy bands in crystals

🔬 From discrete levels to bands

  • In a single isolated atom, energy levels are discrete, permissible steps.
  • In a crystal, each atom is affected by surrounding atoms, causing slight changes in individual energy levels.
  • Taken as a whole, these variations blur discrete levels into broader bands.

Energy bands: continuous ranges of permissible electron energy levels in a crystal, replacing the discrete levels of isolated atoms.

  • Between bands remain non-permissible or forbidden zones.

Band gap: a forbidden energy zone between permissible bands.

⚡ Valence and conduction bands

  • Instead of thin discrete lines, the valence and conduction energy levels appear as thicker bands.
  • These bands still represent permissible electron energy levels, but now as a continuum rather than discrete steps.
  • The band gap represents the amount of energy needed to move an electron from the valence band to the conduction band.
  • The band gap value depends on various factors, particularly the precise semiconductor material used.

🎯 Fermi level

Fermi level: the energy level in a material at which there is a 50% probability that it is filled with electrons.

  • Levels below the Fermi level tend to be filled with electrons.
  • Levels above the Fermi level tend to be empty.

Material classification by Fermi level position:

Fermi level positionMaterial typeReason
Within a bandGood conductorElectrons can easily move within the partially-filled band
Between widely separated bandsGood insulatorLarge energy gap prevents electron movement
Between relatively close bandsSemiconductorModerate energy gap allows controlled conductivity

🧊 Intrinsic semiconductors

🧊 Crystal structure

  • Silicon has a face-centered cubic crystal structure.
  • At each corner of the cube and at the center of each face exists an atom of silicon.

Intrinsic semiconductor: a pure semiconductor crystal with no impurities, such as ideal silicon.

🌡️ Thermal electron generation

  • Without external energy (isolated at absolute zero), the crystal lattice is stable with no electron movement.
  • Adding thermal energy allows valence electrons to jump up to the conduction band.
  • Once in the conduction band, electrons can "wander" through the crystal.

🔄 Electron-hole pairs

How thermal generation creates both electrons and holes:

  • When thermal energy causes an electron to jump to the conduction band, it leaves behind a "hole" (a place devoid of an electron).
  • The hole provides a place for another electron to "fall into."
  • Higher temperature → greater number of freed electrons → greater number of corresponding holes.

Electron-hole recombination: the process when an electron moves into and fills a hole.

Visualizing hole flow:

  • Electrons move in one direction (e.g., left to right) to fill holes.
  • Holes appear to move in the opposite direction (e.g., right to left).
  • Electron movement = movement of negative charge.
  • Hole movement = movement of positive charge.

Charge carriers: electrons carry negative charge; holes carry positive charge.

Key characteristics of intrinsic semiconductors:

  • Number of thermally-produced electrons = number of holes (always equal).
  • Total number of both is quite small compared to total electrons in the crystal, even at room temperature.
  • Not particularly useful by themselves—neither good conductors nor insulators, and conduction is largely temperature-dependent.

🧪 Doped (extrinsic) semiconductors

🧪 What doping does

Dopants: foreign substances or impurities introduced into a semiconductor crystal.

Extrinsic semiconductor or doped material: a crystal with added dopants.

  • Doping alters the properties of the intrinsic material.
  • Amount of impurity is generally small, perhaps one part per million.

Doping methods:

MethodHow it works
Gaseous diffusionCrystal heated in an oven; dopant added in gaseous form; impurities "seep into" target crystal over time
Ion implantationImpurities accelerated and smashed into the target, dislodging and replacing original atoms

⚛️ N-type material (negative)

N-type material: semiconductor doped with pentavalent impurities (dopants with five electrons in the outer shell).

Examples of pentavalent impurities: phosphorus, arsenic, antimony.

How N-type material works:

  • A pentavalent impurity at the center of a silicon crystal has five outer electrons vs. silicon's four.
  • This creates an extra or donor electron.
  • The crystal has a net negative charge.
  • Donor electron energy level is just below the bottom of the conduction band.
  • The difference between donor level and conduction band is much smaller than the band gap itself.
  • Therefore, donor electrons easily jump into the conduction band, becoming free ionized electrons and leaving behind ionized holes.

Ion: an atom or molecule with unequal numbers of protons and electrons, resulting in a net charge. Cation = net positive charge (lost electrons); anion = net negative charge (gained electrons).

Charge carriers in N-type:

  • Number of free electrons is significantly larger than number of holes.
  • Electrons = majority charge carrier (or majority carrier).
  • Holes = minority charge carrier (or minority carrier).

Effect on conductivity:

  • Compared to undoped intrinsic crystal, doped N-type exhibits relatively high number of free electrons.
  • This enhances conductivity.
  • Greater doping level → greater conductivity enhancement.

Effect on Fermi level:

  • Extra electrons add to the number of filled energy states.
  • Being of higher energy than valence electrons, they push the Fermi level to a higher value.
  • Remember: Fermi level represents the point where 50% of states would be filled.

⚛️ P-type material (positive)

P-type material: semiconductor doped with trivalent impurities (dopants with three electrons in the outer shell).

Examples of trivalent impurities: boron, gallium, indium.

  • The excerpt introduces P-type material but does not provide detailed mechanisms.
  • By parallel with N-type: trivalent impurities have one fewer electron than silicon, creating excess holes (positive charge carriers).

Don't confuse:

  • N-type uses pentavalent (5 outer electrons) → excess electrons → negative.
  • P-type uses trivalent (3 outer electrons) → excess holes → positive.
5

Doped Materials

1.4 Doped Materials

🧭 Overview

🧠 One-sentence thesis

Adding small amounts of impurities (dopants) to intrinsic semiconductors creates extrinsic materials with enhanced conductivity and controllable electrical properties by introducing either surplus electrons (N-type) or surplus holes (P-type).

📌 Key points (3–5)

  • Why doping matters: intrinsic semiconductors alone are not particularly useful—they are neither good conductors nor insulators and their conduction depends heavily on temperature; doping alters their properties.
  • Two types of doped material: N-type (pentavalent dopants add extra electrons) and P-type (trivalent dopants create holes).
  • Majority vs minority carriers: in N-type, electrons are majority carriers and holes are minority; in P-type, holes are majority and electrons are minority.
  • Common confusion: the dopant amount is very small (around one part per million), yet it significantly changes conductivity and shifts the Fermi level.
  • Fermi level shift: N-type doping pushes the Fermi level up (closer to the conduction band); P-type doping pushes it down (closer to the valence band).

🔬 What doping is and why it's needed

🔬 Limitations of intrinsic semiconductors

Intrinsic semiconductor: a pure semiconductor crystal with no added impurities.

  • By themselves, intrinsic semiconductors are not particularly useful.
  • They are neither good conductors nor insulators.
  • Their conduction is largely dependent on temperature.
  • Even at room temperature, the total number of thermally produced electrons and holes is quite small compared to the number of electrons in the crystal.

🧪 What doping does

Dopants (impurities): foreign substances introduced into the crystal to alter its properties.

Extrinsic semiconductor (doped material): a crystal with an added dopant.

  • The amount of impurity added is generally small, perhaps in the neighborhood of one part per million.
  • Despite the small amount, doping significantly alters the material's electrical characteristics.
  • Don't confuse: "small amount" does not mean "small effect"—even trace dopants dramatically enhance conductivity.

⚙️ How dopants are added

Two main methods:

MethodHow it works
Gaseous diffusionCrystal is heated in an oven; dopant is added in gaseous form and diffuses ("seeps into") the target crystal over time
Ion implantationImpurities are accelerated and smash into the target, dislodging and replacing some of the original atoms in the crystal

⚡ N-type material: surplus electrons

⚡ How N-type is created

N-type material: semiconductor doped with pentavalent impurities (dopants with five electrons in the outer shell).

  • The "N" stands for Negative.
  • Examples of pentavalent impurities: phosphorus, arsenic, antimony.
  • Compared to an ordinary silicon atom (four electrons in outer shell), the pentavalent impurity creates an extra, or donor, electron.
  • The crystal has a net negative charge.

🔋 Donor electrons and energy levels

  • The energy level of the donor electrons is just below the bottom of the conduction band.
  • The difference between the donor level and the bottom of the conduction band is much, much smaller than the band gap itself.
  • Therefore it is relatively easy for these donor electrons to jump into the conduction band, becoming free ionized electrons and leaving behind ionized holes.

Ion: an atom or molecule that does not have a neutral net charge (the numbers of protons and electrons are not equal). If it loses electrons (net positive charge), it is called a cation; if it gains electrons (net negative charge), it is called an anion.

📊 Conductivity and charge carriers in N-type

  • Compared to the undoped intrinsic crystal, the doped extrinsic crystal exhibits a relatively high number of free electrons.
  • This enhances the conductivity of the material.
  • The greater the doping level, the greater the enhancement.

Majority charge carrier (majority carrier): the type of charge carrier that is significantly more numerous.

Minority charge carrier (minority carrier): the type of charge carrier that is less numerous.

  • In N-type material, electrons are the majority carrier (because the number of free electrons is significantly larger than the number of holes).
  • Holes are the minority carrier in N-type material.

📈 Fermi level shift in N-type

  • The extra electrons add to the number of filled energy states.
  • Being of higher energy than the valence electrons, they push the Fermi level to a higher value.
  • Remember: the Fermi level represents the point where 50% of states would be filled.
  • If we add states above this, then the new 50% point must be higher than the former level.
  • The donor level is very close to the conduction band.
  • The Fermi level has been pushed up, away from the valence band and closer to the conduction band.
  • This shift will be of great significance in upcoming discussions on semiconductor devices.

🔵 P-type material: surplus holes

🔵 How P-type is created

P-type material: semiconductor doped with trivalent impurities (dopants with three electrons in the outer shell).

  • The "P" stands for Positive.
  • Examples of trivalent impurities: boron, gallium, indium.
  • The trivalent impurity creates a hole—a location where an electron is lacking.
  • For this reason, trivalent impurities are sometimes called acceptors.

🔄 The reverse of N-type

  • The resulting situation is essentially the reverse of that of N-type material.
  • In P-type material, holes outnumber free electrons.
  • Consequently, holes are referred to as the majority carrier in P material.
  • Electrons take on the role of minority charge carrier.

📉 Fermi level shift in P-type

  • The Fermi level has been pushed down, closer to the valence band.
  • This is the opposite of N-type, where the Fermi level is pushed up.
  • The acceptor level is close to the valence band.

⚙️ Effect of doping level

  • As with N-type material, the greater the amount of trivalent impurity added, the greater the overall effect.
  • By itself, a doped crystal can be used to create a resistor.
  • The resistivity of the material is a function of the doping level.
  • By setting the cross-sectional area, length, and doping level, we can create well-defined resistor values.

🔮 Why combining N and P matters

🔮 Beyond simple resistors

  • If doped crystals could only be used to create resistors, the solid state semiconductor revolution would not exist.
  • The interesting bits arrive when we combine both N- and P-type materials into a single device.
  • This combination will be explored in the next chapter (PN Junctions and Diodes).

📋 Summary comparison

PropertyN-type materialP-type material
Dopant typePentavalent (5 outer electrons)Trivalent (3 outer electrons)
ExamplesPhosphorus, arsenic, antimonyBoron, gallium, indium
What's addedExtra (donor) electronsHoles (acceptors)
Net chargeNegativePositive
Majority carrierElectronsHoles
Minority carrierHolesElectrons
Fermi level shiftPushed up (closer to conduction band)Pushed down (closer to valence band)
Dopant energy levelJust below conduction band (donor level)Just above valence band (acceptor level)
6

PN Junctions and Diodes

Chapter 2: PN Junctions and Diodes

🧭 Overview

🧠 One-sentence thesis

The PN junction—formed by combining N-type and P-type semiconductor regions within a single crystal—is the fundamental building block of solid state devices including diodes, which serve diverse functions from rectification to light emission.

📌 Key points (3–5)

  • What a PN junction is: a single silicon crystal with adjacent N-type and P-type zones, not mechanically joined separate pieces.
  • How it's made: through diffusion or ion implantation techniques applied repeatedly to one crystal, creating different doped regions.
  • Why single-crystal matters: the junction must maintain mono-crystalline structure, not a poly-crystalline mix of separate materials.
  • Common confusion: PN junctions are not created by soldering, welding, or gluing two pieces together—the crystal remains one continuous piece.
  • What diodes do: the simplest PN junction device, used for rectifying, lighting (LEDs), and photodetection.

🔬 Doped semiconductor fundamentals

🔬 N-type material characteristics

N-type material: semiconductor doped with pentavalent (five-valence-electron) impurities, creating a surplus of electrons.

  • Pentavalent impurities are called donors because they donate extra electrons.
  • The Fermi level is pushed upward, away from the valence band and closer to the conduction band.
  • Majority carrier: electrons (abundant due to donor atoms).
  • Minority carrier: holes (fewer in number).
  • The donor energy level sits very close to the conduction band in the energy diagram.

🔬 P-type material characteristics

P-type material: semiconductor doped with trivalent (three-valence-electron) impurities, creating a surplus of holes.

  • Trivalent impurities are called acceptors because they accept electrons, leaving holes.
  • The Fermi level is pushed downward, closer to the valence band.
  • Majority carrier: holes (abundant due to acceptor atoms).
  • Minority carrier: electrons (fewer in number).
  • The acceptor energy level appears close to the valence band.

⚖️ Comparison of N-type vs P-type

PropertyN-typeP-type
Dopant typePentavalent (donors)Trivalent (acceptors)
Surplus charge carrierElectronsHoles
Fermi level movementPushed up toward conduction bandPushed down toward valence band
Majority carrierElectronsHoles
Minority carrierHolesElectrons

🎚️ Doping level effects

  • Greater impurity concentration → stronger effect on material properties.
  • Doping level, combined with cross-sectional area and length, determines resistivity.
  • A single doped crystal can function as a resistor with well-defined resistance values.
  • However, the real power of semiconductors emerges when N- and P-type materials are combined in one device.

🏗️ Creating the PN junction

🏗️ What makes a valid junction

PN junction: a region of N-type material adjacent to a region of P-type material within a single mono-crystalline silicon structure.

  • The junction is not formed by mechanical joining methods (soldering, welding, bolting, gluing, or friction-fitting).
  • The crystal must remain mono-crystalline (one continuous crystal structure), not poly-crystalline (multiple separate crystals stuck together).
  • This distinction is critical for proper device operation.

🛠️ Fabrication techniques

  • Diffusion or ion implantation techniques are applied repeatedly to a single piece of silicon.
  • These processes create distinct zones or regions within the same crystal.
  • One region becomes N-type, an adjacent region becomes P-type.
  • It's possible to embed one type completely within the opposite type (used in more complex devices).

🚫 Common fabrication misconception

Don't confuse: A PN junction is not two separate pieces of material attached together. The excerpt explicitly lists what is not done:

  • Not soldered
  • Not welded
  • Not bolted
  • Not friction-fitted
  • Not glued
  • Not duct-taped

The material remains a single continuous crystal throughout the doping process.

🔌 The diode and its applications

🔌 What a diode is

Diode: the most basic device built from a PN junction.

  • The PN junction is described as "arguably the fundamental building block of solid state semiconductor devices."
  • Diodes are the simplest application of this junction.

🔌 Types of diodes mentioned

The excerpt lists several diode varieties designed for different purposes:

Diode typePrimary use
Rectifier diodeRectifying (converting AC to DC)
Zener diodeVoltage regulation
LED (Light Emitting Diode)Lighting/illumination
PhotodiodePhotodetection (sensing light)
VaractorVoltage-variable capacitance

🔌 PN junctions in other devices

Beyond simple diodes, PN junctions appear in:

  • BJTs (Bipolar Junction Transistors)
  • JFETs (Junction Field Effect Transistors)

These more complex devices use multiple junctions or embedded regions.

7

PN Junctions and Diodes – Chapter Objectives

2.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

This chapter builds on extrinsic semiconductor knowledge to explain how combining P-type and N-type materials into a single crystal creates PN junctions—the fundamental building block of diodes and other semiconductor devices.

📌 Key points (3–5)

  • What a PN junction is: a single crystal with adjacent N-type and P-type zones, not mechanically joined pieces.
  • The depletion region: thermal energy causes electrons from N-material to recombine with holes in P-material, creating a carrier-free zone at the interface.
  • The energy hill concept: the depletion region creates a barrier that affects current flow through the device.
  • Common confusion: PN junctions are not made by physically attaching two separate pieces; they require a single mono-crystalline structure created through diffusion or ion implantation.
  • Why it matters: PN junctions form the basis of diodes (rectifiers, LEDs, photodiodes, Zeners) and transistors (BJTs, JFETs).

🏗️ Creating a PN junction

🏗️ Manufacturing requirement

  • The excerpt emphasizes that P- and N-type materials must not be joined mechanically (no soldering, welding, bolting, gluing, or tape).
  • Instead, a single piece of mono-crystalline silicon is treated repeatedly using diffusion or ion implantation techniques.
  • This creates distinct zones or regions within one continuous crystal—some N-type, some P-type.
  • One type can even be completely embedded within the opposite type (used in later device designs).

🔗 What defines the junction

PN junction: a single zone of N material adjacent to a zone of P material within one crystal.

  • This structure is described as "arguably the fundamental building block of solid state semiconductor devices."
  • Found in bipolar junction transistors (BJTs), junction field effect transistors (JFETs), and diodes.

⚡ The depletion region

⚡ How it forms

  • When N-material (surplus electrons) abuts P-material (surplus holes), thermal energy causes free electrons to "fall" into the excess holes.
  • This recombination happens at the interface between the two regions.
  • The result: a zone depleted of available charge carriers (neither free electrons nor holes remain mobile there).

🔬 What happens during recombination

In N-materialIn P-material
Electron leaves → positive ion remains (circled +)Electron arrives → negative ion forms (circled −)
Majority carrier (electron) is lostMajority carrier (hole) is filled
  • Example: an electron from the N-side recombines with a hole on the P-side, leaving behind fixed charged ions at the boundary.

🚧 The energy hill

  • The depletion region creates an energy hill that must be overcome to establish current flow.
  • This barrier arises because the interface is now devoid of mobile carriers.
  • The excerpt hints that understanding this hill requires recalling how doping shifts the Fermi level (N-type: Fermi level moves up toward conduction band; P-type: Fermi level moves down).

🔌 Diode applications and chapter scope

🔌 Devices built from PN junctions

The chapter will cover:

  • Rectifier diodes: for converting AC to DC.
  • Zener diodes: for voltage regulation.
  • LEDs (light-emitting diodes): for lighting.
  • Photodiodes: for light detection.
  • Varactor diodes: voltage-variable capacitors.

📐 Analysis tools introduced

  • Energy hill diagrams for PN junctions.
  • Forward-bias and reverse-bias operation regions (graphed).
  • Effective resistance under specific conditions.
  • Simplified circuit models for solving DC resistor-diode circuits.

🎯 Don't confuse

  • Intrinsic vs extrinsic crystals (from prior chapter): intrinsic = pure semiconductor; extrinsic = doped with impurities.
  • Mono-crystalline vs poly-crystalline: PN junctions require a single continuous crystal structure, not multiple pieces joined together.
8

PN Junctions and Diodes

2.1 Introduction

🧭 Overview

🧠 One-sentence thesis

The PN junction—formed by adjoining N-type and P-type regions in a single silicon crystal—creates a depletion region that acts as an energy barrier, allowing current to flow easily in one direction (forward-bias) but blocking it in the other (reverse-bias), making it the fundamental building block of semiconductor devices like diodes.

📌 Key points (3–5)

  • How a PN junction is made: not by mechanical joining, but by diffusion or ion implantation in a single mono-crystalline silicon piece, creating adjacent N and P zones.
  • What happens at the junction: free electrons from N material recombine with holes in P material, forming a depletion region devoid of charge carriers—an "energy hill."
  • Forward-bias vs reverse-bias: applying voltage with the correct polarity (positive to P, negative to N) flattens the energy hill and allows current flow; reversing polarity widens the depletion region and blocks current.
  • Common confusion: the barrier potential (forward voltage drop) is not zero—silicon requires ~0.7 V, germanium ~0.3 V, LEDs 1.5–3 V to overcome the energy hill.
  • Why it matters: this asymmetry makes the diode a one-way valve for current, the basis for rectifiers, LEDs, photodiodes, and transistors.

🏗️ Creating the PN junction

🏗️ Not a mechanical bond

A PN junction is formed by creating adjacent N-type and P-type regions within a single piece of mono-crystalline silicon, not by mechanically joining two separate pieces.

  • The excerpt emphasizes: no soldering, welding, bolting, gluing, or "duct tape."
  • Why mono-crystalline matters: a poly-crystalline amalgam (multiple crystals stuck together) would not work; the junction must be continuous.
  • How it's done: diffusion or ion implantation techniques are applied repeatedly to one silicon crystal, leaving N and P zones.
  • Example: one type can be completely embedded within the opposite type (used in later devices like transistors).

🧱 What the PN junction is

  • By placing an N zone adjacent to a P zone, you create the PN junction.
  • The excerpt calls it "arguably the fundamental building block of solid state semiconductor devices."
  • Found in: bipolar junction transistors (BJTs), junction field effect transistors (JFETs), and diodes.

⚡ The depletion region and energy hill

⚡ How the depletion region forms

  • At the interface between N and P materials, thermal energy causes free electrons (majority carriers in N) to "fall" into excess holes (majority carriers in P).
  • Result: electrons recombine with holes, leaving behind:
    • Positive ions in the N material (an electron left, so the atom is now positive).
    • Negative ions in the P material (an electron arrived, so the atom is now negative).
  • This interface region is depleted of available charge carriers (no free electrons or holes), hence the name depletion region.

🏔️ The energy hill concept

Energy hill: the barrier created by the depletion region that must be overcome for current to flow.

  • Why it forms: doping shifts the Fermi level—up for N material (toward conduction band), down for P material (toward valence band).
  • When N and P regions adjoin, the energy bands adjust so Fermi levels align.
  • Effect: the P material's bands rise relative to the N material's bands, creating a "hill" at the interface.
  • The depletion region is this hill; electrons must have enough energy to climb it.

🔌 Forward-bias operation

🔌 What forward-bias means

  • Connect the positive terminal of a voltage source to the P material and the negative terminal to the N material.
  • Electrons flow from the negative terminal → through N material → across the depletion region → into P material → to the positive terminal.

🔓 How forward-bias enables current flow

  • In N material, majority carriers are electrons, so they move easily.
  • The key: if the supplied voltage is high enough, electrons can diffuse into the P material (where there are many lower-energy holes).
  • The applied voltage "flattens" the energy hill.
  • Once the voltage equals or exceeds the barrier potential, current flows easily.

🚧 Barrier potential (forward voltage drop)

Barrier potential: the minimum voltage that must be dropped across the depletion region to achieve current flow.

MaterialTypical barrier potential
Silicon~0.7 V
Germanium~0.3 V
LEDs1.5–3 V (depends on color)
  • Don't confuse: this is not the total applied voltage; it is the voltage "used up" overcoming the depletion region.
  • Example: if you apply 1 V to a silicon diode in forward-bias, about 0.7 V is dropped across the junction, leaving 0.3 V for the rest of the circuit.

🚫 Reverse-bias operation

🚫 What reverse-bias means

  • Reverse the voltage source polarity: positive terminal to N material, negative terminal to P material.

🔒 How reverse-bias blocks current

  • Electrons in N material are drawn toward the positive terminal (away from the junction).
  • Holes in P material are drawn toward the negative terminal (away from the junction).
  • Result: the depletion region widens, increasing the size of the energy hill.
  • A small, short-lived current flows initially as carriers move away, then current ceases.
  • Further voltage increases only expand the depletion region more.
  • Ideal behavior: the PN junction acts like an open circuit under reverse-bias.

🔄 Asymmetry and the diode

  • Key insight: the PN junction allows current in one direction (forward-bias) but blocks it in the other (reverse-bias).
  • This asymmetry is the basis of the diode: a device that passes current easily one way but prevents flow the other way.
  • Example uses: rectifying AC to DC, light emission (LEDs), light detection (photodiodes).

📐 Quantifying behavior: the Shockley equation

📐 The Shockley equation

  • William Shockley derived an equation to describe PN junction current:
    • I = I_S × (e^(V_D × q / (n × k × T)) − 1)
  • Where:
    • I: diode current
    • I_S: reverse saturation current (a small leakage current in reverse-bias)
    • V_D: voltage across the diode
    • q: charge on an electron (1.6 × 10^−19 coulombs)
    • n, k, T: other parameters (excerpt cuts off here)
  • This equation captures the exponential relationship between applied voltage and current in forward-bias, and the near-zero current in reverse-bias.
9

The PN Junction

2.2 The PN Junction

🧭 Overview

🧠 One-sentence thesis

The PN junction creates an asymmetric device that allows current to flow easily in one direction (forward-bias) but blocks it in the opposite direction (reverse-bias), forming the basis of the diode.

📌 Key points (3–5)

  • Energy band alignment: When P and N materials meet, their Fermi levels align, creating a depletion region that acts as an energy "hill" at the interface.
  • Forward-bias behavior: Applying voltage with the positive terminal to P material flattens the energy hill; if voltage exceeds the barrier potential (~0.7 V for silicon), current flows easily.
  • Reverse-bias behavior: Reversing the voltage polarity widens the depletion region and blocks current flow, making the junction act like an open circuit (until breakdown).
  • Common confusion: Forward voltage drop vs applied voltage—the barrier potential (e.g., 0.7 V for silicon) must be overcome before significant current flows; below this threshold, current is negligible.
  • Asymmetry creates the diode: This one-way current behavior makes the PN junction the fundamental structure of a diode.

🏔️ Formation and structure of the PN junction

🏔️ Energy band alignment at the interface

Depletion region: The interface between P and N materials where energy bands adjust so Fermi levels align, creating an energy "hill."

  • When P material (Fermi level near valence band) adjoins N material (Fermi level near conduction band), the bands shift.
  • The P material's bands rise relative to the N material's bands.
  • This alignment creates a barrier at the junction that appears as a hill in the energy diagram.
  • Don't confuse: The depletion region is not simply empty space—it is a region where the energy bands have adjusted to align Fermi levels.

⚡ Physical charge distribution

  • In the depletion region, positive and negative charges are exposed near the junction.
  • The P side shows positive charges; the N side shows negative charges.
  • This charge distribution maintains the energy hill.

⏩ Forward-bias operation

⏩ How forward-bias works

Forward-bias: Connecting the positive terminal of a voltage source to the P material (anode) and the negative terminal to the N material (cathode).

  • Electrons flow from the negative terminal into the N material.
  • In N material, electrons are majority carriers and move easily.
  • If the applied voltage is high enough, electrons diffuse across the depletion region into the P material.
  • In P material, electrons fill lower-energy holes and migrate toward the positive terminal, completing the circuit.

🚧 Barrier potential (forward voltage drop)

Barrier potential: The minimum voltage required to overcome the depletion region and achieve current flow.

  • The applied voltage must "flatten" the inherent energy hill of the junction.
  • Material-dependent values:
    • Silicon: ~0.7 volts
    • Germanium: ~0.3 volts
    • LEDs: ~1.5 to 3 volts (varies with color)
  • Below the barrier potential, current is virtually non-existent.
  • Above it, current rises rapidly and becomes nearly vertical.
  • Example: A silicon diode with 0.5 V applied shows negligible current; at 0.7 V and above, current increases sharply.

⏪ Reverse-bias operation

⏪ How reverse-bias works

Reverse-bias: Connecting the positive terminal to the N material and the negative terminal to the P material.

  • Electrons in N material are drawn toward the positive terminal.
  • Holes in P material are drawn toward the negative terminal.
  • This creates a small, short-lived current that widens the depletion region.
  • Once the depletion region expands to match the supplied potential, current flow ceases.
  • Effect: The energy hill increases in size; the junction acts like an open circuit.

⚠️ Breakdown region

  • The Shockley equation does not model breakdown behavior.
  • At high reverse voltages (breakdown voltage V_R), the diode starts to conduct.
  • Two mechanisms:
MechanismConditionsHow it works
Zener effectHigh doping levels; breakdown below ~5–6 voltsVery high electric field across depletion region causes electron tunneling, producing high current
AvalancheLower doping levels; higher breakdown voltagesHigh electric field accelerates free electrons; they impact atoms, creating new electron-hole pairs that repeat the process, rapidly increasing current
  • General rule: Diodes should not be operated in breakdown (exception: Zener diodes designed for this purpose).

📐 Quantifying PN junction behavior

📐 The Shockley equation

The behavior of the PN junction is described by:

Shockley equation: I = I_S × (e^(V_D × q / (n × k × T)) − 1)

Where:

  • I = diode current
  • I_S = reverse saturation current
  • V_D = voltage across the diode
  • q = charge on an electron (1.6 × 10^−19 coulombs)
  • n = quality factor (typically 1 to 2)
  • k = Boltzmann constant (1.38 × 10^−23 joules/kelvin)
  • T = temperature in kelvin

Key observations:

  • At 300 kelvin, q/kT ≈ 38.6.
  • For even small forward voltages, the "−1" term can be ignored.
  • I_S is not constant; it approximately doubles for each 10°C rise in temperature.

📊 Characteristic curves

Linear scale (forward-bias):

  • Below ~0.5 volts: current is virtually non-existent.
  • Above ~0.5 volts: current rises rapidly.
  • After ~0.7 volts: curve becomes nearly vertical.
  • Higher temperature shifts the curve left (higher current for a given voltage).

Logarithmic scale (forward-bias):

  • Plotting current on a log scale produces a straight line.
  • This clearly shows the logarithmic relationship between voltage and current.

Complete I-V curve:

  • First quadrant (forward-bias): Shows the forward "knee" voltage V_F (~0.7 V for silicon) where current begins to rise sharply.
  • Third quadrant (reverse-bias): Shows reverse saturation current I_R (ideally zero, but a very small amount flows in reality).
  • Breakdown: At reverse voltage V_R, current increases rapidly.

🔌 The diode as a device

🔌 Basic diode structure and symbol

Diode: A semiconductor device that allows current to pass easily in one direction but prevents current flow in the opposite direction; in its basic form, just a PN junction.

Schematic symbol (ANSI standard):

  • Arrow points toward N material (cathode).
  • Arrow also points in the direction of easy conventional current flow.
  • P material = anode; N material = cathode.
  • General rule for semiconductor symbols: Arrows point toward N material.

Alternate symbol (IEC international standard):

  • Same structure but in outline form.

🎯 Why asymmetry matters

  • The PN junction's one-way behavior is extraordinarily useful.
  • Forward-bias: low resistance, current flows easily (above barrier potential).
  • Reverse-bias: high resistance, acts like an open circuit (until breakdown).
  • This asymmetry is the foundation for rectification, switching, and many other semiconductor applications.
10

2.3 Diode Data Sheet Interpretation

2.3 Diode Data Sheet Interpretation

🧭 Overview

🧠 One-sentence thesis

Reading a diode data sheet reveals critical operating limits and temperature-dependent behaviors that determine whether a device suits a given application.

📌 Key points (3–5)

  • Key parameters on data sheets: maximum reverse voltage, maximum forward current, switching speed, and power dissipation define safe operating boundaries.
  • Temperature effects are significant: reverse current doubles roughly every 10°C, and forward voltage decreases with rising temperature for a given current.
  • Physical packaging matters: different case styles (DO-35, DO-204, DO-4) handle different current and power levels; high-power devices mount to heat sinks.
  • Common confusion: the forward voltage is not fixed—it varies with both current and temperature, not just a constant 0.7 V.
  • Real-world example: the 1N4148 switching diode illustrates how specs (100 V reverse, 450 mA forward, 4 ns switching) match it to high-frequency signal applications.

📋 Understanding the characteristic curve parameters

📋 Forward knee voltage (V_F)

V_F is the forward "knee" voltage, roughly 0.7 volts for silicon.

  • This is the voltage where the diode begins to conduct significantly in the forward direction.
  • The excerpt shows this corresponds to the sharp upward turn in the I-V curve.
  • At room temperature, the 1N4148 data sheet confirms approximately 0.7 volts.

📋 Reverse saturation current (I_R)

I_R is the reverse saturation current, ideally zero but in reality a very small amount of current will flow.

  • Even when reverse-biased, a tiny leakage current exists.
  • This is not a failure—it's normal diode behavior.
  • Temperature strongly affects this parameter (see temperature section below).

📋 Reverse breakdown voltage (V_R)

V_R is the reverse breakdown voltage; current increases rapidly once this reverse voltage is reached.

  • Beyond this voltage, the diode conducts heavily in reverse.
  • Two mechanisms cause breakdown:
    • Zener effect: high doping levels, breakdown below ~5–6 volts, caused by electron tunneling through a strong electric field.
    • Avalanche: lower doping levels, electrons accelerate and create new electron-hole pairs, rapidly multiplying current.
  • General diodes should not operate in breakdown (except Zener diodes, which are designed for it).

🔌 Physical packaging and symbols

🔌 Schematic symbols

StandardDescriptionKey feature
ANSIFilled triangle with barPredominates in North America
IECOutline triangle with barInternational standard
  • The P material is the anode; the N material is the cathode.
  • General rule: arrows point toward N material.
  • The arrow also points in the direction of easy conventional current flow.
  • On physical devices, the cathode end is marked by a band (matching the bar on the symbol).

🔌 Common case styles

  • DO-35 and DO-204: small/medium current, through-hole mounting, size comparable to 1/4 to 1/8 watt resistor, model number stamped on body.
  • DO-4: stud or bolt style for higher currents and powers, facilitates mounting to metal plate or heat sink to dissipate excess heat.
  • Surface mount packages are also available.

Don't confuse: the physical size and mounting style directly relate to power handling—larger packages with heat sink provisions handle more power.

📊 Reading the 1N4148 data sheet

📊 Key specifications

The 1N4148 is a popular switching diode designed for high-speed operation in high-frequency signal applications and general-purpose use.

ParameterValueMeaning
Switching speed4 nanosecondsVery fast turn-on/turn-off
Maximum reverse voltage100 voltsSafe reverse bias limit
Maximum forward current450 milliampsContinuous forward current limit
Pulse currentUp to 4 ampsShort single pulses only
Power dissipation500 milliwattsMaximum heat the device can handle

Example: If your circuit applies 120 volts reverse bias, the 1N4148 is unsuitable (exceeds 100 V limit); if it needs 600 mA continuous forward current, the 1N4148 cannot handle it (exceeds 450 mA).

📊 Power derating and pulse limits

  • The data sheet shows power derating curves: as temperature rises, maximum safe power decreases.
  • Permissible pulse amplitudes are also shown: short pulses can carry much higher current than continuous operation.
  • This explains why 4 amp pulses are possible even though continuous rating is only 450 mA.

🌡️ Temperature effects

🌡️ Reverse current vs temperature

  • The data sheet graph shows reverse current variation with temperature.
  • The excerpt verifies the "doubles every 10°C" rule-of-thumb for reverse current.
  • Example: if reverse current is 10 nanoamps at 25°C, expect roughly 20 nanoamps at 35°C and 40 nanoamps at 45°C.

🌡️ Forward voltage vs temperature

  • The forward voltage curves show variation due to temperature.
  • Key behavior: for a given current, an increase in temperature results in a lower forward voltage.
  • Don't confuse: higher temperature does not mean higher forward voltage—it's the opposite.
  • At room temperature, the knee voltage is approximately 0.7 volts, but this shifts with temperature.

Example: If a diode carries 10 mA at 25°C with 0.7 V forward drop, at 75°C the same 10 mA might produce only 0.6 V forward drop.

🔍 Why data sheet interpretation matters

🔍 Matching device to application

  • The 1N4148's 4 nanosecond switching speed makes it suitable for high-frequency signals.
  • Its 100 V reverse limit and 450 mA forward limit define the circuit voltage and current boundaries.
  • Applications requiring higher current or power need different devices (e.g., DO-4 packaged diodes with heat sinks).

🔍 Avoiding damage

  • Operating beyond maximum reverse voltage causes breakdown (rapid current increase).
  • Exceeding maximum forward current or power dissipation damages the device.
  • Temperature derating must be considered: a device rated 500 mW at 25°C may only handle 300 mW at 75°C.
11

Diode Circuit Models

2.4 Diode Circuit Models

🧭 Overview

🧠 One-sentence thesis

Diodes are nonlinear devices that require simplified circuit models (first, second, and third approximations) to make circuit analysis practical, with each model trading simplicity for accuracy by adding elements like knee voltage and bulk resistance.

📌 Key points (3–5)

  • Why diodes need models: The Shockley equation is too complex for quick circuit analysis, and diodes are nonlinear and non-bilateral (unlike resistors), so superposition doesn't work without knowing bias state first.
  • Three approximation levels: First (simple switch), second (switch + knee voltage), third (switch + knee voltage + bulk resistance), each increasing in accuracy.
  • Common confusion: Don't confuse the model elements (like the 0.7V source or bulk resistor) with actual physical components inside the diode—these are behavioral models only.
  • Effective resistance varies: Diodes don't have a single "resistance"; DC resistance (voltage/current at operating point) differs from AC (dynamic) resistance (tangent slope at operating point), and both change with current level.
  • Bias state determines behavior: Always check if the diode is forward- or reverse-biased first; reverse-biased diodes act as open circuits (zero current), forward-biased ones follow the chosen approximation model.

🔧 Why diodes need special models

🚫 Nonlinearity blocks standard techniques

  • Diodes are not linear bilateral devices like resistors.
    • Linear means the I-V plot is a straight line; bilateral means forward and reverse quadrants are identical.
    • Diodes fail both tests: their characteristic curve is exponential, and forward/reverse behaviors are completely different.
  • Superposition doesn't work unless you already know the diode's bias state.
    • Example: A circuit with two voltage sources and a diode—one source might forward-bias it, the other reverse-bias it. The diode can't be both simultaneously, so you can't analyze each source separately and add results.

🧮 Shockley equation is impractical

  • The Shockley equation accurately describes diode behavior but adds too much computational complexity.
  • For speed and ease, engineers use simplified circuit element models instead.

📊 The three approximation models

🔌 First approximation: simple switch

The diode is modeled as a switch: closed (short circuit) when forward-biased, open when reverse-biased.

  • Simplest model: ignores all voltage drops and resistance.
  • When to use: Quick estimates where diode voltage drop is negligible compared to circuit voltages.
  • Example: In a 12V circuit with a 2kΩ resistor, treating the diode as a closed switch gives current = 12V / 2kΩ = 6 mA.

⚡ Second approximation: switch + knee voltage

Adds the "turn-on" potential (V_knee) required to overcome the energy barrier—0.7V for silicon diodes.

  • More realistic: accounts for the voltage drop across a conducting diode.
  • The knee voltage is the threshold where the diode begins significant conduction.
  • Example: Same 12V circuit, now current = (12V - 0.7V) / 2kΩ = 5.65 mA.
  • Most commonly used: provides sufficient accuracy for many applications without excessive complexity.

🎯 Third approximation: switch + knee voltage + bulk resistance

Adds a small resistive value (R_bulk) to model the non-vertical slope of the characteristic curve beyond the knee.

  • Most accurate: the I-V curve doesn't go perfectly vertical after the knee—voltage continues to increase slightly with current.
  • R_bulk represents this small positive slope.
  • Example: Same circuit with R_bulk = 10Ω gives current = (12V - 0.7V) / (2kΩ + 10Ω) = 5.622 mA.
  • Don't confuse: R_bulk is not "the diode resistance"—it's a minimum modeling value; actual effective resistance varies with operating conditions.

📈 Comparison table

ApproximationModel elementsAccuracyTypical use
FirstSwitch onlyLowestQuick estimates, high-voltage circuits
SecondSwitch + 0.7V sourceMediumMost practical applications
ThirdSwitch + 0.7V + R_bulkHighestPrecision requirements

Important reminder: These are behavioral models. There are no literal 0.7V batteries or little resistors inside the diode.

🔍 Effective resistance concepts

📐 DC resistance: operating point ratio

  • Definition: At a specific DC operating point (a particular current and voltage on the diode curve), DC resistance = voltage / current.
  • Graphically: the reciprocal of the slope of a line from the origin through the operating point.
  • Key insight: DC resistance changes with operating point—higher current produces lower DC resistance.
  • Example: If a diode operates at 0.8V and 0.005A, its DC resistance at that point is 0.8V / 0.005A = 160Ω. At a different current, the resistance would be different.

🌊 AC (dynamic) resistance: tangent slope

AC or dynamic resistance: the ratio of a small AC voltage variation to its associated AC current variation around the operating point.

  • Visualize: a small AC signal riding on top of the DC level, causing the operating point to move back and forth along the diode curve.
  • Graphically: the slope of a line tangent to the operating point (not from the origin).
  • Always smaller than DC resistance: the tangent line must be steeper than the line from the origin.
  • This is an average value across the AC variation.

⚠️ No single "diode resistance"

  • R_bulk models a minimum value but doesn't represent "the" diode resistance.
  • Resistance is a linear concept (straight line), but diodes are nonlinear.
  • We can only talk about effective resistance in a particular circuit under specific conditions (DC or AC).

🧪 Circuit analysis approach

✅ Step-by-step method

  1. Determine bias state first: Is the diode forward- or reverse-biased?
    • Forward: positive terminal of source connected to anode (through any resistors).
    • Reverse: positive terminal connected to cathode.
  2. If reverse-biased: Model as open switch, current = 0, diode voltage = whatever satisfies KVL (often equals source voltage).
  3. If forward-biased: Choose an approximation model and apply KVL.
    • Second approximation: I = (E - V_knee) / R_total
    • Third approximation: I = (E - V_knee) / (R_circuit + R_bulk)

🔄 Multiple diode circuits

  • Check each diode's bias state independently.
  • Forward-biased diodes contribute their knee voltage to the total voltage drop.
  • Reverse-biased diodes act as opens—virtually no current flows through them.
  • Example: Two diodes in series, both forward-biased, subtract 0.7V + 0.7V = 1.4V from the source before calculating current.

⚡ Parallel diode scenarios

  • If a diode is in parallel with a resistor and forward-biased, the resistor voltage equals the diode voltage (approximately 0.7V for second approximation).
  • The remaining source voltage drops across other series elements.
  • If the diode is reversed, it becomes an open and the circuit simplifies (often to a voltage divider).

🚨 Common pitfalls

  • Don't assume bias state: Always verify based on circuit topology and source polarity.
  • Don't forget multiple diodes: Each forward-biased diode adds its knee voltage to the total drop.
  • Reversed diode ≠ short: A reverse-biased diode is an open (unless in breakdown), not a short.
  • Breakdown exception: If reverse voltage exceeds breakdown rating, the diode conducts and voltage equals breakdown value (not covered in basic models).

💡 Practical examples summary

🧮 Single diode, forward-biased

  • Circuit: 12V source, 2kΩ resistor, silicon diode (forward).
  • First approximation: I = 12V / 2kΩ = 6 mA.
  • Second approximation: I = (12V - 0.7V) / 2kΩ = 5.65 mA.
  • Third approximation (R_bulk = 10Ω): I = (12V - 0.7V) / 2010Ω = 5.622 mA.
  • Observation: Second and third differ by less than 1% in this case; third predicts diode voltage slightly above 0.7V (≈0.756V) due to bulk resistance drop.

🔄 Single diode, reverse-biased

  • Circuit: 20V source with positive terminal to cathode, 2kΩ resistor.
  • Diode is reverse-biased → open switch.
  • Current = 0, resistor voltage = 0, diode voltage = 20V (to satisfy KVL).
  • Exception: If breakdown voltage < 20V, diode voltage = breakdown value, remainder drops across resistor.

🔗 Series diodes

  • Circuit: 9V source, two silicon diodes in series (both forward), R1 = 1kΩ, R2 = 2kΩ.
  • Both diodes forward-biased: I = (9V - 0.7V - 0.7V) / (1kΩ + 2kΩ) = 2.533 mA.
  • If either diode reversed: current = 0, all voltage drops across reversed diode.

🔀 Parallel diode and resistor

  • Circuit: 10V source, 1kΩ resistors R1 and R2, diode in parallel with R2.
  • Diode forward-biased: R2 voltage = diode voltage ≈ 0.7V, so R1 drops 9.3V, source current = 9.3 mA.
  • Diode reversed: becomes open, circuit is simple 1:1 divider, each resistor drops 5V.

🎛️ Mixed bias states

  • Circuit: 20V source, D1 forward-biased, D2 reverse-biased in parallel with 2kΩ resistor.
  • D2 acts as open, so I = (20V - 0.7V) / (1kΩ + 2kΩ) = 6.433 mA.
  • R1 drops 6.433V, R2 drops 12.867V (same as D2 voltage since they're parallel).
  • Virtually no current through D2 because it's reverse-biased.
12

2.5 Other Types of Diodes

2.5 Other Types of Diodes

🧭 Overview

🧠 One-sentence thesis

Diodes exploit different aspects of PN junctions beyond simple switching, enabling specialized functions such as voltage regulation (Zener), light emission (LED), light sensing (photodiode), fast switching (Schottky), and variable capacitance (varactor).

📌 Key points (3–5)

  • Zener diodes are used in reverse-bias mode to produce a stable voltage at their rated Zener voltage when breakdown occurs.
  • LEDs convert electrical energy to light with forward voltages typically 1.8–4 volts (depending on color), much higher than silicon's 0.7 V.
  • Photodiodes are the complement of LEDs: they generate current or voltage when exposed to light, operating in either photovoltaic (zero-bias) or photoconductive (reverse-bias) mode.
  • Common confusion: Zener analysis—first check if reverse voltage exceeds the Zener rating; if yes, replace the Zener mentally with a voltage source equal to its rated voltage.
  • Schottky and varactor diodes serve niche roles: Schottky offers fast switching and low forward drop (~0.2–0.3 V); varactor acts as an electrically controlled capacitance in reverse-bias.

🔋 Zener diodes: voltage regulation in reverse-bias

🔋 How Zener diodes work

Zener diode: a diode designed to operate in reverse-bias breakdown, producing a stable voltage at its rated "Zener voltage."

  • When forward-biased, a Zener behaves like an ordinary signal diode (0.7 V drop for silicon).
  • When reverse-biased below the Zener voltage, it acts as an open switch.
  • When reverse-biased at or above the Zener voltage, it enters breakdown (Zener or avalanche conduction) and maintains a stable voltage equal to the rated value.
  • Zener voltages are standardized (e.g., 3.9 V, 5.1 V, 6.8 V) and measured at a test current (I_ZT).
  • A lower current may not fully push the diode into conduction, resulting in lower-than-expected voltage.

🧮 Analysis method for Zener circuits

Step-by-step approach:

  1. Determine if the diode is forward-biased. If yes, treat it as a normal diode (0.7 V drop).
  2. If reverse-biased, treat it as an open switch initially.
  3. Calculate the voltage that would appear across the diode if it were open.
  4. If that voltage exceeds the Zener voltage, the diode is in Zener conduction—replace it mentally with a voltage source equal to the Zener voltage and recompute the circuit.

Example: In a circuit with a 9 V supply, a 5.1 V Zener (reverse-biased), and a 3.3 kΩ resistor in series:

  • If treated as open, the diode would see 9 V (greater than 5.1 V).
  • Therefore, the Zener is conducting; the diode voltage is 5.1 V.
  • By Kirchhoff's Voltage Law (KVL), the resistor drops 9 V − 5.1 V = 3.9 V.
  • Current: I = 3.9 V / 3.3 kΩ = 1.182 mA.

⚙️ Differential resistance and practical behavior

  • The breakdown curve is not infinitely steep; once past the rated Zener voltage, voltage increases slightly with current.
  • This effect is similar to bulk resistance (R_bulk) in forward-biased diodes.
  • On datasheets, this is called differential resistance (R_dif).
  • Application: Zeners placed in parallel with other components limit the voltage those components see to the Zener voltage, providing voltage regulation or protection.

Don't confuse: Zener voltage rating vs. actual voltage—the actual voltage depends on current; higher current yields slightly higher voltage due to differential resistance.

💡 Light emitting diodes (LEDs): electrical energy to light

💡 How LEDs produce light

LED (Light Emitting Diode): a diode that emits light when forward-biased, converting electrical energy into photons.

  • In a forward-biased PN junction, free electrons recombine with valence holes and release energy.
  • In most diodes, this energy is emitted as heat.
  • In LEDs, the energy transition is designed to radiate at shorter wavelengths (visible light or infrared/ultraviolet).
  • LEDs use exotic materials (not just silicon), which determines the color and forward voltage.

🌈 Forward voltage and color

ColorTypical forward voltage
Red / Amber~1.8–2.1 V
Yellow~2.1 V
Green / Blue~3.2–3.4 V
High brightness / UV~3–4 V
  • Forward voltage is noticeably higher than silicon's 0.7 V.
  • Reverse-biased, an LED acts like an open switch.
  • Maximum reverse voltage is relatively low (often just a few volts, e.g., 5 V).

🔆 LED specifications and characteristics

From the Cree C566D datasheet example:

  • Forward current: 50 mA (red/amber), 35 mA (green/blue); nominal operating current 10–30 mA.
  • Reverse voltage: 5 V typical.
  • Luminous intensity: measured in millicandela (mcd); varies by color.
  • Wavelength: specified in nanometers (nm); human vision covers ~400 nm (violet) to ~700 nm (red).
  • Peak/dominant wavelength: the wavelength producing the highest output in the LED's range (LEDs do not produce pure single-wavelength light like lasers).
  • Beam angle: how narrow or broad the illumination pattern is; narrower angles increase on-axis brightness.

Graphical data:

  • Luminous intensity increases roughly linearly with current.
  • Reverse voltage/current plots differ by color.
  • Beam pattern can be shown as a linear graph (relative brightness vs. angle) or a polar plot.

🧪 LED circuit analysis example

Circuit: 5 V supply, 2.1 V LED (forward-biased), 330 Ω resistor in series.

  • By KVL: resistor voltage = 5 V − 2.1 V = 2.9 V.
  • Current: I = 2.9 V / 330 Ω = 8.788 mA.
  • This should produce a relatively bright LED (likely amber or yellow given the 2.1 V drop).
  • The resistor "programs" brightness by controlling current; smaller resistance → higher current → brighter LED.

Don't confuse: Different colors have different forward voltages and conversion efficiencies; a change in current does not always produce a proportional change in brightness across colors.

🎨 Dual-LED application and color mixing

Example circuit: AC source driving two LEDs (one red, one blue) in opposite orientations with a shared current-limiting resistor.

  • For positive half-cycle: red LED is forward-biased (on), blue is reverse-biased (off).
  • For negative half-cycle: blue is on, red is off.
  • At low frequency (e.g., 1 Hz): LEDs blink alternately.
  • At higher frequency (~30 Hz or more): human vision integrates the rapid flashing, and both LEDs appear continuously lit (color mixing effect).
  • Application: This "on-off" trick is used in digital circuits to control LED brightness or motor speed; bi-color LEDs in a single package can achieve color mixing using a common lead and two control leads.

📷 Photodiodes: light sensing

📷 How photodiodes work

Photodiode: a diode that generates current or voltage when exposed to light, acting as the complement of an LED.

  • The photodiode includes a port allowing light to hit the junction.
  • A sufficiently energetic photon knocks loose an electron, creating an electron-hole pair and resulting in current flow.
  • More light energy → increasing current or voltage.
  • Can operate in visible, infrared (IR), or ultraviolet (UV) spectrum.

⚡ Two modes of operation

ModeBiasBehaviorAdvantagesDisadvantages
PhotovoltaicZero-bias (no external potential)Acts as a voltage sourceUsed by solar cellsSlower response
PhotoconductiveReverse-bias (external potential)Acts as a current sourceFaster responseHigher noise and dark current
  • Dark current: current produced even when no light is present; ideally zero. Large dark current reduces effective dynamic range.

Don't confuse: Photovoltaic mode (voltage source, no bias) vs. photoconductive mode (current source, reverse-biased).

⚡ Schottky and varactor diodes: specialized functions

⚡ Schottky diode: fast switching and low forward drop

Schottky diode: a diode using a semiconductor-to-metal contact (not semiconductor-to-semiconductor), offering very fast switching and low turn-on voltage.

  • Named after Walter Schottky, a German physicist.
  • Advantages:
    • Very fast switching times (orders of magnitude faster than traditional diodes).
    • Low turn-on voltage: ~0.2–0.3 V (vs. 0.6–0.7 V for silicon junction diodes).
  • Applications: shunting diodes in switch-mode power supplies, RF detector circuits.

🔧 Varactor diode: electrically controlled capacitance

Varactor diode: a diode used in reverse-bias mode as an electrically controlled capacitance.

  • The schematic symbol is a hybrid of a normal diode and capacitor.
  • Key insight: The depletion region acts as the dielectric of a capacitor, with anode and cathode as the plates.
  • All junction diodes exhibit some capacitance; varactors exploit this effect.
  • Increasing reverse-bias potential widens the depletion region (increases "plate separation"), decreasing capacitance.
  • Thus, DC bias voltage controls the capacitance value.

Characteristics:

  • Capacitance values are small: tens to hundreds of picofarads.
  • Sufficient for radio frequency (RF) work in oscillators and filters.
  • Advantages over mechanically adjustable capacitors: small size, high reliability, low cost, ability to rapidly change capacitance.

Don't confuse: Varactor capacitance decreases with increasing reverse-bias voltage (wider depletion region = larger plate spacing).

📋 Summary of diode types

Diode typePrimary useKey featureTypical bias mode
Switching/RectifyingGeneral switching, rectification0.7 V forward drop (silicon)Forward-bias
ZenerVoltage regulation, limitingStable voltage in breakdownReverse-bias
LEDLight emissionForward voltage 1.8–4 V (color-dependent)Forward-bias
PhotodiodeLight sensingGenerates current/voltage from lightZero-bias or reverse-bias
SchottkyFast switching, low drop~0.2–0.3 V forward drop, very fastForward-bias
VaractorTuning circuitsElectrically controlled capacitanceReverse-bias

Common confusion: All diodes share the same basic symbol structure—the "bar" portion represents the cathode for all types; variations in the symbol indicate the specialized function.

13

Diode Applications

Chapter 3: Diode Applications

🧭 Overview

🧠 One-sentence thesis

Diodes enable AC-to-DC conversion and signal shaping because their asymmetric conductance allows current to flow in only one direction, making rectification and other AC circuit applications possible.

📌 Key points (3–5)

  • What rectification does: converts alternating current (AC) waveforms into direct current (DC) waveforms by creating a signal with only a single polarity.
  • Why rectification matters: most electronic devices (TVs, computers) need fixed DC voltage internally, but power distribution is normally AC, so conversion is required.
  • Key property enabling these circuits: the diode's asymmetry—its sensitivity to the direction of current flow—makes AC applications possible.
  • Common confusion: DC does not mean constant value; pulsating DC varies in amplitude but never changes polarity, unlike a fixed battery voltage.
  • What this chapter covers: AC rectifier circuits, Zener diode regulators, complete AC-to-DC power supplies, and clippers/clampers for limiting and level shifting.

🔄 What rectification means

🔄 Core definition and purpose

Rectification: the process of turning an alternating current waveform into a direct current waveform, i.e., creating a new signal that has only a single polarity.

  • The term mirrors the everyday meaning: "to rectify the situation" means "to set something straight."
  • Rectification is crucial for modern electronics because it bridges the gap between how power is distributed (AC) and how circuits operate (DC).

🔌 Why conversion is necessary

  • The mismatch: residential and commercial power distribution is normally AC.
  • The requirement: electronic devices such as TVs or computers require a fixed, unchanging DC voltage to power their internal circuitry.
  • The solution: some form of AC-to-DC conversion is required, and diodes provide this through their asymmetric conductance.

⚡ Understanding DC vs pulsating DC

⚡ What DC really means

  • A DC voltage or current does not have to exhibit a constant value like a battery.
  • Key distinction: DC means the polarity of the signal never changes.
  • Don't confuse: a signal can vary in amplitude and still be DC, as long as it doesn't reverse polarity.

📈 Pulsating DC

  • When a DC signal varies in amplitude in a regular fashion (rather than staying fixed), it is sometimes called pulsating DC.
  • Example: the output of a rectifier may rise and fall, but it always stays positive (or always negative)—this is pulsating DC, not AC.
TypePolarity behaviorAmplitude behaviorExample
Fixed DCNever changesConstantBattery
Pulsating DCNever changesVaries regularlyRectifier output
ACAlternatesVariesResidential power

🚪 How diodes enable rectification

🚪 Asymmetric conductance

  • The diode's asymmetry—its sensitivity to the direction of current flow—is what makes AC circuit applications possible.
  • A diode conducts easily in one direction (forward bias) and blocks current in the opposite direction (reverse bias).
  • This one-way behavior allows the diode to "select" only one half of an AC waveform, converting it to a single-polarity signal.

🔧 Non-ideal effects

  • The excerpt notes that non-ideal effects such as a diode's forward voltage drop might be ignored in some instances but may be quite important in others.
  • When analyzing rectifier circuits, whether to account for the ~0.7 V drop (silicon) or ~0.3 V drop (germanium) depends on the application and required accuracy.

📚 Chapter scope and learning goals

📚 What the chapter covers

After the overview and introduction, the chapter focuses on:

  • AC rectifier circuits: solving for resulting waveforms.
  • Rectifier configurations: differences between half-wave, full-wave, and full-wave bridge.
  • Zener diode regulators: basic regulator circuits.
  • Complete power supplies: AC-to-DC supply with regulation, describing each component including the power transformer.
  • Clippers and clampers: AC circuits for limiting and level shifting.

🎯 Learning objectives

The chapter aims to enable students to:

  • Solve basic AC rectifier circuits for resulting waveforms.
  • Detail the differences between half-wave, full-wave, and full-wave bridge diode rectifier configurations.
  • Solve basic regulator circuits employing Zener diodes.
  • Outline a complete AC-to-DC power supply with regulation, describing the function of each component.
  • Solve AC clipper and clamper circuits for output waveforms.

🔗 Connection to previous chapter

  • The preceding chapter introduced practical considerations of diodes in DC circuits.
  • This chapter extends the discussion by focusing on AC circuit applications.
  • The transition from DC to AC analysis builds on the same diode properties but applies them to time-varying signals.
14

Diode Applications: Chapter Objectives

3.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

This chapter teaches how diodes convert AC to DC through rectification and enable regulation and signal-shaping circuits by exploiting the diode's asymmetric conductance.

📌 Key points (3–5)

  • Core application: AC-to-DC conversion (rectification) is essential because most electronics need DC power but distribution systems supply AC.
  • What makes it work: the diode's asymmetry—it conducts in one direction and blocks in the other—allows removal of one polarity from an AC waveform.
  • Key configurations: half-wave rectification (simplest), full-wave, and full-wave bridge rectifiers produce different DC outputs.
  • Beyond rectification: Zener diodes enable voltage regulation; clippers and clampers shape or shift AC waveforms.
  • Common confusion: DC does not mean constant voltage—pulsating DC varies in amplitude but never changes polarity, unlike a battery's fixed DC.

🔄 What rectification means

🔄 Definition and purpose

Rectification: the process of turning an alternating current waveform into a direct current waveform—creating a signal with only a single polarity.

  • The term mirrors everyday usage: "to rectify" means "to set something straight."
  • Why it matters: Most electronic devices (TVs, computers) require fixed DC voltage internally, but residential/commercial power is distributed as AC.
  • Some form of AC-to-DC conversion is therefore required in nearly all electronic equipment.

🔋 DC vs pulsating DC

  • DC voltage or current: polarity never changes (does not require constant amplitude).
  • Pulsating DC: varies in amplitude regularly but maintains the same polarity throughout.
  • Don't confuse: a battery provides fixed DC; a rectified AC signal is pulsating DC—both are DC because polarity never reverses.

🔌 Half-wave rectification

🔌 How it works

The chapter introduces the simplest rectifier: a series loop with an AC source, a diode, and a load resistor.

Operation cycle:

  • Positive half of input: diode is forward-biased → acts like a closed switch → all input voltage appears across the load resistor.
  • Negative half of input: diode is reverse-biased → acts like an open switch → circulating current drops to zero → no voltage across the resistor; all applied potential drops across the diode (by Kirchhoff's voltage law).

Example: If the AC source swings ±10 V, the load resistor sees only the positive 10 V peaks; the negative half is blocked.

📉 Result and limitations

  • Output: a pulsating DC waveform—only the positive portion of the input remains.
  • Why "half-wave": only half of the input waveform reaches the load.
  • Practical note: If the AC peak is small (e.g., 3–4 V) and a silicon diode is used, the diode's forward voltage drop (~0.7 V) noticeably reduces the peak output voltage.

🎯 Chapter learning objectives

The chapter prepares you to:

ObjectiveWhat you will do
Solve AC rectifier circuitsCalculate and sketch resulting waveforms for basic rectifier configurations
Compare rectifier typesDetail differences between half-wave, full-wave, and full-wave bridge configurations
Zener regulator circuitsSolve basic voltage regulator circuits using Zener diodes
Complete power supplyOutline an AC-to-DC supply with regulation, describing each component's function (including power transformer)
Clipper circuitsSolve AC clipper circuits for output waveforms (limiting/shaping)
Clamper circuitsSolve AC clamper circuits for output waveforms (level shifting)

🧩 Why these circuits are possible

  • The diode's inherent asymmetry in conductance—its sensitivity to current direction—is the foundation for all these applications.
  • Non-ideal effects (e.g., forward voltage drop) may be ignored in some cases but are critical in others (especially low-voltage circuits).

🔧 Broader context

🔧 Why AC distribution if DC is needed?

The excerpt addresses a natural question: why not distribute DC power directly?

Reasons for AC distribution:

  1. Efficiency: generally more efficient to distribute power via AC than DC.
  2. Voltage mismatch: even if DC were available, it may not be at the required amplitude → DC-to-DC conversion would still be needed, which can be more expensive than AC-to-DC conversion depending on the application.

🛠️ What comes next

  • The chapter extends diode discussion from DC circuits (previous chapter) to AC applications.
  • Focus areas: AC-to-DC conversion, regulation basics, and signal-shaping circuits (clippers and clampers).
  • All rely on the diode's directional conductance property.
15

MOSFET Small Signal Amplifiers – Introduction

3.1 Introduction

🧭 Overview

🧠 One-sentence thesis

MOSFETs can be configured as common source voltage amplifiers and common drain voltage followers (source followers), both offering very high input impedance.

📌 Key points (3–5)

  • Two main amplifier configurations: common source (voltage amplifier) and common drain (voltage follower/source follower).
  • Key advantage: both circuits offer the potential for very high input impedance.
  • Chapter scope: analyzing voltage gain, input impedance, and output impedance for these configurations.
  • AC modeling: the chapter will introduce a basic AC model for MOSFETs to enable small-signal analysis.

🔧 MOSFET amplifier configurations

🔧 Common source amplifier

  • Functions as a voltage amplifier.
  • One of the two main configurations discussed in this chapter.
  • The excerpt does not provide circuit details but establishes this as a primary topology.

🔧 Common drain amplifier (source follower)

  • Functions as a voltage follower.
  • Also called a "source follower" configuration.
  • The second main configuration covered in this chapter.
  • Example: similar to emitter follower in BJT circuits, but using MOSFET topology.

🎯 Analysis approach

🎯 AC modeling

  • The chapter will present a basic AC model for the MOSFET.
  • This model enables small-signal analysis of amplifier circuits.
  • The model is the foundation for calculating circuit parameters.

📊 Key parameters to analyze

The chapter focuses on three main performance metrics:

ParameterWhat it measures
Voltage gainHow much the circuit amplifies the input signal
Input impedanceThe impedance seen by the source driving the amplifier
Output impedanceThe impedance seen by the load connected to the output

⚡ High input impedance advantage

  • Both configurations offer "the potential for very high" input impedance.
  • This is a distinguishing characteristic of MOSFET amplifiers.
  • High input impedance means minimal loading of the signal source.
  • Don't confuse: this is a feature of both the common source and common drain configurations, not just one.
16

Rectification

3.2 Rectification

🧭 Overview

🧠 One-sentence thesis

Rectification converts AC signals into pulsating DC by using diodes to block or flip portions of the waveform, and adding a filter capacitor smooths the output into a more constant voltage suitable for DC applications.

📌 Key points (3–5)

  • Half-wave rectification: uses one diode to pass only the positive (or negative) half of the AC waveform, discarding the other half; inefficient but simple.
  • Full-wave rectification: uses multiple diodes to invert the negative half instead of discarding it, doubling efficiency and reducing ripple.
  • Filtering with capacitors: a parallel capacitor charges during conduction and discharges during the gap, smoothing the pulsating DC; larger capacitors reduce ripple but cause higher current spikes.
  • Common confusion: larger filter capacitors produce smoother output but charge for a shorter time with higher peak currents, not longer charge times.
  • Practical considerations: transformers scale AC voltage to usable levels; the diode forward drop (≈0.7 V for silicon) reduces peak output voltage, especially noticeable with low input voltages.

🔌 Half-wave rectification basics

🔌 How the circuit works

  • A simple series loop: AC sine wave source → diode → load resistor.
  • Positive half-cycle: diode is forward-biased (acts like a closed switch); nearly all input voltage appears across the resistor.
  • Negative half-cycle: diode is reverse-biased (acts like an open switch); current drops to zero, no voltage across the resistor; all applied potential drops across the diode (by Kirchhoff's voltage law).
  • Result: the load sees only the positive pulses—a pulsating DC waveform.

Half-wave rectification: a process that passes only one half of the AC waveform to the load, leaving just positive (or negative) pulses.

⚠️ Effect of diode forward drop

  • If the AC peak input is small (e.g., 3–4 volts) and a silicon diode is used, the 0.7 V forward drop is a large percentage of the peak.
  • The positive pulses are slightly narrowed because current does not flow at reasonable levels until the input reaches ≈0.6–0.7 V.
  • Example: with a 10 V peak source, the load voltage peaks just under 9 V (≈1 V lost to the diode drop).
  • Don't confuse: the diode drop is always present, but it is more noticeable as a percentage when the input voltage is low.

🔄 Reversing the diode

  • If the diode is oriented in reverse, it blocks the positive portion and allows only the negative portion through.
  • The load waveform appears flipped top-to-bottom compared to the standard half-wave output.

🧰 Practical AC-to-DC conversion issues

🧰 Voltage scaling with transformers

  • Standard AC outlets provide 120 V RMS, which is often too high (or sometimes too low) for circuitry.
  • A transformer has a primary (input) and secondary (output) side, each with a coil of wire wound around a common magnetic core.
  • The voltage ratio equals the turns ratio: if the secondary has half as many turns as the primary, the secondary voltage is half the primary voltage and the current is doubled.
  • Ideally, no power is lost (power in = power out); in reality, transformers have VA (volt-amp) ratings specifying maximum secondary voltage × current.
Transformer typeEffect
Step-downDecreases voltage (e.g., 120 V → 12 V)
Step-upIncreases voltage (e.g., 12 V → 120 V)
  • Transformers can have multiple primaries/secondaries or tapped coils for flexibility.

🔋 Why AC distribution is used

  • It is generally more efficient to distribute power via AC rather than DC.
  • Even if DC is available, it may not be at the required amplitude; DC-to-DC conversion can be more expensive than AC-to-DC conversion depending on the application.

🧊 Smoothing (filtering) the pulsating DC

🧊 Adding a filter capacitor

  • A capacitor in parallel with the load charges during the diode conduction phase (storing energy) and discharges when the diode turns off (transferring energy to the load).
  • Larger capacitors store more energy and produce smoother load voltage.
  • The capacitor and load resistance form an RC discharge network; for smooth output, the time constant τ = RC should be much longer than the gap when the diode is off.
  • Example: for 60 Hz operation, the gap is half the period ≈8.3 ms; a time constant several times larger (e.g., 100 ms) is needed.

📉 Ripple voltage

Ripple: the variation in output voltage due to capacitor discharge; it can be modeled as an AC voltage riding on a larger DC output.

  • Ripple worsens as load current increases.
  • Under light load, the output floats near the peak secondary voltage with very little ripple.
  • As load current demand increases, ripple magnitude increases and the nominal output voltage drops.

⚡ Capacitor size trade-offs

  • Larger capacitor (e.g., 1000 μF vs. 50 μF):
    • Longer RC time constant → smoother output, less ripple.
    • Counterintuitive effect: the diode turns on for a shorter time because the capacitor holds the cathode at a high voltage; the diode only conducts when the input exceeds the capacitor voltage by ≈0.7 V.
    • This leads to very large current spikes during the brief charging phase.
  • Example (from simulation):
    • 50 μF capacitor: charging current peaks ≈180 mA over ≈4 ms; discharge current ≈−80 mA.
    • 1000 μF capacitor: charging current peaks ≈800 mA (over 4× higher) over ≈2.5 ms; discharge phase nearly flat (stable output).
  • Don't confuse: larger capacitors do not charge for longer—they charge faster with higher peak currents because they are replenished only when the input voltage exceeds the already-high capacitor voltage.

🔁 Full-wave rectification

🔁 Why full-wave is better

Full-wave rectification: a process that uses both halves of the AC waveform by inverting (flipping) the negative portion instead of discarding it.

  • Half-wave rectification is inefficient because it throws away the negative half.
  • Full-wave rectification doubles efficiency and greatly reduces ripple, allowing much smaller filter capacitors.
  • The circuit is modestly larger and more complicated but offers large performance improvements.

🔧 Two methods for full-wave rectification

🔧 Method 1: Center-tapped secondary with two diodes

  • Uses a transformer with a center-tapped (split) secondary and two diodes.
  • Positive half-cycle: diode D₁ is forward-biased, D₂ is reverse-biased; current flows through D₁ and the upper half of the secondary into the load.
  • Negative half-cycle: D₁ is reverse-biased, D₂ is forward-biased; current flows through D₂ and the lower half of the secondary into the load.
  • Both halves of the input waveform are used.
  • Important: the load only "sees" half of the total secondary voltage at any time (minus one diode drop).
  • Example: a transformer with a 12 V RMS center-tapped secondary delivers ≈6 V RMS (≈8.5 V peak) to the load, ignoring the diode drop.

🔧 Method 2: Four-diode bridge rectifier

  • Uses a four-diode bridge network.
  • Can produce a bipolar output (both positive and negative outputs, typically of the same magnitude).
  • (The excerpt ends before describing the bridge operation in detail.)
MethodDiodesTransformerLoad voltageNotes
Center-tapped2Center-tapped secondary requiredHalf of total secondary (minus one diode drop)Simpler diode count
Bridge4Standard secondaryFull secondary (minus two diode drops)Can produce bipolar output
17

Clippers

3.3 Clippers

🧭 Overview

🧠 One-sentence thesis

Clippers are circuits that limit signal amplitude at programmable voltage levels using diodes and DC bias sources, enabling both signal protection and intentional waveform shaping.

📌 Key points (3–5)

  • Purpose of clipping: limits maximum signal amplitude for protection or intentional wave shaping (e.g., guitar distortion effects).
  • Basic mechanism: a diode in parallel with the load limits voltage swing to its forward turn-on potential (~0.7V for silicon).
  • Biased clippers: adding a DC bias source allows programming the clip point to any desired voltage level (not just 0.7V).
  • Common confusion: the resistor R value is not critical—it simply needs to be much larger than the diode's on-resistance to create an effective voltage divider.
  • Dual-polarity capability: combining positive and negative clippers allows independent limiting of positive and negative signal swings.

🎸 Why clipping matters

🎸 Protection and wave shaping

  • Protection use case: prevents excessive input signals from damaging downstream circuits by limiting amplitude.
  • Aesthetic use case: guitar amplifier "fuzz" sound originally came from power stage clipping when overdriven.
    • Problem: achieving this sound required maximum volume (unpopular with neighbors).
    • Solution: clip the signal earlier in the chain at lower volume levels using dedicated clipper circuits.
  • This led to commercial "fuzz boxes" and "distortion pedals" in the 1970s.

⚡ Basic clipper operation

⚡ Simple parallel diode clipper

  • Simplest form: place a diode (or two opposing diodes) in parallel with the load.
  • The diode limits output voltage swing to its forward turn-on potential (~0.7V for silicon).
  • Limitation: you're stuck with 0.7V as the clip point.
  • While stacking diodes in series can increase the limit, this approach is inflexible.

🔧 Biased diode clippers

🔧 Positive clipper design

A biased diode clipper uses a DC source to program the clip point to any desired voltage level.

Circuit structure (Figure 3.23):

  • Diode and DC bias voltage V_clip in the signal path.
  • Series resistor R between input and output.

How it works:

  • When V_in < V_clip: diode is reverse-biased → high resistance → effectively removed → signal passes through R unimpeded.
  • When V_in > V_clip + 0.7V: diode turns on → very low resistance path to ground → voltage divider with R → input signal above turn-on voltage drops across R, never reaching output.
  • Clip point: approximately V_clip + 0.7V (for silicon diodes).

🔧 Negative clipper design

Circuit structure (Figure 3.24):

  • Both diode and DC bias voltage flipped to opposite polarity compared to positive clipper.

How it works:

  • Diode remains reverse-biased as long as V_in > V_clip - 0.7V.
  • When V_in goes below this voltage, diode turns on, creating shorting path and limiting output voltage.

🔧 Dual-polarity (bipolar) clipper

  • Combines positive and negative clippers in one circuit.
  • Allows independent limiting of positive and negative swings.
  • Example: can clip at +4.7V and -8.7V simultaneously, creating asymmetric waveforms.

🔍 Design considerations

🔍 Resistor R selection

  • Key principle: R must be significantly larger than the diode's on-resistance.
  • Typical diode dynamic resistance when fully on: under 100Ω.
  • A 10kΩ resistor provides more than sufficient resistance ratio for an effective voltage divider.
  • Don't confuse: the precise value of R is unimportant—only the ratio matters.

🔍 Calculating clip levels

For a dual clipper with two bias voltages:

  • Positive clip level = V₁ + 0.7V (forward potential of silicon diode).
  • Negative clip level = -(V₂ + 0.7V).

Example: V₁ = 4V and V₂ = 8V yields clipping at +4.7V and -8.7V.

🔍 Expected waveform behavior

  • A sine wave input clipped at asymmetric levels produces a "lopsided cross between a sine wave and a square wave."
  • The portions exceeding clip thresholds are flattened while the middle portion retains its original shape.
18

Clampers

3.4 Clampers

🧭 Overview

🧠 One-sentence thesis

Clampers automatically add a DC offset to an AC signal to make it uni-polar by using a capacitor that charges to the peak input voltage and creates a level shift that tracks changes in signal amplitude.

📌 Key points (3–5)

  • What clampers do: add a DC offset to shift an AC signal so it becomes uni-polar (all positive or all negative voltages).
  • How they work: a capacitor charges to the peak input value and creates the offset automatically, so the circuit adapts if the input amplitude changes.
  • Positive vs negative clampers: positive clampers shift the negative peak to zero volts; negative clampers shift the positive peak to zero volts.
  • Common confusion: clampers vs clippers—clippers limit voltage range by cutting off peaks; clampers shift the entire waveform without cutting.
  • Biased clampers: adding a DC source in series with the diode moves the clamped peak to a voltage other than zero.

🔄 What clampers are and why they matter

🔄 Definition and purpose

A clamper is a circuit that adds a DC offset to an AC signal in such a way that the resulting voltage is uni-polar.

  • Positive clamper: adds a positive offset so the former negative peak now sits at zero volts.
  • Negative clamper: adds a negative offset so the former positive peak now sits at zero volts.
  • Also called DC restorers because they restore or add a DC level.
  • The key advantage: the circuit automatically determines the needed DC shift, so if the input amplitude changes, the offset tracks with it.

🎯 The challenge

  • The concept is simple—just add a DC voltage to the AC signal.
  • The trick: getting the circuit to figure out what the DC shift needs to be on its own, without manual adjustment.

⚙️ How clampers operate

⚙️ Prototype: fixed DC offset

  • A simple approach uses a fixed DC source in series with the input (Figure 3.27 in the excerpt).
  • If the DC offset equals the peak value of the input, the negative peak rises to zero volts.
  • Limitation: the offset is fixed and won't adapt if the input amplitude changes.

🔋 Capacitor-based clamper (the practical solution)

  • Replace the fixed DC source with a capacitor (Figure 3.28).
  • The capacitor voltage varies with the peak value of the input, so it precisely compensates to produce an ideally clamped output.
  • Critical requirement: the RC time constant (capacitor × parallel resistor) must be much longer than the period of the input waveform.

🔄 Step-by-step operation (positive clamper)

  1. Initial positive cycle: capacitor is uncharged, diode is reverse-biased. Output follows input because the RC time constant is long.
  2. Input swings negative: diode turns on, bypassing the parallel resistor. This drastically reduces the charge time constant, so the capacitor voltage tracks the negative portion of the input while the output stays near zero volts.
  3. At negative peak: capacitor voltage reaches the negative peak value of the input (polarity: minus-to-plus from left to right, per Kirchhoff's voltage law).
  4. Input reverses and rises: diode turns off due to the voltage now held on the capacitor. The capacitor now behaves like a fixed DC source equal to the negative peak.
  5. Result: as the input swings positive from its negative peak, the output starts at zero volts and tracks the input upward, producing the desired level shift.

Example: If the input is a 10 V peak sine wave (swinging from +10 V to −10 V), the capacitor charges to 10 V. When the input is at its negative peak (−10 V), the capacitor adds +10 V, so the output is at 0 V. When the input swings to its positive peak (+10 V), the output is at +20 V. The waveform is now uni-polar (0 V to +20 V).

⚠️ Real-world imperfections

  • Diode forward voltage drop: the negative peak won't be exactly at zero volts but around −0.7 V (for silicon diodes).
  • Settling time: it may take more than one cycle to "grab" the peak value, depending on the period and the charge/discharge time constants.

🔀 Variations: negative and biased clampers

🔀 Negative clamper

  • Flip the polarity of the diode.
  • The circuit now shifts the positive peak to zero volts instead of the negative peak.

⚡ Biased clamper

  • Add a DC source in series with the diode (like the biased clipper approach).
  • This moves the clamped peak to a voltage other than zero.
  • Figure 3.29 shows a biased positive clamper.
Clamper typeDiode polarityDC sourceResult
PositiveStandardNoneNegative peak at ~0 V
NegativeFlippedNonePositive peak at ~0 V
Biased positiveStandardIn series with diodeNegative peak at DC source voltage (minus diode drop)

📐 Example 3.4 walkthrough

Given: 10 V peak sine wave at 1 kHz, C = 10 μF, R = 10 kΩ, V_clamp = 5.7 V, silicon diode.

Analysis:

  • Period = 1/f = 1 ms.
  • Discharge time constant τ = R × C = 10 kΩ × 10 μF = 100 ms.
  • τ is much longer than the period ✓
  • DC clamping source produces +5 V offset (5.7 V − 0.7 V diode drop).
  • Input is 10 V peak, so 20 V peak-to-peak (−10 V to +10 V).
  • Expected output: 20 V peak-to-peak sine wave swinging between +5 V and +25 V.

Simulation result (Figure 3.31): output matches prediction precisely after the initial charge phase (first ~10 ms).

🆚 Don't confuse: clampers vs clippers

  • Clippers (section 3.3): limit the voltage range by cutting off peaks above/below a threshold. The waveform is distorted (flattened at the clip points).
  • Clampers (section 3.4): shift the entire waveform up or down without cutting. The shape is preserved; only the DC level changes.

🔧 Design considerations

🔧 Time constant requirement

  • The discharge time constant (R × C) must be much longer than the input period.
  • If too short, the capacitor will discharge significantly between cycles and fail to hold the offset.
  • Rule of thumb: τ should be at least 10–100 times the period.

🔧 Component selection

  • Capacitor: larger values increase the time constant and improve clamping stability.
  • Resistor: must be significantly larger than the on-resistance of the diode (typically the diode's dynamic resistance is under 100 Ω when fully on, so R should be in the kΩ range).
  • Diode: silicon switching diodes (e.g., 1N914) are commonly used; the 0.7 V forward drop affects the final clamped level.

🔧 Settling behavior

  • The circuit may require several input cycles to reach steady-state operation.
  • Simulations often delay the observation window (e.g., 10 ms in Example 3.4) to skip the initial charge phase and view stable clamping.
19

Bipolar Junction Transistors (BJTs)

Chapter 4: Bipolar Junction Transistors (BJTs)

🧭 Overview

🧠 One-sentence thesis

Bipolar junction transistors are three-layer semiconductor devices that act as current-boosting components, requiring forward-reverse biasing to operate and exhibiting stable performance in switching applications when driven into saturation.

📌 Key points (3–5)

  • What a BJT is: A three-layer (NPN or PNP) semiconductor device with emitter, base, and collector terminals that amplifies current through forward-reverse biasing.
  • How it works: The thin, lightly-doped base allows most injected carriers to reach the collector rather than recombine, creating current gain (β = IC/IB).
  • Key parameters: Alpha (α = IC/IE, typically >0.95) and beta (β = IC/IB, typically 100–200 for small signal devices) describe current relationships.
  • Common confusion: Saturation vs. constant-current operation—saturation provides stable current independent of β variation, while constant-current region is used for linear amplification.
  • Why it matters: BJTs enable switching, driving, and amplification applications; understanding saturation and load lines is critical for reliable circuit design.

🔬 Physical structure and operation

🧱 Basic construction

Bipolar junction transistor (BJT): A three-layer semiconductor device with alternating N-type and P-type materials forming two PN junctions.

  • NPN configuration: N-emitter, P-base, N-collector sandwich structure
  • PNP configuration: P-emitter, N-base, P-collector (inverse of NPN)
  • The base region is thin and lightly doped (critical for operation)
  • The collector is the largest region
  • Each layer has an external lead attached

Don't confuse: The simplified diagram with actual construction—real BJTs are built in "layer cake" fashion (bottom to top), not side-by-side, though spatial orientation doesn't affect operation.

⚡ Forward-reverse bias operation

The key operating condition requires:

  • Base-emitter junction: forward-biased (VBE ≈ 0.7V for silicon)
  • Base-collector junction: reverse-biased (VCB is large)
  • For NPN: VC > VB > VE (collector most positive, emitter at lowest potential)

How current flows (using electron flow for NPN):

  1. Electrons exit the emitter supply and enter the N-emitter (majority carriers)
  2. Forward bias pushes electrons over the energy hill into the P-base
  3. Only a small percentage (1–5%) recombine with base holes → base current (IB)
  4. The vast majority (95–99%+) reach the collector-base depletion region
  5. Electrons flow through the collector back to the positive supply terminal

Example: If 100 electrons enter from the emitter, perhaps 2 exit through the base terminal while 98 continue to the collector—this creates the current gain effect.

🔄 PNP operation

  • All current directions reverse compared to NPN
  • Conventional current flows INTO the emitter, OUT of collector and base
  • Voltage polarities reverse: VBE ≈ −0.7V
  • Same equations (α, β) apply
  • Any NPN circuit has a PNP counterpart with reversed supplies

📊 Key parameters and relationships

📐 Alpha (α)

Alpha (α): The ratio of collector current to emitter current.

  • Formula: α = IC / IE
  • Typical value: greater than 0.95
  • Represents the fraction of emitter current that reaches the collector
  • Related to β by: α = β / (β+1)

📈 Beta (β)

Beta (β): The ratio of collector current to base current, also called current gain or hFE.

  • Formula: β = IC / IB
  • Typical values: 100–200 for small signal transistors; 25–50 for power transistors
  • Also appears on datasheets as hFE (one of four hybrid parameters)
  • Related to α by: β = α / (1−α)
  • Critical insight: IC = β · IB

🔗 Current relationships

From Kirchhoff's Current Law:

  • IE = IC + IB (emitter current equals sum of collector and base currents)
  • Since IC >> IB, therefore IE ≈ IC
  • The base current is much smaller than the other two currents

Don't confuse: The simple two-diode model with actual BJT behavior—while an ohmmeter test would show two diodes, the thin base creates the current gain effect that two separate diodes cannot produce.

📉 Collector curves and operating regions

📊 Family of collector curves

A plot of IC versus VCE at various fixed IB levels reveals device behavior:

RegionLocationCharacteristicsUse
SaturationExtreme left (VCE < VCE(sat))Current rises rapidly; VCE(sat) ≈ 0.1–0.2VSwitching applications
Constant currentMiddle regionIC relatively flat, modest positive slopeLinear amplifiers
BreakdownExtreme right (VCE > BVCEO)Current rises rapidly; device damage riskAvoid this region

🌡️ Beta variation and Early voltage

Why IC rises with VCE in constant-current region:

  • Increased VCE → increased VCB (reverse bias on collector-base junction)
  • Wider depletion region penetrates further into the base
  • Effectively narrower base → less recombination → reduced IB → increased β

Early Voltage (VA): The voltage where extended constant-current region traces intersect in the second quadrant.

🔍 Determining β from curves

Method using curve tracer display:

  1. Locate the operating point (VCE, IC) on the graph
  2. Find the nearest trace to that point
  3. Read the precise IC value at the intersection
  4. Count traces upward to determine IB (number of traces × step size)
  5. Calculate β = IC / IB

Example: At VCE = 30V, if IC = 4.2mA on the fourth trace with 10μA steps, then IB = 40μA, so β = 4.2mA / 40μA = 105.

🔌 DC load lines and circuit analysis

📐 Load line concept

DC load line: A plot of all possible coordinate pairs of IC and VCE for a transistor in a given circuit.

Starting from KVL in the collector-emitter loop:

  • VCE = VCC − IC·RC
  • Rearranging: IC = (−1/RC)·VCE + VCC/RC
  • This is a linear equation (y = mx + b)

Key points on the load line:

  • Y-intercept (IC(sat)): Maximum collector current = VCC / RC (when VCE = 0)
  • X-intercept (VCE(cutoff)): Maximum voltage = VCC (when IC = 0)
  • Slope: −1 / RC
  • Q point: The actual operating point for a specific transistor (ICQ, VCEQ)

🎯 Using load lines

All possible operating points lie on the load line:

  • If calculated IC > IC(sat), the actual current is IC(sat) (transistor saturated)
  • If calculated IC < 0, the actual current is ≈0 (transistor in cutoff)
  • Different β values produce different Q points along the same load line

Example: For VCC = 15V and RC = 1kΩ:

  • IC(sat) = 15V / 1kΩ = 15mA
  • VCE(cutoff) = 15V
  • With β = 100: Q point might be at IC = 4.65mA, VCE = 10.35V
  • With β = 200: Q point shifts to IC = 9.3mA, VCE = 5.7V

⚠️ Beta variation problems

The simple base-bias circuit (fixed base voltage) suffers from instability:

  • β varies widely (3:1 or more) between devices and with temperature
  • β increases with temperature → IC increases → power increases → temperature rises further
  • Creates inadvertent thermal positive feedback loop
  • Result: IC and VCE vary significantly between devices or operating conditions

Don't confuse: Calculated current with actual current when saturation occurs—if the math suggests IC = 18.6mA but VCC/RC = 15mA, the transistor cannot exceed 15mA (it saturates).

🔧 Switching and driver applications

🔀 Saturating switch principle

Key insight: Saturation is a fixed, stable value independent of β variation.

When saturated:

  • IC = IC(sat) = (VCC − VLED) / RC (for LED driver example)
  • β effectively forced to whatever value produces IC(sat)
  • Design rule: Make IB large enough that IC(sat) / IB << β (ratio of 10:1 or more guarantees hard saturation)

💡 Saturating LED driver (positive logic)

Circuit operation:

  • Logic low (0V): No base current → no collector current → LED off (cutoff)
  • Logic high: Base current flows → transistor saturates → LED current = IC(sat) → LED on

Advantages:

  • Stable LED current regardless of β variation
  • Offloads current demand from logic circuit (logic only supplies IB, not ILED)
  • Reliable on/off switching

Example calculation:

  • Logic voltage = 5V, RB = 4.7kΩ → IB = (5V − 0.7V) / 4.7kΩ = 915μA
  • VCC = 5V, VLED = 1.8V, RC = 330Ω → IC(sat) = (5V − 1.8V) / 330Ω = 9.7mA
  • Ratio = 9.7mA / 915μA ≈ 10.6:1 → guarantees saturation

Inverting logic: Use PNP transistor—logic low turns LED on, logic high turns it off.

🎛️ Non-saturating driver

Circuit uses emitter resistor (RE) to "bootstrap" the emitter voltage:

Operation:

  • Logic low: No base-emitter voltage → no current → LED off
  • Logic high: VE follows logic input (VE = Vlogic − VBE) → IE = (Vlogic − VBE) / RE
  • Since IC ≈ IE, the LED current is directly programmed by RE and logic voltage

Advantages:

  • Requires less current from logic circuit
  • Stable current independent of β (if β varies, IB changes inversely with no change in IC)

Disadvantages:

  • Higher transistor power dissipation (VCE is larger)
  • Requires VCC higher than logic level

Example: Vlogic = 5V, RE = 270Ω → IC ≈ (5V − 0.7V) / 270Ω = 15.9mA (β not needed in calculation).

⚡ Motor drive application

Pulse width modulation (PWM) technique:

  • Fast pulses saturate the BJT (acts as switch)
  • Motor responds to averaged pulse value, not individual pulses
  • Narrow, widely-spaced pulses → low average → slow speed
  • Wide, closely-spaced pulses → high average → fast speed

Critical component: Snubbing diode (also called flyback diode, commutating diode, clamp diode)

  • Motor winding has high inductance
  • When BJT turns off, current cannot stop instantly → large flyback voltage (inductive kick)
  • Flyback voltage appears across BJT and could damage it
  • Snubbing diode short-circuits the winding during voltage reversal, preventing spike
  • Otherwise reverse-biased and out of circuit

🔋 Zener follower

Improved voltage regulation circuit:

How it works:

  • Zener diode reverse-biased through resistor R → establishes fixed VZ
  • Output voltage = VZ − VBE (both fixed → stable output)
  • Any input variation (e.g., ripple) drops across the BJT (VCE = VCB + VBE)
  • Current draw reflects load demand (more efficient than simple Zener regulator)

Advantages over simple Zener regulator:

  • Low Zener current → modest power dissipation
  • Efficient: draws current only as needed by load
  • BJT absorbs input variations

🔬 Ebers-Moll model

🧩 Model components

Ebers-Moll model: A simplified BJT model using an ideal diode for the base-emitter junction and a current-controlled current source at the collector-base.

Components:

  • Ideal diode from base to emitter (represents forward-biased B-E junction)
  • Current-controlled current source from collector to base (source value = β·IB)
  • Sufficient for good DC and low-frequency circuit analysis

📝 Circuit analysis approach

Start in the base-emitter loop (not collector-emitter):

  • B-E junction has known voltage (≈0.7V for silicon)
  • C-E voltage is unknown (depends on circuit elements)

Basic equations:

  • Base-emitter loop (KVL): VBB = IB·RB + VBE → IB = (VBB − VBE) / RB
  • Collector current: IC = β·IB
  • Collector-emitter loop (KVL): VCE = VCC − IC·RC

Example: VBB = 10V, VCC = 15V, RB = 200kΩ, RC = 1kΩ, β = 100

  1. IB = (10V − 0.7V) / 200kΩ = 46.5μA
  2. IC = 100 × 46.5μA = 4.65mA
  3. VCE = 15V − 4.65mA × 1kΩ = 10.35V

Limitation: Simple base-bias circuit lacks stability—β variation causes large changes in IC and VCE.

📚 Practical considerations

📄 Datasheet interpretation

Key specifications (example: 2N3904):

  • Case style: TO-92 plastic (through-hole mounting)
  • Maximum power dissipation: 625mW at 25°C ambient
  • Maximum collector current: 200mA
  • Maximum VCE: 40V (BVCEO)
  • β (hFE) range: 100–300 at IC = 10mA (3:1 variation)

Important graphs:

  • Normalized β vs. IC and temperature: shows β variation under different conditions
  • VCE(sat) vs. IC and IB: critical for switching circuit design
  • β generally increases with temperature and varies with collector current

🔄 Emitter and collector not interchangeable

Although the structure might suggest symmetry:

  • Emitter and collector regions are physically optimized differently
  • Swapping leads usually results in unpredictable behavior
  • The arrow in the schematic symbol indicates the emitter (points to N material for NPN, toward base for PNP)

⚠️ Operating region summary

  • Cutoff: IB = 0 → IC ≈ 0 (transistor off)
  • Saturation: IB large enough that IC = IC(sat) (transistor fully on, acts as closed switch)
  • Constant current (active): IC = β·IB, used for linear amplification
  • Breakdown: VCE > BVCEO (avoid—device damage)
20

Bipolar Junction Transistors (BJTs)

4.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

The bipolar junction transistor (BJT) is a three-layer semiconductor device that functions as a current-boosting component by extending the basic diode structure with an additional PN junction.

📌 Key points (3–5)

  • What a BJT is: a three-layer structure (NPN or PNP) with three terminals—emitter, base, and collector—forming two PN junctions.
  • Core function: all BJTs are current-boosting devices; boosting current enables voltage and power amplification depending on circuit impedances.
  • Two types exist: NPN and PNP configurations; circuits often work best when both types are used together.
  • Common confusion: the simplified two-diode model predicts ohmmeter behavior but is very limited for understanding actual BJT operation.
  • Biasing matters: external voltage sources determine whether junctions are forward- or reverse-biased, controlling current flow.

🔧 Structure and basic operation

🧱 Physical structure

Bipolar junction transistor (BJT): a three-layer semiconductor device with alternating N-type and P-type materials, creating two PN junctions.

  • The BJT extends the basic diode by adding another segment of oppositely doped material to one end, creating a second PN junction.
  • Two configurations are possible:
    • NPN: N-type, P-type, N-type layers
    • PNP: P-type, N-type, P-type layers
  • Real BJTs are built in a "layer cake" fashion (bottom to top), though diagrams often show them as side-by-side for clarity.

📍 Three terminals

The three regions have different sizes and doping levels:

TerminalRegion sizeCharacteristics
CollectorLargestOne end of the sandwich
BaseThinLightly doped, middle layer
EmitterStandardOther end of the sandwich

⚡ Depletion regions

  • Above absolute zero, recombination occurs and two depletion regions form at the two PN junctions.
  • These depletion regions are devoid of free charges, similar to a basic diode.
  • The dissimilar Fermi levels of N-type and P-type materials create an "energy hill" at each junction.

🔌 The two-diode model

🔌 Simplified representation

Two-diode model: a very limited model representing the BJT as two diodes connected at the base terminal.

  • Because the BJT contains two depletion regions, it can be modeled (in a simplified way) as two diodes sharing a common terminal (the base).
  • Important limitation: the excerpt explicitly warns this is "a very limited model."

🧪 Ohmmeter test predictions

The two-diode model successfully predicts ohmmeter readings when testing a BJT:

  • Red (positive) lead on base, black (negative) on emitter or collector: low resistance (forward-biases the junction).
  • Leads reversed: high resistance (reverse-biases the junction).
  • Leads on emitter and collector (any polarity): high resistance (one junction is always reverse-biased, blocking current in the series path).

Example: Testing an NPN BJT with an ohmmeter, connecting positive to base and negative to emitter will show low resistance because the base-emitter junction is modestly forward-biased.

⚠️ Don't confuse

  • The two-diode model is useful for understanding basic junction behavior and ohmmeter tests.
  • It does not fully explain BJT operation in active circuits (the excerpt states "as we shall soon see").

⚙️ Biasing and current flow

⚙️ What biasing means

  • External DC voltage sources are added to control the polarity across each junction.
  • The circuit forms two loops: base-emitter (B-E) and base-collector (B-C).
  • Each loop can forward-bias or reverse-bias its respective junction.

🔋 Reverse-bias scenario

When both junctions are reverse-biased:

  • In the B-E loop, the emitter supply (V_EE) reverse-biases the base-emitter diode.
  • In the B-C loop, the collector supply reverse-biases the base-collector diode.
  • Result: virtually no current flows anywhere in the circuit.

🔄 Forward-bias scenario

  • The excerpt mentions that if the two supplies are reversed in polarity, the situation changes (the text cuts off before completing the explanation).
  • Implication: proper biasing (forward-biasing the correct junctions) is necessary for current flow and active operation.

📌 Why biasing matters

  • Without an external potential of the proper polarity, the junction will not allow current to flow.
  • The required magnitude is a function of the material used.
  • For a PN junction to conduct, the P material (anode) must be positive with respect to the N material (cathode).

🎯 Applications and importance

🎯 Wide range of uses

BJTs serve as the basis for:

  • Simple amplifiers
  • Device control circuits
  • Complex digital computing circuitry

🔧 Variations and optimization

BJTs are optimized for different applications:

Application domainExamples
FrequencyVery low to very high frequency work
PowerLow, medium, and high power
Cost/specializationInexpensive general purpose through highly specialized niche items

💡 Fundamental capability

  • No matter what a BJT has been optimized for, all BJTs can be considered current boosting devices.
  • If you can boost current, you can also boost voltage and power, depending on the associated impedances.
  • This makes the BJT a foundational electronic component.
21

Bipolar Junction Transistor Structure and Operation

4.1 Introduction

🧭 Overview

🧠 One-sentence thesis

The bipolar junction transistor (BJT) extends the basic diode by adding a third layer to create two PN junctions, enabling a small base current to control a much larger collector current through a thin, lightly doped base region.

📌 Key points (3–5)

  • Structure: BJT consists of three alternating layers (NPN or PNP) forming two PN junctions, with terminals named emitter, base, and collector.
  • Key mechanism: The thin, lightly doped base allows most injected electrons (95–99%) to reach the collector rather than recombine, creating current amplification.
  • Biasing requirement: Forward-biasing the base-emitter junction while reverse-biasing the base-collector junction produces high current in both loops.
  • Common confusion: A simple two-diode model predicts ohmmeter behavior but fails to explain actual BJT operation—the unified thin base (not two separate pieces) is critical.
  • Current relationships: Collector current approximately equals emitter current (both much larger than base current), with current gain β typically 100–200 for small-signal devices.

🏗️ Physical structure and basic configuration

🏗️ Three-layer sandwich

  • BJT adds a second portion of oppositely doped material to a basic diode, creating three alternating layers.
  • Two configurations exist: PNP or NPN (both are useful; circuits often combine both types).
  • Real BJTs are built in "layer cake" fashion (bottom to top), though diagrams often show them spatially for clarity.

Three terminals: emitter, base, and collector.

  • Collector: largest of the three regions.
  • Base: relatively thin and lightly doped (this is crucial for operation).
  • Emitter: third region completing the sandwich.

⚡ Two depletion regions

  • Above absolute zero, recombination creates two depletion regions at the two PN junctions.
  • Each depletion region is devoid of free charges, just like in a single diode.
  • The dissimilar Fermi levels create "energy hills" at both junctions.

🔌 Simple two-diode model and its limitations

🔌 What the model predicts correctly

A simplified model using two diodes back-to-back can predict ohmmeter test results:

Ohmmeter connectionResultWhy
Red (+) to base, black (−) to emitter or collectorLow resistanceForward-biases that junction
Reversed leadsHigh resistanceReverse-biases that junction
Between emitter and collector (any polarity)High resistanceOne junction always reverse-biased (series connection blocks current)

⚠️ Where the model fails

  • The two-diode model treats the base as two separate pieces of material.
  • In a real BJT, the base is a single thin, lightly doped region—this unified structure makes all the difference.
  • Don't confuse: the model works for static resistance checks but cannot explain active transistor operation (current amplification).

🔋 Biasing scenarios and current flow

🔋 Double reverse-bias

  • If both base-emitter and base-collector junctions are reverse-biased, virtually no current flows anywhere.
  • This is expected behavior, matching the diode model.

🔋 Double forward-bias

  • If both junctions are forward-biased, currents flow in both loops depending on supply voltages and resistors.
  • Again, this matches the diode model—no surprises.

⚡ Forward-reverse bias (active mode)

  • Base-emitter junction: forward-biased.
  • Base-collector junction: reverse-biased.
  • What happens: High current flows in both loops, and the two currents are nearly equal in magnitude.
  • This contradicts the simple diode model (which would predict negligible current in the reverse-biased loop).

🧲 The key mechanism: thin, lightly doped base

🧲 Electron injection and recombination

Using electron flow (dashed lines in diagrams):

  1. Electrons exit the emitter supply and enter the N-type emitter (majority carriers there).
  2. Sufficient emitter supply potential pushes electrons over the base-emitter energy hill into the base.
  3. In the base, electrons attempt to recombine with majority holes.
  4. Critical point: Because the base is thin and lightly doped, only a small percentage (1–5%) recombine and exit via the base terminal as base current (recombination current).
  5. The vast majority (95–99%) of electrons reach the base-collector depletion region and enter the collector.
  6. In the collector, electrons are again majority carriers and flow back to the positive terminal of the collector supply.

Example: If 100 electrons are injected from the emitter, roughly 1–5 exit through the base terminal, and 95–99 exit through the collector terminal.

🔄 Why emitter and collector are not interchangeable

  • At first glance, the NPN structure looks symmetric.
  • In real devices, emitter and collector regions are optimized differently and not physically identical.
  • Swapping emitter and collector leads usually results in unpredictable behavior.

📐 Current relationships and performance parameters

📐 Kirchhoff's current law

From KCL at the transistor node:

I_E = I_C + I_B

  • Emitter current equals the sum of collector and base currents.

📐 Approximations for analysis

  • I_C >> I_B (collector current much greater than base current).
  • Therefore, I_E ≈ I_C (emitter and collector currents are approximately equal).
  • Conventional current flows into the collector and base, out of the emitter.

📐 Junction voltages

  • Base-emitter junction is forward-biased: V_BE ≈ 0.7 V (for silicon).
  • Base-collector junction is reverse-biased: V_CB is large.

📊 Performance parameters

ParameterSymbolDefinitionTypical value
AlphaαRatio of collector current to emitter current (I_C / I_E)Greater than 0.95
Beta (current gain)β or h_FERatio of collector current to base current (I_C / I_B)100–200 (small-signal); smaller for power transistors
  • β is also called current gain: if base current is the input signal and collector current is the output, β represents the signal boost.
  • h_FE is one of four hybrid parameters found on transistor spec sheets.
  • Don't confuse: α is always less than 1 (since I_C < I_E), but β is typically much greater than 1 (since I_B is very small).

🎯 Summary of transistor operation

When properly biased (base-emitter forward, base-collector reverse):

  • A small base current controls a much larger collector current.
  • The thin, lightly doped base is the key: it allows most charge carriers to "leak through" to the collector rather than recombining.
  • This creates current amplification: β = I_C / I_B is typically 100–200.
  • The energy diagram shows two depletion regions with energy hills, but the forward-biased base-emitter junction allows carriers to overcome the first hill and reach the collector.
22

4.2 The Bipolar Junction Transistor

4.2 The Bipolar Junction Transistor

🧭 Overview

🧠 One-sentence thesis

The bipolar junction transistor (BJT) operates fundamentally differently from a simple pair of diodes because its thin, lightly doped base allows most charge carriers injected from the emitter to reach the collector rather than recombine in the base, creating a controllable current amplification effect.

📌 Key points (3–5)

  • The key structural difference: the BJT's base is thin and lightly doped, unlike a dual-diode model where the base would be split into two separate pieces—this makes all the difference in operation.
  • Forward-reverse bias behavior: when the base-emitter junction is forward-biased and the base-collector junction is reverse-biased, high currents flow in both loops (not just the forward-biased loop), and these currents are nearly equal.
  • Current relationships: 95% to over 99% of electrons injected into the base reach the collector; only a small percentage recombine in the base, so collector current is much larger than base current and approximately equal to emitter current.
  • Common confusion: emitter and collector cannot simply be swapped—real devices optimize these regions differently, so reversing leads causes unpredictable behavior.
  • Two transistor types: NPN and PNP are logical inverses regarding current directions and voltage polarities, but all fundamental equations remain applicable.

🔬 How the BJT differs from simple diodes

🔬 The dual-diode model fails

  • A BJT looks like two diodes back-to-back (base-emitter and base-collector junctions).
  • With simple diodes in forward-reverse bias, you'd expect high current in the forward-biased loop and negligible current in the reverse-biased loop.
  • With a BJT, this is not what happens: high current flows in both loops, and the currents are very nearly equal in magnitude.

🧱 The thin, lightly doped base is the key

The base of the BJT is thin and lightly doped.

  • The dual-diode model splits the base into two separate pieces of material—that structural difference is critical.
  • The thin, lightly doped base allows most charge carriers to pass through rather than recombine.
  • Example: if the base were thick and heavily doped (like in separate diodes), most electrons would recombine and exit through the base terminal; the thin base prevents this.

⚡ Current flow mechanism in forward-reverse bias

⚡ Electron injection from emitter

  • Electrons exit the emitter supply and enter the N-type emitter, where they are the majority carrier.
  • The base-emitter depletion region creates an energy hill (just like a single PN junction).
  • Sufficient potential from the emitter supply pushes electrons over the hill into the base.

🔄 What happens in the base

  • Injected electrons attempt to recombine with the majority base holes.
  • Because the base is physically thin and lightly doped, only a small percentage (1% to 5%) of injected electrons recombine with base holes.
  • These recombining electrons exit the base terminal back to ground, forming the base current (also called recombination current).

➡️ Most electrons reach the collector

  • The vast majority of remaining electrons (95% to over 99%) find their way to the base-collector depletion region and then to the collector.
  • Once in the collector, electrons are again the majority carrier and flow back to the positive terminal of the collector power supply.
  • This creates a high collector current even though the base-collector junction is reverse-biased.

⚠️ Don't confuse: emitter and collector are not interchangeable

  • At first glance, it might seem the emitter and collector leads could be swapped with no change.
  • With real-world devices this is not possible: the emitter and collector regions are optimized and not physically identical.
  • Placing transistors backwards (emitter and collector swapped) will usually result in unpredictable behavior.

📐 BJT performance summary and parameters

📐 Key operating characteristics

  • From Kirchhoff's Current Law: emitter current equals collector current plus base current.
  • Collector current is much greater than base current, therefore emitter current approximately equals collector current.
  • The base-emitter junction is forward-biased, therefore base-emitter voltage is approximately 0.7 V (for silicon).
  • The base-collector junction is reverse-biased, therefore collector-base voltage is large.
  • Conventional current flows into the collector and base, and out of the emitter.

📊 Alpha and beta parameters

ParameterSymbolDefinitionTypical value
AlphaαRatio of collector current to emitter currentGreater than 0.95
Beta (current gain)β or h_FERatio of collector current to base current100–200 (small signal); 25–50 (power transistors)
  • Alpha: collector current divided by emitter current; typically greater than 0.95.
  • Beta: collector current divided by base current; also called h_FE (one of four hybrid parameters) or current gain.
  • If base current is the input signal and collector current is the output signal, beta represents the amount of signal boost or gain.
  • Relationships: alpha equals beta divided by (beta plus 1); beta equals alpha divided by (1 minus alpha); collector current equals beta times base current.

🔣 Schematic symbol for NPN

  • The standard symbol may place the body within a circle (common variation).
  • Following the standard, the arrow points to N material and in the direction of easy conventional current flow.
  • The arrow is on the emitter terminal.

🔁 The PNP bipolar junction transistor

🔁 How PNP differs from NPN

  • The PNP version is created by swapping the material for each layer.
  • The outcome is the logical inverse of the NPN regarding current directions and voltage polarities.

⬅️ Current and voltage reversals

  • Conventional current flows into the emitter, and out of the collector and base (echoing the electron flow of the NPN).
  • Voltages across the device have reversed polarity; for example, base-emitter voltage is approximately −0.7 V.
  • All other characteristics remain unchanged, so the alpha and beta equations are still applicable.

🔣 PNP schematic symbol

  • Just about any NPN-based circuit has its PNP counterpart.
  • The schematic symbol reverses the emitter arrow.
  • As the base is now the N material, the arrow points toward the base.

📈 BJT collector curves

📈 What collector curves show

A family of collector curves: a series of plots of collector current versus collector-emitter voltage at varying levels of base current.

  • To generate these curves: drive the base terminal with a fixed current source (establishing base current), attach a DC power supply from collector to emitter and sweep it from zero volts to some upper value (establishing collector-emitter voltage), and track the resulting collector current.
  • This process is repeated at increasing levels of base current, with each new base current stepping up a fixed amount (e.g., 0 μA, 10 μA, 20 μA, 30 μA, etc.).
  • The bottom curve results when base current equals 0; ideally collector current would be 0, but a small leakage current occurs (called I_CEO: collector-emitter current with the base terminal open).

🗺️ Three distinct regions of operation

RegionLocationCharacteristicApplication
SaturationExtreme leftCurrent rises rapidly; break-over at a few tenths of a volt (V_CE(sat))Transistor switching
Constant currentMiddleCollector current relatively constant with only modest positive slopeLinear amplifiers
BreakdownExtreme rightCollector current rises rapidly at breakdown voltage (BV_CEO)Avoid—damage may result

🔧 Practical notes

  • Saturation region: the break-over point (V_CE(sat)) is fairly small at just a few tenths of a volt; found on data sheets.
  • Breakdown region: same effect as with individual diodes; breakdown voltage denoted as BV_CEO (collector to emitter voltage with an open base); typically 30–60 volts for general purpose devices, but can be much higher.
  • Constant current region: where the transistor should operate for linear amplifier applications.
  • A device called a curve tracer can be used to generate this family of curves in the lab; a very good approximation for beta can be determined using these curves.
23

4.3 BJT Collector Curves

4.3 BJT Collector Curves.

🧭 Overview

🧠 One-sentence thesis

The family of collector curves reveals three distinct operating regions—saturation, constant current, and breakdown—that determine how BJTs behave in switching versus amplifier applications.

📌 Key points (3–5)

  • What collector curves plot: collector current (I_C) versus collector-emitter voltage (V_CE) at different fixed base currents (I_B), creating a family of traces.
  • Three regions: saturation (extreme left, rapid current rise), constant current (middle, nearly flat), and breakdown (extreme right, rapid rise again).
  • Application match: saturation region is used for switching; constant current region is used for linear amplifiers.
  • Common confusion: β is not constant—it varies with collector current, temperature, and V_CE; the constant current region shows only modest slope, not perfectly flat.
  • Why β rises with V_CE: increased reverse bias on the collector-base junction widens the depletion region, narrows the effective base, reduces recombination, and thus increases β.

📊 How collector curves are generated

📊 Measurement procedure

  • Drive the base terminal with a fixed current source to establish I_B.
  • Attach a DC power supply from collector to emitter and sweep it from zero volts to some upper value to establish V_CE.
  • Simultaneously track the resulting collector current and plot I_C versus V_CE—this produces one trace.
  • Increase the base current by a fixed step (e.g., 10 μA) and repeat the sweep for the next trace.
  • Repeat to create a family of curves.

🔍 What each trace represents

  • The bottom curve corresponds to I_B = 0; ideally I_C would be zero, but a small leakage current occurs, called I_CEO (Collector-Emitter current with base Open).
  • Each curve above corresponds to a higher base current, stepping up by a fixed amount (e.g., 0 μA, 10 μA, 20 μA, 30 μA, etc.).
  • Example: if the step is 10 μA, the fourth trace up (not counting the bottom I_B = 0 trace) corresponds to I_B = 40 μA.

🗺️ Three operating regions

⚡ Saturation region (extreme left)

Saturation region: the zone at the extreme left of the curve where collector current rises rapidly with small increases in V_CE.

  • The break-over point is just a few tenths of a volt, found on data sheets as V_CE(sat).
  • This region is used in transistor switching applications.
  • Don't confuse: saturation here means the transistor is "fully on," not that current has stopped increasing.

🌊 Constant current region (middle)

Constant current region: the zone between saturation and breakdown where collector current is relatively constant, showing only a modest positive slope.

  • This is where the transistor operates for linear amplifiers.
  • The current is not perfectly flat; there is a modest positive slope due to the Early effect (explained below).
  • Example: for a given base current trace, I_C changes only slightly as V_CE increases from a few volts to tens of volts.

💥 Breakdown region (extreme right)

Breakdown region: the zone at the extreme right where collector current rises rapidly, similar to diode breakdown.

  • The breakdown voltage is denoted BV_CEO (Collector to Emitter voltage with an Open base).
  • For general-purpose devices, BV_CEO is typically 30 to 60 volts, but can be much higher.
  • Do not operate devices in this region—damage may result.
  • This is the same effect seen with individual diodes.

🔧 Using collector curves to find β

🔧 Graphical method

A curve tracer can generate the family of curves in the lab, and you can approximate β from the graph:

  1. Determine the approximate circuit values for I_C and V_CE, and locate this point on the graph.
  2. Find the nearest plot line (trace) to that point.
  3. From the intersection of V_CE and this trace, track back to the vertical axis to find the precise value of I_C for that trace.
  4. Count the number of traces up to the selected trace (not including the bottom I_B = 0 trace) and multiply by the base current step size to determine the corresponding I_B.
  5. Divide: β = I_C / I_B.

📐 Example scenario

Example: A BJT operates at V_CE = 30 V and I_C = 4 mA. The curve tracer shows base current increases by 10 μA per trace.

  • Draw a vertical line at 30 V and find the trace nearest to 4 mA—suppose it's the second trace from the top.
  • Track back to the vertical axis: the precise I_C is roughly 4.2 mA.
  • Count up: the selected trace is the fourth one up (not counting the bottom I_B = 0 trace).
  • So I_B = 4 × 10 μA = 40 μA.
  • β = 4.2 mA / 40 μA = 105.

🌀 Why collector current rises with V_CE (Early effect)

🌀 Physical mechanism

  • Increased V_CE means increased collector-base voltage (by definition, V_CE = V_CB + V_BE).
  • V_CB is the reverse-bias potential on the collector-base PN junction.
  • As this reverse potential increases, the collector-base depletion region widens.
  • The widening depletion region penetrates further into the base layer, effectively narrowing the base.
  • A narrower base means fewer chances for recombination, thus reducing base current and effectively increasing β.

📏 Early voltage (V_A)

Early voltage (V_A): the point in the second quadrant where the extended constant current region traces intersect, named after James Early.

  • If you extend the constant current region traces back into the second quadrant (negative V_CE), they intersect at a single point called V_A.
  • This voltage quantifies the slope of the constant current region: a larger V_A means flatter curves (less variation in I_C with V_CE).
  • Don't confuse: V_A is not an operating voltage; it is a model parameter derived from the slope of the curves.

📉 Variation of β with operating conditions

📉 β is not constant

From the 2N3904 data sheet example:

  • Nominal β (h_FE) varies under different conditions.
  • At particularly small or large collector currents, β tends to drop off.
  • At 10 mA, the data sheet shows a wide 3:1 variance (range of 100 to 300).

🌡️ Normalized β graph

The data sheet includes a graph of normalized β versus collector current and temperature:

  • The vertical axis plots normalized β, which is a ratio used to compare β under varying conditions, not the absolute value.
  • Example: at room temperature (25°C) and 10 mA, normalized β = 1.0. If a particular transistor has β = 200 at these conditions, then at 0.2 mA and 25°C (where normalized β = 0.7), the actual β would be (0.7 / 1.0) × 200 = 140.
  • Generally, β tends to increase with increasing temperature.

⚙️ V_CE(sat) graph

The data sheet also plots collector-emitter saturation voltage (V_CE(sat)) for various current conditions:

  • This is an important parameter for transistor switching circuits.
  • V_CE(sat) is the voltage across the transistor when it is fully "on" in the saturation region.

🔄 Summary table of regions

RegionLocationCharacteristicApplicationKey parameter
SaturationExtreme leftCurrent rises rapidly with small V_CESwitching (transistor "on")V_CE(sat) (few tenths of a volt)
Constant currentMiddleI_C relatively constant, modest positive slopeLinear amplifiersEarly voltage (V_A)
BreakdownExtreme rightCurrent rises rapidly, damage riskAvoid operation hereBV_CEO (30–60 V typical)
24

4.4 BJT Data Sheet Interpretation

4.4 BJT Data Sheet Interpretation

🧭 Overview

🧠 One-sentence thesis

BJT data sheets provide maximum ratings, beta (β) values under various conditions, and characteristic graphs that reveal how transistor parameters vary with current and temperature, enabling designers to predict real-world device behavior.

📌 Key points (3–5)

  • Maximum ratings define safe operating limits: power dissipation, collector current, and collector-emitter voltage cannot all be at maximum simultaneously.
  • Beta (β or hFE) varies widely: even for the same transistor model, β can vary 3:1 at a given current, and it changes with collector current and temperature.
  • Normalized graphs show relative changes: data sheets use normalized β plots (ratio format) to show how β changes under different conditions relative to a reference point.
  • Common confusion: the data sheet shows a range of β values (e.g., 100 to 300), not a single fixed value—each individual transistor will have its own β within that range.
  • Saturation voltage (VCE(sat)) is critical for switching: graphs show how this parameter varies with current, which is important for transistor switching circuit design.

📋 Understanding maximum ratings

📋 What the data sheet specifies

The excerpt uses the 2N3904 NPN transistor as an example:

  • Case style: TO-92 plastic case for through-hole mounting, commonly used for small signal transistors.
  • Maximum power dissipation: 625 mW in free air at 25°C ambient temperature.
  • Maximum collector current: 200 mA.
  • Maximum collector-emitter voltage: 40 V.

⚠️ Simultaneous limits constraint

The device cannot withstand maximum current and voltage simultaneously.

  • This is a critical design constraint: you cannot operate at 200 mA and 40 V at the same time.
  • The power dissipation limit (625 mW) restricts the product of voltage and current.
  • Example: If you operate at maximum voltage (40 V), the current must be limited to about 15.6 mA (625 mW ÷ 40 V) to stay within the power limit.

📊 Beta (β) characteristics and variation

📊 Beta notation and nominal values

  • The data sheet lists β as hFE (a standard notation in transistor specifications).
  • The excerpt mentions "nominal values for β under various conditions" found in Figure 4.13b.
  • At particularly small or large collector currents, β tends to drop off.

📏 Wide variation range

  • At 10 mA collector current, the 2N3904 shows a 3:1 variance in β.
  • The data sheet indicates a range of 100 to 300 for β at room temperature and 10 mA.
  • Don't confuse: this is not measurement error—different individual transistors of the same model will have different β values within this range.

🌡️ Normalized β graphs

Normalized β is plotted on the vertical axis. That is, this is not the expected value but is a ratio used to compare β under varying conditions.

How to read normalized β graphs:

  • The graph shows β variation with both collector current and temperature.
  • At the reference point (room temperature, 10 mA), the normalized value is 1.0.
  • Other conditions show a ratio relative to this reference.

Example from the excerpt:

  • Suppose you measure a particular 2N3904 and find β = 200 at 10 mA and 25°C (the reference point where normalized β = 1.0).
  • If you operate this same transistor at 0.2 mA and 25°C, the graph shows normalized β = 0.7.
  • The actual β under these new conditions = (0.7 ÷ 1.0) × 200 = 140.

🔥 Temperature effects

  • The graph shows that, generally speaking, β tends to increase with increasing temperature.
  • This is important for circuit design because ambient temperature changes will affect transistor gain.

🔌 Saturation voltage characteristics

🔌 VCE(sat) parameter

  • The middle graph (mentioned in Figure 4.13c) plots collector-emitter saturation voltage for various current conditions.
  • This is the voltage across the collector-emitter junction when the transistor is fully "on" (saturated).

🔄 Why saturation voltage matters

This is an important parameter when dealing with transistor switching circuits.

  • In switching applications, the transistor operates either fully off or fully on (saturated).
  • A lower VCE(sat) means less power dissipation when the transistor is conducting.
  • The graph shows how VCE(sat) varies with current, helping designers predict voltage drops in switching circuits.

🔧 Practical implications for circuit design

🔧 Beta variability requires robust design

FactorEffect on βDesign implication
Collector current (very low or very high)β drops offAvoid extreme current ranges for stable gain
Temperature increaseβ generally increasesCircuit must tolerate gain changes with temperature
Device-to-device variation3:1 range (100–300)Design must work across the entire β range

🔧 Using data sheet information

  • Start with the nominal β range (e.g., 100 to 300 at 10 mA).
  • Use normalized graphs to adjust for your actual operating current and temperature.
  • Check that your design stays within maximum ratings (power, current, voltage).
  • For switching circuits, verify VCE(sat) at your operating current to calculate power dissipation.

Don't confuse: The data sheet β range is not a tolerance or error band—it represents the actual spread of β values you will encounter across different transistors of the same part number.

25

4.5 Ebers-Moll Model

4.5 Ebers-Moll Model

🧭 Overview

🧠 One-sentence thesis

The Ebers-Moll model provides a simplified functional representation of the BJT that enables DC circuit analysis, but the base bias circuit it helps analyze suffers from poor stability due to β variation, which can cause large swings in collector current and voltage.

📌 Key points (3–5)

  • What the Ebers-Moll model is: a simplified BJT model using an ideal diode for the base-emitter junction and a current-controlled current source at the collector-base, sufficient for DC and low-frequency analysis.
  • Proper biasing requirement: collector-base must be reverse-biased and base-emitter forward-biased (V_C > V_B > V_E).
  • Analysis strategy: start with the base-emitter loop (where V_BE is known, ~0.7 V for silicon) to find base current, then use β to find collector current and solve the collector-emitter loop.
  • Common confusion: β is not constant—it varies device-to-device, with temperature, collector current, and collector-emitter voltage; a 10:1 range is possible.
  • Major stability problem: base bias circuits lack stability because fixed base current combined with varying β causes large changes in collector current and voltage, and can create thermal runaway.

🔧 The Ebers-Moll model structure

🔧 Model components

The simplified Ebers-Moll model: uses an ideal diode to model the base-emitter junction and a current-controlled current source located at the collector-base.

  • The model is "sufficient to achieve good analysis results with a variety of DC and low frequency circuits."
  • It is a functional approximation, not a complete physical model.
  • The excerpt emphasizes that β varies with multiple factors, so the model's accuracy depends on knowing the operating conditions.

⚡ Biasing requirements

  • To properly bias the BJT: make the collector-base reverse-biased and the base-emitter forward-biased.
  • In other words: V_C > V_B > V_E.
  • Example: place emitter at ground, modest DC source in base-emitter loop, higher DC source at collector.

🔌 Common emitter configuration

  • When the emitter is at ground (the common point), the circuit has a "common emitter configuration."
  • The specific circuit with resistors in base and collector loops is called "base bias."
  • Resistors serve to limit the transistor's currents and voltages.

📐 Circuit analysis using the model

📐 Base-emitter loop analysis

  • Start with the base-emitter loop because the forward-biased base-emitter junction has a known potential (approximately 0.7 V for silicon).
  • KVL for the base-emitter loop: V_BB = V_RB + V_BE
  • Expanding with Ohm's law: V_BB = I_B × R_B + V_BE
  • Solving for base current: I_B = (V_BB − V_BE) / R_B

Why start here: The base-emitter voltage is known, unlike the collector-emitter voltage which depends on other circuit elements.

📐 Collector-emitter loop analysis

  • After finding base current, use β to find collector current: I_C = β × I_B
  • KVL for the collector-emitter loop: V_CC = V_RC + V_CE
  • Expanding: V_CC = I_C × R_C + V_CE
  • Solving for collector-emitter voltage: V_CE = V_CC − I_C × R_C

Don't confuse: The collector-emitter loop includes the reverse-biased collector-base junction, so V_CE is initially unknown and must be calculated after finding I_C.

🧮 Worked scenario

Example from the excerpt: V_BB = 10 V, V_CC = 15 V, R_B = 200 kΩ, R_C = 1 kΩ, β = 100, silicon transistor.

Step 1 - Base current:

  • Voltage across R_B: 10 V − 0.7 V = 9.3 V
  • I_B = 9.3 V / 200 kΩ = 46.5 μA

Step 2 - Collector current:

  • I_C = 100 × 46.5 μA = 4.65 mA

Step 3 - Collector-emitter voltage:

  • V_CE = 15 V − (4.65 mA × 1 kΩ) = 10.35 V

Additional values: V_RC = 4.65 V, V_CB = 9.65 V, I_E = 4.6965 mA.

🔄 Single supply conversion

  • The two-supply circuit can be redesigned for a single supply by keeping base current unchanged.
  • If V_BB changes from 10 V to 15 V, voltage across R_B becomes 14.3 V.
  • New R_B = 14.3 V / 46.5 μA = 307.5 kΩ.
  • Collector-emitter loop remains unchanged because base current (and thus collector current) stays the same.

⚠️ Stability problems with base bias

⚠️ β variation impact

  • The major problem: base bias circuits lack stability of collector current and collector-emitter voltage.
  • β can vary widely at a given operating point; adding temperature and other factors creates a possible 10:1 range.
  • Example: if β doubles from 100 to 200 in the worked scenario:
    • Base-emitter loop unchanged (I_B still 46.5 μA)
    • Collector current doubles: I_C = 9.3 mA
    • Voltage drop across R_C increases to 9.3 V
    • V_CE drops to 5.7 V

Production consequences: A typical production run might exhibit collector currents from less than 4 mA to more than 10 mA.

🌡️ Thermal runaway mechanism

  • When the circuit is powered on, the BJT warms up as it dissipates power.
  • From the data sheet: β increases with increasing temperature.
  • Because I_B is fixed, any rise in β means I_C must also rise.
  • Increased current causes further rise in power dissipation and temperature.
  • This causes further increase in β, creating an inadvertent thermal positive feedback loop.
  • Danger: Left unchecked, devices could overheat and be destroyed.

Observable effect: In the lab, you could watch I_C slowly rise on an ammeter after turning on power.

💡 Application example - LED brightness

  • Suppose an LED is placed in series with R_C.
  • LED brightness depends on its current level.
  • With base bias, brightness will depend on the β of the specific BJT used.
  • In a larger display made up of similar circuits, illumination will be uneven, causing the entire display to appear "off kilter."

🚫 Physical limits and impossible calculations

  • What if β increases to very high values (e.g., 400)?
  • Calculated I_C would be 46.5 μA × 400 = 18.6 mA.
  • Problem: This current would develop 18.6 V across the 1 kΩ R_C, but V_CC is only 15 V.
  • This is impossible unless the BJT magically becomes a 3.6 V battery (which won't happen).
  • Implication: There must be a maximum collector current limit, leading to the concept of saturation.

📊 DC load lines

📊 Load line definition and equation

DC load line: a plot of all possible coordinate pairs of I_C and V_CE for a transistor in a given circuit.

  • Starting from V_CE = V_CC − I_C × R_C, solve for I_C:
  • I_C = (1/R_C) × (V_CC − V_CE)
  • Rearranging: I_C = −(1/R_C) × V_CE + V_CC / R_C
  • This is a linear equation of the form y = mx + b.

📊 Key load line parameters

ParameterValueMeaning
I_C(sat)V_CC / R_CY-intercept; maximum collector current when V_CE = 0; transistor is saturated
V_CE(cutoff)V_CCX-intercept; largest possible V_CE when I_C = 0; current is cut off
Slope−1 / R_CNegative slope of the load line

📍 Q point (quiescent point)

  • The operating point for a specific transistor on the load line.
  • Associated device current and voltage are called I_CQ and V_CEQ.
  • All possible Q points lay on the load line.

Example from the excerpt (V_CC = 15 V, R_C = 1 kΩ):

  • I_C(sat) = 15 mA
  • V_CE(cutoff) = 15 V
  • Q point for β = 100: I_C = 4.65 mA, V_CE = 10.35 V
  • Q point for β = 200: I_C = 9.3 mA, V_CE = 5.7 V

🔒 Saturation constraint

  • If calculated collector current exceeds saturation current, the actual current will be the saturation current maximum.
  • For the example circuit, any calculated value greater than 15 mA indicates the transistor would produce only 15 mA.
  • This resolves the "impossible" β = 400 scenario: the transistor saturates at 15 mA, not 18.6 mA.
26

DC Load Lines

4.6 DC Load Lines

🧭 Overview

🧠 One-sentence thesis

The DC load line graphically shows all possible operating points for a BJT in a given circuit, defining the limits of collector current and collector-emitter voltage and revealing whether the transistor will saturate or cut off.

📌 Key points (3–5)

  • What the load line plots: all possible coordinate pairs of collector current (I_C) and collector-emitter voltage (V_CE) for a transistor in a specific circuit.
  • Two key limits: saturation current I_C(sat) when V_CE = 0, and cutoff voltage V_CE(cutoff) when I_C = 0.
  • The Q point (quiescent point): the actual operating point for a specific transistor, determined by beta and base current; all possible Q points lie on the load line.
  • Common confusion: calculated currents above saturation are impossible—the transistor cannot produce more than I_C(sat), and V_CE cannot exceed V_CC.
  • Why saturation matters: forcing saturation makes operation stable and independent of beta variations, useful for switching applications like LED drivers.

📐 Understanding the DC load line equation

📐 Deriving the load line

The load line comes from the basic voltage equation for a BJT circuit:

  • Start with: V_CE = V_CC − I_C × R_C
  • Rearrange to solve for I_C: I_C = (−1/R_C) × V_CE + V_CC/R_C
  • This is a linear equation of the form y = mx + b, where:
    • y-axis = I_C (collector current)
    • x-axis = V_CE (collector-emitter voltage)
    • Slope = −1/R_C
    • y-intercept = V_CC/R_C
    • x-intercept = V_CC

🔢 The two intercepts define operating limits

InterceptValuePhysical meaningTransistor state
y-intercept (V_CE = 0)I_C(sat) = V_CC/R_CMaximum possible collector currentSaturated
x-intercept (I_C = 0)V_CE(cutoff) = V_CCMaximum possible collector-emitter voltageCut off
  • Don't confuse: these are limits, not typical operating points; actual operation (the Q point) lies somewhere on the line between them.

🎯 The Q point and beta dependence

🎯 What is the Q point

The quiescent point (Q point): the actual operating point for a specific transistor, with associated current I_CQ and voltage V_CEQ.

  • All possible Q points for different transistors (or different betas) lie on the same load line.
  • The Q point shifts along the line depending on beta and base current.

🔄 How beta affects the Q point

The excerpt gives a concrete example:

  • Circuit parameters: V_CC = 15 V, R_C = 1 kΩ
  • Load line limits: I_C(sat) = 15 mA, V_CE(cutoff) = 15 V
  • For beta = 100: I_C = 4.65 mA, V_CE = 10.35 V
  • For beta = 200: I_C = 9.3 mA, V_CE = 5.7 V

Key observation: higher beta moves the Q point upward along the load line (more current, less voltage across the transistor).

⚠️ The impossible current problem

The excerpt describes what happens with very high beta (e.g., beta = 400):

  • Naive calculation: I_C = 46.5 μA × 400 = 18.6 mA
  • Problem: this would require 18.6 V across R_C, but V_CC is only 15 V
  • Reality: the transistor cannot produce more than I_C(sat) = 15 mA
  • Don't confuse: calculated values above saturation do not mean the transistor will produce that current; it will be limited to I_C(sat).

🔌 Saturation details and practical limits

🔌 Real saturation voltage

The excerpt notes that V_CE does not actually reach zero in saturation:

  • Typical V_CE(sat) is about 0.1 V for small signal devices
  • Precise values can be found in device datasheets (e.g., the "Collector Saturation Region" graph)
  • Example from the excerpt: if I_C = 10 mA and I_B = 0.3 mA, then V_CE(sat) ≈ 0.15 V

🛡️ Why saturation is useful for stability

The excerpt explains a key advantage:

  • Saturation is a fixed value, inherently stable
  • Beta no longer matters in saturation—it is "forced to drop to whatever value is needed to produce I_C(sat)"
  • Strategy: design so that even the smallest expected beta is large enough to cause saturation
  • This solves the thermal feedback problem mentioned earlier (where rising temperature increases beta, which increases current, which increases temperature, etc.)

🔀 Saturating switch applications

🔀 The LED driver circuit concept

The excerpt describes a saturating LED driver as a practical application:

  • Purpose: offload current demand from a logic circuit or microcontroller that can only supply limited current (e.g., 5 mA) when the LED needs more (e.g., over 10 mA)
  • How it works: the logic circuit supplies only base current, not LED current; the BJT acts as a switch to complete the circuit between the DC supply, LED, and current-limiting resistor R_C

💡 Positive logic driver operation

Circuit behavior (Figure 4.19 in the excerpt):

  • Logic input = 0 V: no base current → no collector current → LED off (BJT in cutoff)
  • Logic input = high: voltage drops across R_B (except for V_BE) → creates I_B → BJT saturates → LED turns on
  • Design rule: ratio of saturation current to base current should be much less than beta (a value around 10 guarantees hard saturation)

🔄 Negative logic driver (inverted)

The excerpt mentions a PNP version (Figure 4.20):

  • Logic low turns on the LED
  • Logic high turns off the LED
  • This inverts the logic compared to the NPN version

📋 Worked example from the excerpt

Given: logic "on" voltage = 5 V, V_LED = 1.8 V, V_CE(sat) = 0, R_B = 4.7 kΩ, R_C = 330 Ω, V_CC = 5 V

Step 1: Find base current:

  • I_B = (V_logic − V_BE) / R_B
  • I_B = (5 V − 0.7 V) / 4.7 kΩ = 915 μA

Step 2: Find saturation current (this will be the LED current):

  • I_C(sat) = (V_CC − V_LED) / R_C
  • I_C(sat) = (5 V − 1.8 V) / 330 Ω = 9.7 mA

Step 3: Check the ratio:

  • Ratio = I_C(sat) / I_B ≈ 10:1
  • This guarantees hard saturation

🔧 Broader applications

The excerpt concludes:

  • Saturating switches can be used in many applications where relays might traditionally be used
  • The key advantage is stability: operation does not depend on beta variations
27

BJT Switching and Driver Applications

4.7 BJT Switching and Driver Applications

🧭 Overview

🧠 One-sentence thesis

By intentionally operating a BJT in saturation, designers can create stable switching and driver circuits that are immune to beta variation and reliably control loads like LEDs and motors.

📌 Key points (3–5)

  • Why saturation matters: Saturation is a fixed, stable state where beta no longer affects collector current, solving the problem of performance variance caused by beta variation.
  • Saturating vs non-saturating drivers: Saturating switches require more base current but are stable and simple; non-saturating drivers need less input current but dissipate more power and require higher supply voltages.
  • Common confusion: In saturation, beta doesn't disappear—it's forced to drop to whatever value is needed to produce the saturation current; you just need to ensure even the smallest beta is large enough.
  • Practical applications: BJT switches can replace relays for controlling LEDs, motors (via pulse width modulation), and other loads with advantages of small size, no moving parts, and fast switching.
  • Protection requirements: Inductive loads like motors require snubbing diodes to prevent damaging voltage spikes when the transistor switches off.

🔄 Understanding Saturation in Switching

🔄 The beta variation problem

  • Beta variation causes collector current changes, leading to performance issues.
  • Example: When driving an LED, beta variation causes brightness to vary unpredictably.
  • This instability makes circuits unreliable across different transistors or temperature conditions.

🔒 How saturation solves the problem

Saturation is a fixed value that is inherently stable, and beta no longer matters.

  • When a BJT saturates, beta is effectively forced to drop to whatever value is needed to produce the saturation current.
  • The key design rule: make sure even the smallest beta is large enough to cause saturation.
  • A ratio of saturation current to base current of about 10:1 guarantees hard saturation.
  • The collector-emitter voltage in saturation (V_CE(sat)) is very low, typically around 0.1 to 0.15 V for small signal devices, not quite zero.

💡 Saturating LED Driver Circuits

💡 Positive logic saturating driver

The circuit offloads current demand from logic gates or microcontrollers that can only supply limited current (e.g., 5 mA) when more current is needed (e.g., over 10 mA).

How it works:

  • Logic input zero → no base current → no collector current → LED off (BJT in cutoff).
  • Logic input high → voltage drops across base resistor R_B (except for V_BE) → creates base current I_B.
  • If properly designed, this base current saturates the BJT, which acts as a switch completing the circuit between DC supply, LED, and current limiting resistor R_C.
  • The logic circuit only needs to supply base current, not LED current.

🔄 Negative logic saturating driver

  • Uses a PNP transistor instead of NPN.
  • Logic low turns on the LED; logic high turns it off.
  • Inverts the logic compared to the positive version.

📐 Design example calculation

For a circuit with 5 V logic, V_LED = 1.8 V, and V_CE(sat) = 0:

Base current:

  • I_B = (V_logic - V_BE) / R_B
  • I_B = (5 V - 0.7 V) / 4.7 kΩ = 915 μA

LED current (saturation current):

  • I_C(sat) = (V_CC - V_LED) / R_C
  • I_C(sat) = (5 V - 1.8 V) / 330 Ω = 9.7 mA

The ratio is just over 10:1, guaranteeing hard saturation.

🔌 Motor Control and Protection

🔌 Pulse width modulation motor drive

  • Motor speed depends on average voltage applied.
  • Instead of continuously variable voltage, apply fast pulses of varying width.
  • Narrow, widely-spaced pulses → low average → slow motor speed.
  • Wide, closely-spaced pulses → high average → fast motor speed.
  • Pulses are fast enough that motor inertia keeps it running smoothly rather than starting and stopping.

⚡ Snubbing diode protection

A snubbing diode (also called commutating diode, clamp diode, or flyback diode) prevents damaging transient spikes across the switching transistor.

Why it's needed:

  • When BJT is on, current flows through motor armature (a large coil with high inductance).
  • When BJT turns off, current through inductor cannot change instantaneously.
  • Motor winding generates large flyback voltage (inductive kick) of opposite polarity.
  • Via KVL, this voltage appears across the BJT collector-emitter and could damage it.

How it works:

  • Snubbing diode effectively short-circuits the winding when voltage polarity reverses, preventing the spike.
  • Rest of the time, diode is reverse-biased and out of the circuit.

🎯 Non-Saturating Driver Circuits

🎯 How non-saturating drivers differ

Advantages:

  • Requires less current from the logic circuit.

Disadvantages:

  • Higher transistor power dissipation.
  • Requires a DC source higher than the logic level.

⚙️ Operation principle (bootstrapping)

  • Logic level zero → no base-emitter voltage rise → collector current is zero.
  • Logic level high → via KVL, all logic voltage drops across emitter resistor R_E (except V_BE).
  • This creates I_E, which is virtually the same as I_C (and I_LED).
  • The circuit "programs" emitter current via the resistor and logic voltage, making it fixed and stable.

The emitter voltage is "bootstrapped" to within 0.7 volts of the logic input level, keeping it stable.

Key stability feature:

  • If beta varies, it causes an inverse change in base current with no change in collector current.
  • Beta is not used in the calculation: I_C = (V_logic - V_BE) / R_E
  • Higher beta simply leads to lower base current.

📊 Comparison table

FeatureSaturating DriverNon-Saturating Driver
Input current requiredHigherLower
Transistor power dissipationLowerHigher
Supply voltage requirementSame as logic level OKMust be higher than logic
Beta dependenceNone (in saturation)None (bootstrapped)
V_CEVery low (~0.1-0.15 V)Higher (not saturated)

🔋 Zener Follower Voltage Regulation

🔋 Improvement over simple Zener regulator

Previous Zener diode regulators were inefficient because they drew significant current even under light load. The Zener Follower solves this problem.

⚡ How the Zener Follower works

Circuit structure:

  • Input: positive rectified and filtered output from AC-to-DC power supply.
  • Zener diode is reverse-biased via resistor R.
  • Current flows down through R and into the Zener, which presents fixed potential V_Z.

Voltage relationships:

  • Difference between input voltage and V_Z drops across R and V_CB.
  • Output voltage at BJT emitter = V_Z - V_BE.
  • Both V_Z and V_BE are fixed, stable potentials → output is fixed and stable.
  • Since V_CE = V_CB + V_BE, any input variation (e.g., ripple) drops across the BJT.

🎯 Efficiency advantages

  • Zener diode current is kept low → modest power dissipation.
  • Current draw from input circuit directly reflects load current demand.
  • Low load current → very little current through transistor and from input circuit.
  • Makes for a more efficient system compared to simple Zener regulation.

Don't confuse: The Zener Follower is not just a Zener regulator with a transistor added—the transistor actively buffers the load from the Zener, allowing the Zener to operate at low current while the transistor supplies the load current as needed.

28

BJT Biasing

Chapter 5: BJT Biasing

🧭 Overview

🧠 One-sentence thesis

BJT amplifiers require DC biasing to establish proper forward-reverse bias conditions and overcome the base-emitter voltage threshold, ensuring stable operation despite variations in transistor parameters like β.

📌 Key points (3–5)

  • Why biasing is necessary: Without DC bias, an AC signal alone cannot maintain the required 0.7V forward bias on the base-emitter junction throughout the entire signal cycle.
  • The β variation problem: The current gain β varies with temperature, collector-emitter voltage, and other factors, which can lead to circuit instability.
  • Goal of biasing circuits: Establish proper forward-bias of base-emitter and reverse-bias of collector-base junctions while maintaining stability.
  • Common confusion: Simply applying an AC signal to the base won't work—only the portion exceeding 0.7V would forward-bias the junction, cutting off the negative half of the signal.
  • Design challenge: Various circuit topologies exist to bias the BJT while minimizing sensitivity to parameter variations.

🔌 Why BJTs Need DC Biasing

⚡ The fundamental requirement

  • A BJT requires two specific conditions to operate properly:
    • Forward-bias of the base-emitter junction
    • Reverse-bias of the collector-base junction
  • Current gain (β) is an outgrowth of this forward-reverse bias configuration.
  • Without an additional source of energy beyond the AC signal, amplification cannot occur.

🚫 Why AC alone fails

The base-emitter junction needs approximately 0.7 volts to achieve forward-bias.

  • If only an AC signal were applied to the base:
    • Forward-bias would only occur when the signal exceeded 0.7 volts
    • The entire negative half of the AC signal would fail to forward-bias the junction
    • Result: severe distortion and loss of half the signal

Example: An AC signal swinging from +1V to -1V would only forward-bias during the portion above +0.7V, completely cutting off during the negative swing and most of the positive swing below 0.7V.

🔧 The Stability Challenge

📉 β variation issues

The excerpt emphasizes that β (current gain) is problematic because:

  • It can have a considerable impact on circuit operation
  • It varies with multiple factors:
    • Changes in temperature
    • Collector-emitter voltage changes
    • Other environmental and operating conditions
  • This variation can lead to circuit instability

🎯 Design goal

  • The chapter investigates various circuit topologies to bias the BJT
  • The consistent focus: designing with an eye toward stability
  • Different biasing configurations offer different trade-offs in managing β variation

📊 Chapter Scope

🛠️ What will be covered

The chapter addresses several key topics:

TopicPurpose
DC biasing circuitsSolve for device currents and voltages
DC load linesPlot operating points for various biasing circuits
Stability methodsIncrease circuit stability against transistor parameter variation
Circuit topologiesInvestigate variety of biasing configurations

🔍 The analytical approach

  • Solve various BJT biasing circuits for:
    • Device currents
    • Device voltages
  • Use DC load line analysis as a graphical tool
  • Compare methods for their stability characteristics
29

BJT Biasing

5.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

BJT amplifiers require DC biasing to establish proper forward-reverse junction conditions and overcome the base-emitter voltage threshold, ensuring stable amplification despite variations in transistor parameters like β.

📌 Key points (3–5)

  • Why biasing is needed: a BJT requires forward-bias of the base-emitter junction and reverse-bias of the collector-base junction to operate; without DC biasing, AC signals alone cannot reliably forward-bias the base-emitter (which needs ~0.7 V).
  • The β variation problem: current gain β varies with temperature, collector-emitter voltage, and other factors, leading to circuit instability.
  • What this chapter covers: various circuit topologies for biasing BJTs, methods to solve for device currents and voltages, DC load lines, and techniques to increase stability against transistor parameter variation.
  • Common confusion: thinking that current gain alone (β) is enough to amplify an AC signal—amplification requires an additional energy source and proper bias conditions, not just β.

🔌 Why BJTs need biasing

🔌 The junction bias requirement

A bipolar junction transistor requires forward-bias of the base-emitter junction and reverse-bias of the collector-base junction in order to operate properly.

  • Current gain (β) is an outgrowth of this forward-reverse bias condition.
  • Without proper bias, the transistor cannot amplify.
  • Don't confuse: β is not a standalone property—it depends on the bias state.

⚡ The energy and threshold problem

  • Energy source: Without an additional source of energy beyond the input signal, amplification cannot be produced.
  • Voltage threshold: The base-emitter junction needs approximately 0.7 volts to forward-bias.
  • If you simply apply an AC signal to the base without DC bias:
    • Only the portion of the AC signal that exceeds 0.7 V will forward-bias the base-emitter.
    • The entire negative half of the AC signal will not forward-bias the junction at all.
  • Example: An AC signal swinging ±0.5 V around 0 V will never reach 0.7 V, so the transistor will not turn on.

🔄 Why not just use AC alone?

  • The excerpt emphasizes that applying an AC signal directly to the base (hoping to get an amplified version at the collector) will fail because:
    • The signal must first overcome the 0.7 V barrier.
    • Without DC bias, much of the AC waveform (especially negative portions) will be cut off.
  • DC biasing "lifts" the operating point so that the AC signal can swing around a stable, forward-biased condition.

🌡️ The stability challenge

🌡️ β variation and instability

  • The current gain β is a prime BJT parameter, but it is not constant.
  • β varies with:
    • Changes in temperature.
    • Changes in collector-emitter voltage.
    • Other operating conditions.
  • This variation can lead to circuit instability: the operating point drifts, and the amplifier may not work reliably.

🛠️ Design goal: stability

  • The chapter investigates a variety of circuit topologies "always with an eye toward stability."
  • Stability means designing circuits that maintain consistent operation even when β and other parameters change.
  • Methods to increase stability will be discussed (specific techniques are not detailed in this excerpt, but the goal is stated).

📐 What you will learn

📐 Chapter objectives

The excerpt lists four learning objectives:

ObjectiveWhat it means
Explain the need for DC biasingUnderstand why BJTs require DC bias (junction conditions, threshold, energy source).
Solve various BJT biasing circuitsCalculate device currents and voltages for different circuit topologies.
Plot DC load linesGraphically represent the operating point and constraints for BJT circuits.
Discuss stability methodsLearn techniques to reduce sensitivity to transistor parameter variation (especially β).

🔍 Scope of the chapter

  • The chapter will cover a variety of circuit topologies for biasing BJTs.
  • Each topology will be analyzed for:
    • How it sets the operating point.
    • How stable it is against parameter changes.
  • The excerpt does not provide specific circuit examples yet, but it sets the stage for detailed analysis.

🧩 Key takeaways

🧩 Biasing is essential

  • You cannot simply apply an AC signal to a BJT and expect amplification.
  • Proper DC biasing ensures:
    • The base-emitter is forward-biased (~0.7 V).
    • The collector-base is reverse-biased.
    • The transistor operates in the active region where β is meaningful.

🧩 β is variable, not fixed

  • Don't treat β as a constant in real circuits.
  • Temperature and voltage changes cause β to vary, so circuits must be designed to tolerate this variation.

🧩 Stability is a design priority

  • The chapter emphasizes stability throughout: every biasing topology will be evaluated for how well it handles β variation and other parameter changes.
  • Example: A circuit that works perfectly at 25°C but fails at 125°C is not stable; the chapter will teach how to avoid this.
30

Introduction to Insulated Gate Bipolar Transistors (IGBTs)

5.1 Introduction

🧭 Overview

🧠 One-sentence thesis

The IGBT is a power semiconductor switch that combines the low on-state loss and high current/voltage handling of BJTs with the voltage-controlled ease of driving found in MOSFETs, making it a competitive choice for many power switching applications despite being slower and more costly than current-generation power MOSFETs.

📌 Key points (3–5)

  • What the IGBT is: a power semiconductor device introduced commercially in the 1980s, designed primarily as a high voltage/high current switch (not for linear applications like audio amplifiers).
  • Hybrid performance: combines BJT advantages (low on-state power loss, high current/voltage capability) with MOSFET advantages (voltage-controlled, easier to drive than current-controlled devices).
  • Trade-offs: not as fast as modern power MOSFETs and more expensive than both power BJTs and power MOSFETs.
  • Common confusion: choosing between IGBT, power BJT, and power MOSFET depends on application specifics—no single device is always best.
  • Market position: has largely replaced older thyristor devices (e.g., SCR) in many areas due to speed and simpler driving circuits.

🔌 What the IGBT is and its history

🔌 Device definition and purpose

Insulated Gate Bipolar Transistor (IGBT): a power semiconductor designed to be used as a high voltage/high current switch.

  • First became commercially available in the 1980s.
  • Initial devices had performance issues; subsequent generations have largely resolved these problems.
  • Today in wide use, competing with power BJTs and power E-MOSFETs across a range of applications.

🚫 What the IGBT is not used for

  • Typically not used for linear applications such as audio class B power amplifiers.
  • The excerpt emphasizes it is a switch, not a linear device.
  • Don't confuse: the IGBT's design focus is switching, not amplification in the linear region.

⚖️ Hybrid characteristics: combining BJT and MOSFET strengths

⚖️ Advantages inherited from power BJTs

  • Low on-state power loss: when the device is conducting, it dissipates less power.
  • High current and voltage handling: can manage large currents and voltages effectively.

⚖️ Advantages inherited from power E-MOSFETs

  • Voltage-controlled device: easier to drive than current-controlled devices.
  • The excerpt contrasts this with BJTs, which are current-controlled and require more complex drive circuits.
  • Example: a driver circuit for an IGBT is simpler than one for a BJT because it controls via voltage rather than supplying continuous base current.

⚠️ Trade-offs and limitations

AspectIGBT characteristicComparison
SpeedNot as fast as current-generation power E-MOSFETsSlower switching times
CostMore costly than both power BJT and power E-MOSFETHigher price point
  • These trade-offs mean the IGBT is not always the best choice; selection depends on design specifics.

🏆 Market position and application context

🏆 Replacement of older technologies

  • The IGBT has overtaken older thyristor devices (e.g., SCR) in many areas.
  • Two reasons given:
    • Speed: faster than thyristors.
    • Simpler driving circuits: easier to control than thyristors.

🏆 Choosing between IGBT, power BJT, and power MOSFET

  • The excerpt states that the choice depends on the specifics of the design.
  • Example scenario mentioned (incomplete in excerpt): "a medium to high power design that focuses on lowest [cost/loss/etc.]" would guide the choice.
  • Don't confuse: there is no universal "best" device—each has strengths suited to different applications.

🔍 Context from surrounding material

🔍 Preceding content (PWM amplifiers)

The excerpt includes review questions and problems from a previous section on PWM (pulse width modulation):

  • PWM vs PDM (pulse density modulation)
  • Shoot-through and dead time concepts
  • Output LC filters and device capacitance effects
  • Half-bridge and full-bridge configurations
  • Effects of on-resistance (r_DS(on)) and drive current capacity

Note: These topics are not part of the IGBT introduction itself; they belong to the prior chapter on PWM amplifiers.

🔍 What comes next

The excerpt indicates that subsequent sections will cover:

  • Basic operation of the IGBT
  • Internal structure of the IGBT
  • Differences between PT (Punch-Through) and NPT (Non-Punch-Through) IGBTs
  • Detailed comparisons with power BJTs and power MOSFETs
  • Data sheet parameter interpretation
  • Basic IGBT power control circuits
31

The Need For Biasing

5.2 The Need For Biasing

🧭 Overview

🧠 One-sentence thesis

DC biasing is essential for transistor amplifiers because AC signals alone can drive the transistor into cutoff or ignore portions of the signal, so a stable DC bias (Q point) must be established to maintain proper transistor function and AC performance.

📌 Key points (3–5)

  • Why biasing is needed: AC signals swing both positive and negative, but transistors require specific voltage thresholds (e.g., 0.7 V for silicon) to turn on; without DC bias, negative portions or small signals would be ignored.
  • The biasing solution: Apply a DC voltage to the transistor and superimpose the AC signal on top, so even the negative AC peaks remain above the transistor's threshold.
  • Stability matters: A stable Q point (one that doesn't shift when parameters like β change) is critical; unstable biasing causes gain instability, increased distortion, or reduced output power.
  • Common confusion: Don't confuse "any DC voltage" with "stable DC voltage"—the goal is a Q point that remains fixed despite parameter variations like changes in β.
  • Two-supply emitter bias advantage: By making R_E much larger than R_B/β, collector current becomes nearly independent of β, achieving high stability.

🔌 Why transistors need DC bias

⚡ The AC signal problem

  • AC signals alternate between positive and negative voltages.
  • Transistors have threshold requirements: for example, a silicon transistor needs about 0.7 volts to turn on.
  • Problem 1: The negative half of an AC signal would try to reverse-bias the transistor, potentially driving it into cutoff.
  • Problem 2: If the entire AC signal is below the 0.7 V threshold, the transistor ignores it completely.

Example: A microphone generates only a few hundred millivolts. Without bias, this entire signal could fall below 0.7 V and be ignored.

🎯 The superposition solution

  • Apply a large DC voltage to the transistor first.
  • Then add (superimpose) the AC signal on top of this DC level.
  • Result: Even when the AC signal swings negative, the net voltage (DC + AC) remains positive and above the threshold.
  • This maintains proper transistor function throughout the entire AC cycle.

🎚️ Stability and the Q point

📍 What is the Q point

Q point (quiescent point): the DC operating point of the transistor, defined by collector current (I_C) and collector-emitter voltage (V_CE).

  • "Quiescent" means at rest, i.e., with no AC signal applied.
  • The Q point sets the baseline around which the AC signal swings.

⚠️ Why stability matters

  • An unstable Q point shifts when transistor parameters (especially β) change.
  • Consequences of instability:
    • Gain becomes unstable
    • Distortion increases
    • Output power is reduced
  • Goal: Establish a Q point that doesn't move despite parameter changes.

🔧 The base bias problem

  • The excerpt mentions that base bias (examined in a prior chapter) has major stability problems.
  • The desired circuit establishes collector current that does not shift even when β changes.

🔋 Two-supply emitter bias configuration

🏗️ Circuit structure (NPN version)

  • Collector: connected to positive supply (V_CC) through resistor R_C (highest potential).
  • Base: tied to ground through resistor R_B (intermediate potential).
  • Emitter: connected to negative supply (V_EE) through resistor R_E (lowest potential).
  • This arrangement ensures:
    • Collector-base junction is reverse-biased ✓
    • Base-emitter junction is forward-biased ✓

📐 Collector current equation

Applying KVL to the base-emitter loop yields:

I_C = (|V_EE| − V_BE) / (R_E + R_B/β)

  • The absolute value notation avoids confusion about the negative supply sign.
  • Key insight: β only partly determines collector current.

🎯 Achieving stability

  • Condition: Make R_E much greater than R_B/β.
  • In practice: R_E approximately equal to or larger than R_B achieves this (given typical β values).
  • Simplified equation when R_E >> R_B/β:

I_C ≈ (|V_EE| − V_BE) / R_E

  • Now β plays virtually no role in determining I_C.
  • Almost all of the emitter supply voltage drops across R_E to establish a stable I_C.
  • If β changes, base current (I_B) changes inversely, but I_C remains largely unchanged.

📊 Other circuit parameters

Once I_C is known, find other values using Ohm's law and KVL:

ParameterFormulaNotes
V_C (collector to ground)V_CC − I_C·R_CVoltage at collector node
V_CEV_CC + |V_EE| − I_C(R_C + R_E)Transistor's collector-emitter voltage
V_E (emitter voltage)−I_B·R_B − V_BENegative value for NPN with negative emitter supply

📈 DC load line endpoints

The load line provides a "sanity check" for calculations:

  • Saturation current (when V_CE = 0): I_C(sat) = (V_CC + |V_EE|) / (R_C + R_E)
    • All supply voltage drops across the resistors.
  • Cutoff voltage (when I_C = 0): V_CE(cutoff) = V_CC + |V_EE|
    • No current means no voltage drop across resistors; V_CE absorbs entire supply.

✅ Stability verification example

  • Doubling β from 100 to 200 produces only about 1% change in collector current.
  • The Q point barely moves despite a 100% change in β.
  • This demonstrates very high stability.

🔄 PNP version and practical considerations

🔃 PNP two-supply emitter bias

  • Common practice: "flip" the entire circuit so emitter is on top, collector on bottom.
  • Advantage: In multi-transistor schematics, all DC bias currents "run down the page."
  • Key differences from NPN:
    • Voltage polarities are reversed (positive becomes negative, vice versa).
    • Current directions are reversed (e.g., current flows out of PNP collector, into NPN collector).
    • Base current flows out of PNP base (vs. into NPN base).
    • Emitter is 0.7 V more positive than base (vs. more negative in NPN).

💡 Design approach

  • Don't memorize all the equations—there are too many variations.
  • Do remember: How to find collector current first.
  • Then apply Ohm's law and KVL to derive whatever else is needed.
  • This approach works across all biasing configurations.

🔀 Voltage divider bias introduction

🏗️ Alternative configuration

Voltage divider bias: a configuration that provides high bias stability by using a voltage divider (R_1 and R_2) off the collector supply to set the base voltage, with the emitter resistor returned to ground instead of a negative supply.

  • Avoids the need for a second (negative) power supply.
  • Base voltage is derived from V_CC through the voltage divider.
  • Emitter resistor connects to ground, and the base voltage is raised above ground.

📐 Key equations

  • Load line endpoints:
    • I_C(sat) = V_CC / (R_C + R_E)
    • V_CE(cutoff) = V_CC
  • Thevenin equivalent of the voltage divider simplifies analysis:
    • V_TH = V_CC · R_2 / (R_1 + R_2)
    • R_TH = R_1 || R_2
  • Collector current: I_C = (V_TH − V_BE) / (R_E + R_TH/β)

🎯 Quick approximation condition

  • If the voltage divider is lightly loaded (base current << divider current), a simpler approach works:
    • Base voltage ≈ divider voltage
    • Subtract 0.7 V to get emitter voltage
    • Voltage across R_E determines I_E ≈ I_C
  • The excerpt hints that examining equation 5.8 reveals when this approximation is valid (text cuts off).
32

Two-Supply Emitter Bias

5.3 Two-Supply Emitter Bias

🧭 Overview

🧠 One-sentence thesis

Two-supply emitter bias achieves a stable Q point (operating point) by using a negative supply on the emitter and grounding the base through a resistor, making collector current nearly independent of transistor beta changes.

📌 Key points (3–5)

  • Why bias is needed: AC signals riding on DC bias prevent the transistor from cutting off during negative signal swings and keep the base-emitter junction properly forward-biased above 0.7 V.
  • The stability problem: Base bias (previous chapter) produces unstable Q points when beta changes, causing gain instability, distortion, or reduced output power.
  • How two-supply emitter bias works: Collector connects to positive supply, base ties to ground through a resistor, and emitter connects to a negative supply to establish forward bias.
  • The key stability mechanism: When the emitter resistor R_E is much larger than R_B divided by beta, collector current becomes almost independent of beta—changes in beta cause inverse changes in base current while collector current stays constant.
  • Common confusion: Don't confuse voltage polarities and current directions between NPN and PNP versions—PNP circuits are often flipped top-to-bottom so DC currents "run down the page," but the underlying equations remain valid with reversed polarities.

🔧 Why DC bias is necessary

🔧 The AC signal problem

  • Many input devices (microphones, sensors) generate signals of only a few hundred millivolts.
  • Without DC bias, the negative half of an AC signal would drive the transistor below the 0.7 V base-emitter threshold, causing the signal to be ignored or cut off.
  • Solution: Apply a DC bias voltage and superimpose the AC signal on top of it.
    • Example: If an AC signal rides on a much larger DC voltage, even the negative peak remains a net positive voltage, maintaining proper transistor function.

⚠️ The stability requirement

A stable bias establishes a Q point (quiescent operating point) that doesn't move despite parameter changes such as changes in beta.

  • An unstable Q point has negative effects on AC amplifier performance:
    • Makes gain unstable
    • Increases distortion
    • Reduces output power
  • Base bias (from the prior chapter) suffers from major instability problems.
  • Goal: Find a circuit that establishes collector current that does not shift even when beta changes.

⚡ Two-supply emitter bias configuration

⚡ How the circuit is arranged (NPN version)

For proper NPN transistor operation:

  • Collector: highest potential (connected to positive supply V_CC)
  • Base: somewhat lower potential (tied to ground through resistor R_B)
  • Emitter: lowest potential (connected to negative supply V_EE)

This arrangement ensures:

  • Collector-base junction is reverse-biased
  • Base-emitter junction is forward-biased

🧮 The collector current equation

Applying Kirchhoff's Voltage Law (KVL) to the base-emitter loop yields:

I_C = (|V_EE| − V_BE) / (R_E + R_B/β)

  • The absolute value notation avoids confusion about the sign of the emitter supply voltage.
  • Key insight: Beta only partly determines collector current.

🎯 Achieving stability through resistor sizing

When R_E is much greater than R_B divided by beta, the equation simplifies to:

I_C ≈ (|V_EE| − V_BE) / R_E

  • This condition is easy to achieve: if R_E is approximately equal to or larger than R_B, the requirement is met (given typical beta values).
  • Result: Almost all of the emitter supply voltage drops across R_E to establish a stable I_C, with beta playing virtually no role.
  • Stability mechanism: If beta changes, base current changes inversely while collector current remains largely unchanged.

📐 Other circuit voltages and the load line

📐 Finding other voltages

Once collector current is known, apply Ohm's law and KVL:

Collector voltage (from collector to ground):

  • V_C = V_CC − I_C × R_C

Collector-emitter voltage:

  • V_CE = V_CC + |V_EE| − I_C × (R_C + R_E)
  • Alternative: V_CE = V_C − V_E

Emitter voltage:

  • V_E = −I_B × R_B − V_BE

📊 DC load line endpoints

The load line serves as a "sanity check" for computations.

EndpointConditionFormulaMeaning
I_C(sat)V_CE = 0(V_CC + |V_EE|) / (R_C + R_E)Maximum current when all supply voltage drops across resistors
V_CE(cutoff)I_C = 0V_CC + |V_EE|Maximum voltage when no current flows through resistors

Don't memorize all equations—remember how to find collector current, then apply Ohm's law and KVL to derive whatever else is needed.

🔄 PNP version and practical considerations

🔄 Creating the PNP circuit

  • Replace NPN with PNP transistor
  • Change power supply signs
  • Common practice: Flip the entire circuit top-to-bottom so emitter is on top and collector on bottom
    • Advantage: In multi-transistor schematics, all DC bias currents "run down the page"

🔄 Key differences in PNP analysis

All device current equations and component voltage equations remain valid, but:

AspectNPNPNP
Voltage polaritiesWhat was positiveBecomes negative
Current directionsInto collectorOut of collector
Base currentFlows into baseFlows out of base
Base voltageSmall negativeSmall positive
Emitter voltageMore negative (by 0.7 V)More positive (by 0.7 V)

Example: In NPN, base current flows into the base creating a small negative voltage at the base and a more negative voltage at the emitter. In PNP, base current flows out, creating a small positive voltage at the base with the emitter slightly more positive.

✅ Verification of stability

The excerpt demonstrates stability with a numerical example:

  • Doubling beta from 100 to 200 produces only about 1% change in collector current (from 3.38 mA to 3.41 mA)
  • This represents a 1% current change for a 100% change in beta
  • Conclusion: This configuration produces very small Q point changes despite very large beta changes.

💻 Computer simulation expectations

For a properly designed two-supply emitter bias circuit:

  • V_B should be pretty close to 0 V
  • V_E should be about −0.7 volts
  • Collector voltage can be estimated from V_CC minus the drop across R_C
  • Base current is typically small (tens of microamps for typical beta values)
33

Voltage Divider Bias

5.4 Voltage Divider Bias

🧭 Overview

🧠 One-sentence thesis

Voltage divider bias achieves high bias stability by deriving the base voltage from the collector supply through a resistor divider, avoiding the need for a second power supply while maintaining a stable Q point even when transistor beta changes significantly.

📌 Key points (3–5)

  • Core mechanism: the base voltage is raised above ground using a voltage divider (R₁ and R₂) connected to the collector supply, while the emitter resistor returns to ground.
  • Stability advantage: the Q point shifts less than 1% when beta changes by a factor of two, because the divider current is much larger than the base current.
  • Analysis shortcut: if R₂ is not much larger than Rₑ (i.e., Rₑ >> Rₜₕ/β), the divider is "lightly loaded" and a quick approximation (ignoring Rₜₕ/β in the denominator) is accurate.
  • Common confusion: ground-referenced voltages (Vᵦ, Vᴄ, Vₑ) differ between NPN and PNP versions, but component voltages and currents have the same magnitudes—only signs and directions reverse.
  • PNP adaptation: flipping the circuit and offsetting the supply reference allows a positive supply to be used for PNP voltage divider bias, though all single-subscript voltages change because the reference point moves.

🔧 Circuit structure and operation

🔧 How voltage divider bias is built

  • The configuration returns the emitter resistor (Rₑ) to ground instead of to a negative supply.
  • The base voltage is raised by connecting the base to a voltage divider made of R₁ (top) and R₂ (bottom) across the collector supply Vᴄᴄ.
  • This avoids the need for a second power supply (unlike two-supply emitter bias).

🔌 Load line endpoints

The saturation and cutoff points define the limits of operation:

EndpointConditionFormulaMeaning
SaturationVᴄₑ → 0Iᴄ(sat) = Vᴄᴄ / (Rᴄ + Rₑ)Maximum current when collector-emitter voltage collapses
CutoffIᴄ = 0Vᴄₑ(cutoff) = VᴄᴄMaximum voltage when no current flows through Rᴄ and Rₑ
  • At saturation, both Rᴄ and Rₑ limit the current.
  • At cutoff, no potentials drop across Rᴄ or Rₑ, so Vᴄₑ equals the full supply voltage.

🧮 Finding the Q point

🧮 Thevenizing the voltage divider

To simplify analysis, replace R₁ and R₂ with their Thevenin equivalent:

  • Thevenin voltage: Vₜₕ = Vᴄᴄ × R₂ / (R₁ + R₂)
  • Thevenin resistance: Rₜₕ = R₁ || R₂ = (R₁ × R₂) / (R₁ + R₂)

This transforms the divider into a single voltage source and series resistor, making the base-emitter loop easier to analyze.

⚡ Collector current formula

Applying Kirchhoff's voltage law to the base-emitter loop (with the Thevenin equivalent):

Iᴄ = (Vₜₕ − Vᵦₑ) / (Rₑ + Rₜₕ/β)

  • Vₜₕ is the Thevenin voltage from the divider.
  • Vᵦₑ is the base-emitter drop (0.7 V for silicon).
  • The denominator includes Rₑ (emitter resistance) and Rₜₕ/β (base resistance reflected to the emitter side, since Iᵦ = Iᴄ/β).

Example (from Example 5.3):
With Vₜₕ = 4.8 V, Rₑ = 3.3 kΩ, Rₜₕ = 3.2 kΩ, β = 200, and Vᵦₑ = 0.7 V:
Iᴄ = (4.8 − 0.7) / (3.3 + 3.2/200) = 1.236 mA.

🚀 Quick approximation

If the voltage divider is "lightly loaded" (base current << divider current), ignore Rₜₕ/β:

Iᴄ ≈ (Vₜₕ − Vᵦₑ) / Rₑ

  • When to use: as long as Rₑ >> Rₜₕ/β, or equivalently, R₂ is not much larger than Rₑ.
  • Why it works: the divider current is much larger than the base current, so the base draws negligible current and Vᵦ ≈ Vₜₕ.
  • Example: using the same values, Iᴄ ≈ (4.8 − 0.7) / 3.3 = 1.242 mA, very close to the exact 1.236 mA.

📐 Collector-emitter voltage

Once Iᴄ is known:

Vᴄₑ = Vᴄᴄ − Iᴄ(Rᴄ + Rₑ)

  • The supply voltage Vᴄᴄ minus the drops across both Rᴄ and Rₑ.
  • Example: Vᴄₑ = 15 − 1.236 × (3.9 + 3.3) = 6.1 V.

🔍 Base voltage

Two ways to find Vᵦ:

  1. Approximation (if divider is lightly loaded): Vᵦ ≈ Vₜₕ.
  2. More accurate: Vᵦ = Vᵦₑ + Iᴄ × Rₑ (the drop across Rₑ plus the base-emitter drop).

Example: Vᵦ = 0.7 + 1.236 × 3.3 = 4.78 V, close to Vₜₕ = 4.8 V.

🛡️ Stability verification

🛡️ How stable is the Q point?

The excerpt demonstrates stability by recalculating with β cut in half (from 200 to 100):

  • Result: Iᴄ changes from 1.236 mA to 1.23 mA, and Vᴄₑ from 6.1 V to 6.14 V.
  • Shift: less than 1% in both current and voltage.
  • Why: the divider current (~1 mA) is much larger than the base current (~12 µA even after β halves), so changes in base current barely affect the divider voltage.

Don't confuse: high stability does not mean Iᵦ is unchanged—Iᵦ nearly doubles when β halves—but because the divider is lightly loaded, this has minimal impact on the Q point.

🔄 PNP voltage divider bias

🔄 Creating the PNP version

To convert from NPN to PNP:

  1. Replace the NPN transistor with a PNP.
  2. Change the sign of the power supply.
  3. Flip the circuit top-to-bottom so DC current flows down the page.

Result: all currents and component voltages have the same magnitudes, but directions and polarities are reversed.

🔌 Positive supply adaptation

The direct PNP conversion results in ground being the most positive potential and a negative supply at the bottom, which is awkward if a positive supply is used elsewhere.

Solution: add the magnitude of the negative supply voltage to both the ground and power connections—this "offsets" the reference point.

  • What changes: all ground-referenced (single-subscript) voltages change because the reference moves.
  • What stays the same: individual component voltages (e.g., the voltage across Rₑ or Rᴄ) remain unchanged.
  • Example: in the NPN version, Vᵦ is the voltage across R₂; in the offset PNP version, Vᵦ is the voltage across R₁.

🧭 Analysis methods for PNP

The excerpt illustrates two methods for finding the Q point in a PNP circuit (Example 5.4):

Method One (focus on the base-emitter loop):

  1. Find the voltage across R₂ (now on top) using the voltage divider rule: Vᴿ₂ = Vₑₑ × R₂ / (R₁ + R₂).
  2. Subtract Vᵦₑ to get the voltage across Rₑ: Vᴿₑ = Vᴿ₂ − Vᵦₑ.
  3. Use Ohm's law: Iᴄ = Vᴿₑ / Rₑ.

Method Two (ground-referenced voltages):

  1. Find Vᵦ using the divider: Vᵦ = Vₑₑ × R₁ / (R₁ + R₂).
  2. The base-emitter voltage is a rise (+ to −), so Vₑ = Vᵦ + Vᵦₑ.
  3. The voltage across Rₑ is Vᴿₑ = Vₑₑ − Vₑ.
  4. Then Iᴄ = Vᴿₑ / Rₑ.

Both methods yield the same result (Iᴄ = 1.24 mA in the example).

📊 Finding Vᴄₑ in PNP circuits

Two approaches:

  1. Modified formula: Vᴄₑ = −(Vₑₑ − Iᴄ(Rᴄ + Rₑ)). The negative sign indicates the collector is negative relative to the emitter; to avoid this, refer to Vₑᴄ instead.
  2. Voltage difference: Find Vᴄ = Iᴄ × Rᴄ, then Vᴄₑ = Vᴄ − Vₑ.

Example: Vᴄ = 1.24 × 3.9 = 4.84 V, Vₑ = 10.9 V, so Vᴄₑ = 4.84 − 10.9 = −6.06 V.

🔁 Comparing NPN and PNP results

The excerpt compares Example 5.3 (NPN) and Example 5.4 (PNP):

  • Same: device currents (Iᴄ) and component voltage magnitudes (|Vᴄₑ|, voltage across Rₑ).
  • Different: ground-referenced potentials (Vᵦ, Vᴄ, Vₑ) because the reference point and polarities are reversed.

Don't confuse: identical component behavior does not mean identical node voltages—the reference frame matters.

34

Feedback Biasing

5.5 Feedback Biasing

🧭 Overview

🧠 One-sentence thesis

Feedback biasing configurations use negative feedback to partially offset output changes and achieve modest Q-point stability with fewer components than high-stability circuits, though they cannot match the stability of two-supply emitter bias or voltage divider bias.

📌 Key points (3–5)

  • What feedback biasing is: a group of bias configurations that reflect output changes back to the input to partially offset those changes, using fewer components than high-stability designs.
  • How negative feedback works: changes in collector current alter voltages that feed back to the base, causing base current changes that oppose the original collector current change.
  • Three main types: collector feedback bias, emitter feedback bias, and combination feedback bias—each with different feedback paths but similar modest stability.
  • Common confusion: these circuits are simpler but less stable than voltage divider or two-supply emitter bias; the stability condition (making one resistor much larger than another) is harder to achieve in feedback configurations.
  • Trade-off: feedback biasing offers superior stability to simple base bias while using fewer parts than high-stability alternatives.

🔄 Negative Feedback Mechanism

🔄 Core principle of negative feedback

Negative feedback: a technique where a change in the output can be reflected back to the input in such a way that it tends to partially offset the output change.

  • The output change creates a voltage or current change that travels back to the input.
  • This feedback signal opposes (rather than reinforces) the original change.
  • The result is improved stability compared to circuits without feedback.

🎯 Why feedback improves stability

  • Without feedback, parameter changes (like beta increasing due to temperature) directly cause large Q-point shifts.
  • With feedback, the same parameter change triggers a compensating change that reduces the overall shift.
  • Example: if beta increases and tries to raise collector current, feedback reduces base current, partially canceling the increase.

🔌 Collector Feedback Bias

🔌 Circuit structure

  • Takes the basic base bias configuration and moves the base resistor (R_B) connection.
  • Instead of connecting R_B to the power supply, it connects to the lower part of the collector resistor (R_C).
  • This small change creates a feedback path from collector to base.

⚙️ How collector feedback works

The feedback mechanism operates through this sequence:

  1. Assume beta increases (perhaps due to temperature change)
  2. This should increase collector current (I_C)
  3. Increased I_C increases the voltage drop across R_C (by Ohm's law)
  4. By KVL, V_CE = V_C = V_CC - I_C · R_C, so V_C must drop
  5. Key insight: V_C also equals the drop across R_B plus the fixed V_BE (≈0.7V)
  6. Since V_BE is fixed, any decrease in V_C means a decrease in voltage across R_B
  7. By Ohm's law, decreased voltage across R_B means decreased base current (I_B)
  8. Decreased I_B tends to offset the initial tendency of collector current to increase

📐 Collector current equation

The collector current is given by:

I_C = (V_CC - V_BE) / (R_C + R_B/β)

  • This equation resembles those for two-supply emitter bias and voltage divider bias.
  • For good stability, we need R_C >> R_B/β (collector resistance much greater than base resistance divided by beta).
  • Problem: this condition is not nearly as easy to meet in collector feedback circuits.
  • Result: collector feedback has only modest stability.

📏 Load line endpoints

  • Cutoff (maximum V_CE): V_CE(cutoff) = V_CC
  • Saturation (maximum I_C): I_C(sat) = V_CC / R_C

🔍 Stability example

From Example 5.5 with beta = 100 → 50 (halved):

  • Collector current drops from 1.21 mA to 1.05 mA (about 13% reduction)
  • V_CE changes from 2.9 V to 4.5 V (larger change)
  • This is clearly better than base bias but not as stable as two-supply emitter bias or voltage divider bias.

🔄 PNP version

  • Uses the same power supply shifting technique as PNP voltage divider.
  • All currents and component voltages have the same magnitudes but opposite directions and polarities.
  • Ground-referenced voltages differ from NPN counterparts due to the changed reference point.

🔋 Emitter Feedback Bias

🔋 Circuit concept

  • Uses the same overall negative feedback idea as collector feedback.
  • Difference in focus: collector feedback uses collector current establishing V_C via R_C; emitter feedback uses emitter current establishing V_E via R_E.
  • In both cases, these voltages change the voltage across R_B, which changes I_B to oppose the original collector current change.

📐 Emitter current equation

The collector current is:

I_C = (V_CC - V_BE) / (R_E + R_B/β)

  • For stable Q-point despite beta changes, need R_E >> R_B/β.
  • Same problem as collector feedback: this stipulation is not easy to achieve.
  • Consequence: emitter feedback also has only modest stability.

📏 Load line endpoints

  • Cutoff: V_CE(cutoff) = V_CC
  • Saturation: I_C(sat) = V_CC / (R_C + R_E)

🔍 Stability comparison

From Example 5.6 when beta drops from 100 to 50:

  • I_C changes from 6.66 mA to 3.45 mA
  • V_CE changes from 6.68 V to 13.1 V
  • Less than 2:1 change in I_C and V_CE, but stability is not dramatic.

🔀 Combination Feedback Bias

🔀 Dual feedback approach

  • Combines both collector feedback and emitter feedback in one circuit.
  • Applies feedback to R_B "from both ends."
  • Uses feedback from both the collector (via R_C) and emitter (via R_E).

📈 Stability improvement

  • Tends to have slightly better stability than either collector feedback or emitter feedback alone.
  • Still only one resistor shy of the voltage divider circuit, which is considerably more stable.
  • Represents a middle ground in the complexity-stability trade-off.

📐 Equations summary

ParameterEquation
Collector currentI_C = (V_CC - V_BE) / (R_C + R_E + R_B/β)
Collector-emitter voltageV_CE = V_CC - I_C(R_C + R_E)
Saturation currentI_C(sat) = V_CC / (R_C + R_E)
Cutoff voltageV_CE(cutoff) = V_CC

🔄 PNP version available

  • Example 5.7 shows the "upside down" PNP version.
  • Same principles apply with reversed polarities and current directions.

📊 Comparison of Bias Configurations

📊 Stability ranking

From highest to lowest Q-point stability:

  1. Two-supply emitter bias (very high stability)
  2. Voltage divider bias (very high stability)
  3. Combination feedback bias (modest stability)
  4. Collector feedback bias (modest stability)
  5. Emitter feedback bias (modest stability)
  6. Simple base bias (poor stability)

🧩 Complexity vs. stability trade-off

ConfigurationComponent countStabilityKey feature
Two-supply emitter biasMore parts, two suppliesVery highBipolar supplies
Voltage divider biasMore parts, one supplyVery highResistive divider
Feedback configurationsFewest partsModestSingle supply, negative feedback

⚠️ Don't confuse

  • Feedback biasing vs. high-stability biasing: feedback configurations use fewer components but cannot achieve the same stability as voltage divider or two-supply emitter bias.
  • Different feedback types: collector feedback uses V_C, emitter feedback uses V_E, combination uses both—but all three have similar modest stability levels.
  • Stability condition difficulty: the requirement (one resistor >> another/beta) appears in all configurations, but it's much harder to satisfy in feedback circuits than in voltage divider or two-supply emitter bias.
35

Amplifier Concepts

Chapter 6: Amplifier Concepts

🧭 Overview

🧠 One-sentence thesis

An amplifier multiplies the input signal amplitude by a constant gain factor, and its performance depends on input/output impedances interacting with source and load impedances, along with noise, distortion, frequency range, and output compliance limits.

📌 Key points (3–5)

  • What amplification is: multiplication of signal amplitude by a constant factor without changing frequency or shape.
  • Three types of gain: voltage gain (Aᵥ), current gain (Aᵢ), and power gain (G), each expressing the ratio of output to input.
  • Impedance model: amplifiers are characterized by input impedance (Zᵢₙ), output impedance (Zₒᵤₜ), and a controlled source.
  • Loading effects: voltage dividers form between source/input impedances and output/load impedances, reducing final signal.
  • Common confusion: voltage amplifiers need high Zᵢₙ and low Zₒᵤₜ to minimize loading, while current amplifiers need the opposite (low Zᵢₙ, high Zₒᵤₜ).

🎯 What amplification means

🎯 Core definition and ideal behavior

Amplification is multiplication of the input signal amplitude by a constant.

  • The ideal amplifier should:
    • Multiply amplitude by a constant factor
    • Not change the signal frequency
    • Not alter the signal shape
    • Not add noise
    • Not warp or distort the signal in any way

🔢 Gain as a ratio

  • Gain is defined as the ratio of output signal to input signal.
  • It is a unit-less quantity (a pure number).
  • Example: If input power is 10 milliwatts and output power is 50 milliwatts, power gain = 50 mW / 10 mW = 5.
  • Example: If input voltage is 2 volts and output voltage is 16 volts, voltage gain = 16 V / 2 V = 8.

↕️ Signal inversion

  • Some amplifiers flip the waveform upside down (invert the signal).
  • For sine waves, this is equivalent to a 180° phase shift (−sine output for sine input).
  • Inversion is shown by a negative gain.
  • Example: Aᵥ = −10 means amplification factor of 10 with signal inversion.

📊 Three types of gain

Gain typeSymbolDefinitionExample
Power gainGOutput power / Input power50 mW / 10 mW = 5
Voltage gainAᵥOutput voltage / Input voltage16 V / 2 V = 8
Current gainAᵢOutput current / Input current(not given in excerpt)
  • A stands for "Amplification factor" (used for voltage and current gain).
  • G is historically used for power gain.
  • Choice of gain type depends on the application.

🔌 Amplifier model and impedances

🔌 Functional block model

  • Amplifiers vary in size (single transistor to dozens), so simplified models ease system design.
  • The model uses:
    • A resistor representing input impedance (Zᵢₙ)
    • A controlled source (voltage or current)
    • An internal resistance representing output impedance (Zₒᵤₜ)

🔌 Voltage amplifier model

  • The model specifies Vᵢₙ and Vₒᵤₜ with a controlled voltage source.
  • Output side: controlled voltage source + series Zₒᵤₜ is the Thevenin equivalent viewed from the load.
  • Input side: Zᵢₙ is the equivalent impedance seen by the driving source.
  • The model does not care how the circuit creates the boost, only that it does.

⚡ Impedance design for voltage vs. current amplifiers

Amplifier typeZᵢₙZₒᵤₜReason
Voltage amplifierHighLowMaximize voltage transfer; minimize loading (like a voltmeter input and ideal voltage source output)
Current amplifierLowHighMaximize current transfer
  • Don't confuse: voltage amplifiers need high Zᵢₙ to avoid loading the source, while current amplifiers need low Zᵢₙ to sense current effectively.

⚠️ Loading effects

⚠️ What loading effects are

Loading effects are signal losses caused by interactions between the amplifier's impedances and those of the circuits and loads connected to it.

  • Two voltage dividers form in the system:
    1. Between source impedance (Zₘₑₙ) and input impedance (Zᵢₙ)
    2. Between output impedance (Zₒᵤₜ) and load impedance (Zₗₒₐ𝒹)
  • Each divider reduces the final output voltage.

⚠️ Input loading

  • The voltage reaching the amplifier input is reduced by the first divider:
    • Vᵢₙ₋ₐₘₚ = [Zᵢₙ / (Zᵢₙ + Zₘₑₙ)] × Vₘₑₙ
  • If Zᵢₙ is much larger than Zₘₑₙ, the loss is small.
  • Example: A high Zᵢₙ (like a voltmeter) minimizes this loss.

⚠️ Output loading

  • The voltage at the load is reduced by the second divider:
    • Vₗₒₐ𝒹 = [Zₗₒₐ𝒹 / (Zₗₒₐ𝒹 + Zₒᵤₜ)] × Aᵥ × Vᵢₙ₋ₐₘₚ
  • If Zₒᵤₜ is much smaller than Zₗₒₐ𝒹, the loss is small.
  • Example: A low Zₒᵤₜ (like an ideal voltage source) minimizes this loss.

⚠️ Combined system gain

  • The overall gain from source to load is the product of:
    • Input divider ratio
    • Amplifier gain (Aᵥ)
    • Output divider ratio
  • Both loading effects compound to reduce the final signal.

🛠️ Other amplifier characteristics

🛠️ Output compliance

Output compliance: the maximum output level the amplifier can produce.

  • This is a limit on how large the output signal can be.
  • Beyond this limit, the amplifier cannot increase the output further.

🛠️ Frequency range

  • Amplifiers have a useful frequency range (frequency limits).
  • The excerpt mentions this in general terms but does not provide specifics.

🛠️ Noise and distortion

  • Noise: unwanted signals added to the output.
  • Waveform distortion: changes to the signal shape.
  • The excerpt distinguishes these two concepts but does not elaborate further.

🛠️ Miller's Theorem

  • The excerpt defines Miller's Theorem as a concept but provides no details.
  • (This is listed as a learning objective but not explained in the provided text.)
36

Amplifier Concepts

6.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

Amplifiers multiply input signals by a constant gain factor, and their performance depends on input/output impedances interacting with source and load impedances, along with factors like noise, distortion, frequency range, and output compliance.

📌 Key points (3–5)

  • What amplification is: multiplication of signal amplitude by a constant factor without changing frequency or shape.
  • Three types of gain: voltage gain (Aᵥ), current gain (Aᵢ), and power gain (G), each expressing the ratio of output to input.
  • Impedance model: amplifiers are characterized by input impedance (Zᵢₙ), output impedance (Zₒᵤₜ), and a controlled source.
  • Loading effects: signal losses occur through voltage dividers formed between the amplifier's impedances and the source/load impedances.
  • Common confusion: voltage amplifiers need high Zᵢₙ and low Zₒᵤₜ to maximize voltage transfer, while current amplifiers need the opposite (low Zᵢₙ, high Zₒᵤₜ).

🎯 What amplifiers do

🎯 Core function

Amplification: multiplication of the input signal amplitude by a constant factor.

  • The ideal amplifier only increases amplitude—it does not:
    • Change the signal's frequency
    • Alter its shape
    • Add noise
    • Distort the signal in any way
  • Real amplifiers are characterized by several parameters beyond just gain: output compliance (maximum output level), frequency range, noise, and distortion.

🔢 Gain definitions

Gain: the ratio of output signal to input signal; a unit-less quantity.

Three types of gain:

Gain typeSymbolDefinitionExample from excerpt
Power gainGOutput power / Input power50 mW / 10 mW = 5
Voltage gainAᵥOutput voltage / Input voltage16 V / 2 V = 8
Current gainAᵢOutput current / Input current(not given in excerpt)
  • The "A" in Aᵥ and Aᵢ stands for "Amplification factor."
  • Gain is expressed as a simple number, not in units.

🔄 Signal inversion

  • Some amplifiers flip the waveform upside down (invert it).
  • For sine waves, this is equivalent to a 180° phase shift: a sine input produces a −sine output.
  • Inversion is shown by a negative gain value.
  • Example: Aᵥ = −10 means amplification factor of 10 with signal inversion.
  • Don't confuse: the negative sign indicates inversion, not a reduction in amplitude.

🧱 Amplifier model structure

🧱 Functional block model

  • Amplifiers range from single transistors to dozens of transistors.
  • To simplify system design, use functional models instead of full circuit details.
  • The model uses:
    • A resistor for input impedance (Zᵢₙ)
    • A controlled source (voltage or current)
    • A resistor for output impedance (Zₒᵤₜ)

🔌 Voltage amplifier model components

The excerpt describes a voltage amplifier model with three key elements:

  1. Input impedance (Zᵢₙ): the equivalent impedance seen by the driving source.
  2. Controlled voltage source: represents the amplification mechanism (multiplies input by Aᵥ).
  3. Output impedance (Zₒᵤₜ): the Thevenin equivalent resistance in series with the controlled source, as seen from the load.
  • The model specifies Vᵢₙ and Vₒᵤₜ, with a controlled voltage source inside.
  • The model does not care how the circuit creates the boost, only that it does.

⚡ Design for voltage vs. current transfer

Voltage amplifiers (maximize voltage transfer):

  • High input impedance (like a voltmeter—minimizes loading on the source)
  • Low output impedance (like an ideal voltage source—delivers voltage efficiently to the load)

Current amplifiers (maximize current transfer):

  • Low input impedance
  • High output impedance

Example: if you want to transfer voltage efficiently, you need Zᵢₙ to be high so the source isn't loaded down, and Zₒᵤₜ to be low so the load receives the full voltage.

📉 Loading effects

📉 What loading effects are

Loading effects: signal losses caused by interactions between the amplifier's impedances and the impedances of connected circuits and loads.

  • Two voltage dividers form in the complete system:
    1. Between the source impedance (Zₐₑₙ) and the amplifier's input impedance (Zᵢₙ)
    2. Between the amplifier's output impedance (Zₒᵤₜ) and the load impedance (Zₗₒₐ𝒹)
  • Each divider reduces the signal, lowering the final output voltage.

📐 Input loading

The voltage actually reaching the amplifier input is reduced by the first divider:

Vᵢₙ₋ₐₘₚ = [ Zᵢₙ / (Zᵢₙ + Zₐₑₙ) ] × Vₐₑₙ

  • If Zᵢₙ is much larger than Zₐₑₙ, the fraction approaches 1 and loss is small.
  • If Zᵢₙ is comparable to or smaller than Zₐₑₙ, significant signal is lost before amplification.

📐 Output loading

The voltage delivered to the load is reduced by the second divider:

Vₗₒₐ𝒹 = [ Zₗₒₐ𝒹 / (Zₗₒₐ𝒹 + Zₒᵤₜ) ] × Aᵥ × Vᵢₙ₋ₐₘₚ

  • If Zₗₒₐ𝒹 is much larger than Zₒᵤₜ, the fraction approaches 1 and loss is small.
  • If Zₗₒₐ𝒹 is comparable to or smaller than Zₒᵤₜ, the amplified signal is significantly reduced at the load.

🔗 Combined system gain

The overall gain from source to load is the product of:

  • The input divider ratio
  • The amplifier's voltage gain (Aᵥ)
  • The output divider ratio

This combined gain is always less than the amplifier's intrinsic gain Aᵥ due to the two loading losses.

Don't confuse: the amplifier's internal gain (Aᵥ) is a property of the amplifier itself; the system gain includes losses from impedance mismatches at input and output.

🎛️ Other amplifier characteristics

🎛️ Key parameters beyond gain

The excerpt lists additional characteristics that matter for amplifier performance:

  • Output compliance: the maximum output level the amplifier can produce.
  • Frequency range: the useful frequency limits of the amplifier (mentioned in learning objectives).
  • Noise: unwanted random signals added to the output (mentioned as a concept to distinguish).
  • Waveform distortion: changes to the signal shape (mentioned as distinct from noise).

These are introduced as concepts in the chapter objectives but not detailed in the excerpt provided.

🧩 Amplifier types by sensing and sourcing

  • Amplifiers can be designed to sense voltage or current at the input.
  • They can be modeled as controlled voltage sources or controlled current sources at the output.
  • The choice depends on the application and the type of signal transfer desired.
37

Decibels and Bode Plots

6.1 Introduction

🧭 Overview

🧠 One-sentence thesis

Decibels provide an alternative measurement scheme widely used in audio and communications, and Bode plots extend frequency response analysis to its logical conclusion by graphing circuit behavior across a range of frequencies.

📌 Key points (3–5)

  • What decibels offer: an alternative to ordinary measurement systems, with particular advantages in audio and communications fields.
  • What Bode plots reveal: how a circuit responds to input signals over a range of frequencies—the circuit's frequency response taken to its logical conclusion.
  • Two conversion skills needed: translating between ordinary and decibel-based power gains, and between ordinary and decibel-based voltage gains.
  • Common confusion: decibels are not just for power—they apply to both power and voltage measurements during circuit analysis.
  • System-level analysis: combining effects of multiple lead and lag networks to determine an overall system Bode plot.

📏 The decibel measurement scheme

📏 What decibels are

Decibel: a measurement scheme that serves as an alternative to the ordinary system of measurement used previously.

  • The excerpt does not define the mathematical relationship yet, but emphasizes that decibels are a different way to express the same quantities.
  • They are "in wide use," especially in audio and communications.
  • The chapter will cover conversion methods between the two systems.

🔄 Converting between systems

The excerpt states you will learn to:

  • Convert ordinary power gain ↔ decibel-based power gain.
  • Convert ordinary voltage gain ↔ decibel-based voltage gain.

Don't confuse: decibels can represent both power and voltage; they are not limited to one type of measurement.

🛠️ Using decibels in circuit analysis

  • Decibel-based voltage and power measurements can be used during circuit analysis, not just as final results.
  • The excerpt mentions "advantages over the ordinary system"—these will be detailed later in the chapter.

📊 Bode plots and frequency response

📊 What a Bode plot is

Bode plot: a graph that represents a circuit's frequency response—the way the circuit responds to input signals over a range of frequencies.

  • It is described as taking frequency response analysis "to its logical conclusion."
  • The excerpt notes that frequency response has been investigated "to some extent in prior work," so Bode plots build on earlier concepts.

📈 What you graph

  • The plot shows how the circuit behaves across a range of frequencies, not just at one frequency.
  • Example: an amplifier's gain might change as the input signal frequency sweeps from low to high; the Bode plot captures this entire behavior.

🔀 Lead and lag networks

🔀 Two types of networks

The excerpt mentions:

  • Lead networks
  • Lag networks

The chapter will "detail the differences" between them and show how to graph Bode plots for each.

Don't confuse: these are distinct network types with different frequency behaviors; the excerpt does not yet explain the difference, but signals that distinguishing them is important.

🧩 Combining multiple networks

  • Real systems often contain several lead and lag networks.
  • You will learn to "combine the effects" of these networks to determine a single system Bode plot.
  • This implies that individual network responses can be analyzed separately and then integrated.

🎯 Chapter scope summary

TopicWhat you will learn
DecibelsConvert ordinary ↔ decibel for power and voltage; use decibels in analysis
Bode plotsDefine and graph general Bode plots; understand frequency response visualization
Lead/lag networksDistinguish the two types; graph Bode plots for each
System analysisCombine multiple network effects into one system Bode plot

Note: The excerpt is an introduction and chapter learning objectives section; detailed mechanisms, formulas, and examples will appear in subsequent sections.

38

Amplifier Model

6.2 Amplifier Model

🧭 Overview

🧠 One-sentence thesis

An amplifier multiplies the input signal by a constant gain factor without altering frequency or shape, and its performance depends on input/output impedances that cause loading effects and compliance limits that cause distortion.

📌 Key points (3–5)

  • What amplification is: multiplication of signal amplitude by a constant factor (gain), ideally without changing frequency, shape, or adding noise.
  • How amplifiers are modeled: as controlled sources with input impedance, output impedance, and a gain factor (voltage, current, or power).
  • Loading effects: signal losses caused by voltage dividers formed between the amplifier's impedances and the source/load impedances.
  • Common confusion: ideal gain vs. system gain—loading effects reduce the actual output below the ideal gain × input calculation.
  • Compliance and clipping: all amplifiers have maximum output limits set by power supply; exceeding this causes waveform distortion and adds new frequency components.

🔧 What amplifiers do and how they're characterized

🔧 Core function

Amplification is just multiplication—the ideal amplifier multiplies the amplitude of the input signal by a constant.

  • The amplifier should not change frequency, alter shape, add noise, or distort the signal.
  • It increases signal strength while preserving all other characteristics.

📊 Key specifications

ParameterWhat it describesNotes
GainAmplification factor (output/input ratio)Unit-less; denoted G for power, A_v for voltage, A_i for current
Input impedance (Z_in)Resistance seen looking into the amplifierHigher is better for voltage amplifiers (minimizes loading)
Output impedance (Z_out)Internal resistance of the outputLower is better for voltage amplifiers (acts like ideal source)
ComplianceMaximum output signal levelSet by DC power supply and design
Frequency rangeUseful operating frequencies
Noise and distortionUnwanted signal alterations

➕ Gain calculation

  • Power gain example: 10 milliwatts input → 50 milliwatts output = gain of 5
  • Voltage gain example: 2 volts input → 16 volts output = A_v of 8
  • Gain is a ratio (output/input), so it has no units.

🔄 Signal inversion

  • Some amplifiers flip the waveform upside down (180° phase shift for sine waves).
  • This is shown by a negative gain: A_v = −10 means gain of 10 with inversion.
  • Example: sine input produces −sine output.

🧩 Functional model and impedances

🧩 Simplified circuit model

  • Amplifiers range from one transistor to dozens, but a simplified model helps system design.
  • The model uses:
    • Z_in: resistor representing input impedance
    • Controlled source: produces the amplified signal (voltage or current)
    • Z_out: series resistance representing output impedance

🎯 Design goals by amplifier type

Amplifier typeInput impedanceOutput impedanceWhy
Voltage amplifierHigh Z_inLow Z_outMinimize loading on input (like voltmeter); act like ideal voltage source at output
Current amplifierLow Z_inHigh Z_outMaximize current transfer
  • The model doesn't care how the circuit creates the boost, only that it does.

⚠️ Loading effects

⚠️ What causes loading

Loading effects are signal losses caused by interactions between the amplifier's impedances and those of the circuits and loads connected to it.

  • Two voltage dividers form:
    1. Between source impedance (Z_gen) and input impedance (Z_in)
    2. Between output impedance (Z_out) and load impedance (Z_load)
  • Each divider reduces the signal.

📐 How loading reduces gain

  • Input voltage reduction: The voltage at the amplifier input is less than the source voltage because Z_gen and Z_in form a divider.
    • Input voltage = (Z_in / (Z_in + Z_gen)) × source voltage
  • Output voltage reduction: The load voltage is less than the amplifier's internal output because Z_out and Z_load form a divider.
    • Load voltage = (Z_load / (Z_load + Z_out)) × gain × input voltage
  • System gain (actual gain from source to load) = (Z_in / (Z_in + Z_gen)) × A_v × (Z_load / (Z_load + Z_out))

✅ Minimizing loading losses

  • Want Z_in much greater than Z_gen (input impedance >> source impedance)
  • Want Z_load much greater than Z_out (load impedance >> output impedance)

🧮 Worked scenario

Example from the excerpt: amplifier with A_v = 20, Z_in = 10 kΩ, Z_out = 200 Ω; driven by 30 mV source with 600 Ω impedance; driving 1 kΩ load.

  • Step 1: Input divider reduces 30 mV to 28.3 mV at amplifier input
  • Step 2: Amplifier multiplies by 20 → 566 mV internally
  • Step 3: Output divider reduces to 471.7 mV at load
  • Without loading: would be 30 mV × 20 = 600 mV (ideal)
  • Don't confuse: The excerpt notes that with lab equipment (50 Ω source, 1 MΩ scope load), loading would be minimal and output would approach the ideal 600 mV.

🚫 Compliance and distortion

🚫 What compliance means

Compliance: the maximum output signal (typically maximum output voltage) that an amplifier can produce.

  • Set by the DC power supply and amplifier design.
  • All amplifiers have limits—the idealization that output = input × gain eventually fails.

✂️ Clipping distortion

Clipping: when the output signal is abruptly limited to the compliance level, removing portions of the waveform that would exceed it.

  • It's as if electronic scissors clipped off the top of the waveform.
  • Example: a sine wave with severe clipping can look more like a square wave.
  • This is extreme waveform distortion with important consequences.

🌊 Frequency content changes

  • Key principle: Whenever a signal is altered in the time domain, its frequency content changes.
  • Clipping adds new frequency components (harmonics) to the signal.
  • Levels of existing components may change or be deleted.

🎵 Harmonics and wave shaping

  • Fundamental: the base frequency (original sine wave)
  • Harmonic: another sine wave at an integer multiple of the fundamental frequency (e.g., 3× the fundamental)
  • Adding harmonics changes the wave shape:
    • One harmonic (3× fundamental) creates a "lumpy" waveform
    • More harmonics smooth out variations, approaching a square wave
  • Conclusion from excerpt: A clipped sine wave has new frequency components added; if an amplifier clips, it introduces distortion by adding harmonics.
39

Compliance and Distortion

6.3 Compliance and Distortion

🧭 Overview

🧠 One-sentence thesis

All amplifiers have a maximum output signal limit (compliance) beyond which the output waveform becomes distorted, and this distortion adds new frequency components (harmonics) that alter the signal's character.

📌 Key points (3–5)

  • Compliance: the maximum output signal an amplifier can produce, set by the DC power supply and design; exceeding it causes distortion.
  • Clipping: the most common distortion form where the output waveform is abruptly cut off at the compliance level, adding many new harmonics.
  • Half-wave symmetry: a test to determine whether distortion produces only odd harmonics (symmetric) or includes even harmonics (asymmetric).
  • Common confusion: distortion vs. noise—distortion is correlated with the input signal and adds harmonics; noise is random, broad-band, and uncorrelated with the input.
  • Measurement: Total Harmonic Distortion (THD) quantifies distortion by measuring all added harmonics as a percentage of the total signal.

🚧 Compliance and clipping

🚧 What compliance means

Compliance: the maximum output signal (typically voltage) that an amplifier can produce, determined by the DC power supply and amplifier design.

  • Every amplifier has a limit on output signal size.
  • The idealization "output = input × gain" breaks down when you try to exceed this limit.
  • Example: An amplifier with ±10V compliance cannot produce output voltages beyond that range.

✂️ Clipping mechanism

  • When the desired output exceeds compliance, the waveform is strictly and abruptly limited to the compliance level.
  • Any portion that would extend beyond is removed—like electronic scissors cutting off the top.
  • The excerpt shows an example where severe clipping transforms a sine wave into something resembling a square wave.
  • Don't confuse: clipping with simple amplitude reduction—clipping changes the waveform shape, not just its size.

🎵 Frequency content and harmonics

🎵 Time domain changes create frequency changes

  • Whenever a signal is altered in time, its frequency content changes.
  • New frequency components may be added, existing ones may change level, or some may be deleted.
  • The excerpt emphasizes: extreme clipping adds a large number of new frequency components.

🎼 How harmonics build waveforms

Fundamental: the base frequency, a simple sine wave.

Harmonic: another sine wave at an integer multiple of the fundamental frequency, which may differ in amplitude and phase.

The excerpt demonstrates this with examples:

  • Starting with a fundamental (green sine wave)
  • Adding a third harmonic (three times the fundamental frequency, blue)
  • The sum (red) looks somewhat like a square wave with "lumpy" top and bottom
  • Adding more harmonics (up to seven) smooths out variations, approaching a square wave
  • Conclusion: the clipped sine wave contains new frequency components; clipping music adds audible harmonics that change perception

🎧 Practical implications

  • For music signals: clipping adds harmonics that are likely audible and can change perception subtly or drastically.
  • Example: A high-fidelity audio amplifier might have THD below 0.1%, while an overdriven guitar amplifier might exceed 20%.

🔀 Subtle distortion and symmetry

🔀 Non-clipping distortion

  • Amplifiers can exhibit more subtle distortion due to internal nonlinearity.
  • Example: gain varying slightly as the signal swings from low to high or negative to positive.
  • The excerpt shows a distorted wave that initially appears merely offset but is truly distorted when vertically shifted.

⚖️ Half-wave symmetry test

Half-wave symmetry: when the negative portion of a wave is a mirror image of the positive portion (rotated around the time axis).

How to test (three steps from the excerpt):

  1. Consider the waveform (e.g., sawtooth)
  2. Rotate the negative portion around the time axis
  3. Slide the negative portion over the positive portion and check if they're identical

What it reveals:

  • Waves with half-wave symmetry contain only odd harmonic distortion (odd integer multiples of fundamental)
  • Waves lacking half-wave symmetry have at least one even harmonic
  • The distorted wave in the excerpt lacks half-wave symmetry (asymmetric), so it must contain at least one even harmonic

🔍 Distortion vs. offset

  • Don't confuse: a distorted waveform with a merely offset waveform.
  • The excerpt shows that after vertical shifting, true distortion becomes apparent—the wave is no longer a pure sine.

📏 Measuring distortion

📏 Total Harmonic Distortion (THD)

Measurement procedure:

  1. Apply a very pure, low-distortion sine wave (the fundamental) to the amplifier
  2. At the output, use a selective filter to remove the fundamental
  3. What remains are the added distortion harmonics
  4. Treat these harmonics as a lumped value and present as a percentage of the total signal

Interpretation:

  • Double-digit THD levels are relatively easy to discern on an oscilloscope
  • THD much below 1% is very difficult or impossible to discern by eye
  • What matters is what we can hear, not how the waveform looks

Limitations of THD:

  • All distortion products are lumped together
  • Says nothing about which harmonics are strong or their distribution
  • Doesn't address what happens when multiple frequencies interact

📊 Intermodulation Distortion (IMD)

  • Method: apply two sine waves at different frequencies simultaneously
  • Attempts to quantify how multiple frequencies interact
  • Also expressed as a percentage
  • Complements THD by revealing interaction effects

🔊 Distortion vs. noise comparison

CharacteristicDistortionNoise
CorrelationCorrelated with input signalNot correlated with input signal level
NatureAdds harmonics (integer multiples of input frequencies)Broad-band, random signal
PitchHas discernible pitch relationshipsNo discernible pitch
RemovalTheoretically predictableTruly random, cannot be accurately predicted, no easy removal
SourcesClipping, nonlinearity, gain variationProcess issues in semiconductors, thermal effects in resistors

Don't confuse: distortion with noise—they are fundamentally different phenomena requiring different analysis and mitigation strategies.

40

Frequency Response and Noise

6.4 Frequency Response and Noise

🧭 Overview

🧠 One-sentence thesis

All amplifiers are limited by the range of frequencies they can amplify at full gain and by internal noise that appears at the output, both of which constrain practical performance.

📌 Key points (3–5)

  • Frequency response limits: amplifiers have a mid-band where nominal gain is accurate, bounded by corner frequencies f₁ (lower) and f₂ (upper); outside this range, gain drops off.
  • DC vs AC amplifiers: all amplifiers have an upper frequency limit f₂, but only some have a lower limit f₁—DC amplifiers can amplify down to zero frequency.
  • What causes frequency limits: lower limits usually come from coupling capacitors or transformers that block DC; upper limits are unavoidable due to stray capacitances and inductances.
  • Noise characteristics: noise is an undesired, broad-band, random signal not correlated with input level; it cannot be easily removed once added.
  • Common confusion: frequency response is not a single gain number—it varies with frequency, and the "amplifier" may even reduce signal level at extreme frequencies.

📉 Frequency response fundamentals

📉 Mid-band and corner frequencies

Mid-band: the region where the nominal gain is accurate.

Corner (break) frequencies: the frequencies f₁ (lower limit) and f₂ (upper limit) that define the mid-band range; at these frequencies, output power has dropped to half the power exhibited by a mid-band frequency at the same input level.

  • The amplifier's stated gain or amplification factor is true only for frequencies within the mid-band.
  • Outside the mid-band, gain begins to drop off as frequency moves farther away.
  • Eventually, gain falls to practically zero and virtually no trace of the input signal appears at the output.
  • Example: a signal at mid-band frequency receives full amplification; a signal far below f₁ or far above f₂ may be almost completely lost.

🎚️ Gain behavior at extreme frequencies

  • At extreme frequencies (very low or very high), the gain may be much less than the nominal value.
  • If you go far enough, the gain may even be fractional, meaning the "amplifier" is actually reducing the signal level.
  • This is not a malfunction—it is an inherent limitation of all amplifiers.

🔄 DC amplifiers vs frequency-limited amplifiers

🔄 Lower frequency limit (f₁)

  • Not all amplifiers have a lower frequency limit f₁.
  • Amplifiers without a lower limit can amplify signals all the way down to DC (zero frequency).

Direct coupled (DC) amplifiers: amplifiers with no lower frequency limit; they can amplify frequencies down to DC.

  • The lower frequency limit is usually caused by in-line coupling capacitors and sometimes transformers.
  • These components are added to purposely block DC.
  • If an amplifier is designed without these components, it will have no limit on how low a frequency it can amplify.

⚡ Upper frequency limit (f₂)

  • Without exception, all amplifiers have an upper limit frequency f₂.
  • Even if no tailoring is desired, the amplifier would still have an upper limit due to small and unavoidable capacitances and inductances in the circuit (e.g., stray wiring capacitance).
  • These reactances cause signal level reduction that worsens as frequency increases.
  • They also cause varying phase shifts between the input and output signals.
  • Don't confuse: the upper limit is not optional—it is unavoidable, unlike the lower limit which can be eliminated by design.

📊 Application-specific frequency ranges

Applicationf₁ (lower limit)f₂ (upper limit)Notes
High fidelity audioBelow 20 HzAbove 20 kHzCovers the range of frequencies heard by a typical healthy young human (20 Hz – 20 kHz)
Telephone systems300 Hz4 kHzNot hi-fi, but sufficient for voice communication
Radio frequency(varies)Orders of magnitude higherOperating at much higher frequencies than audio

🔊 Noise in amplifiers

🔊 What noise is

Noise: an undesired signal that appears at the output of an amplifier.

  • Unlike distortion, noise is usually not correlated with the input signal level.
  • Noise is broad-band, meaning it contains a very wide range of frequencies.
  • Because it is broad-band, noise does not have a discernible pitch.
  • Noise is best thought of as a truly random signal.
  • Examples in nature: the sound of leaves rustling in the wind or the sound of a waterfall.

⚠️ Why noise is a problem

  • Because noise is truly random, it cannot be accurately predicted.
  • Therefore, there is no easy way to remove it once it has been added to a desired signal.
  • Noise is unavoidable in absolute terms.
  • What matters is whether the noise level is low enough for a given application—i.e., significantly lower than the signal level.

🔬 Sources and factors affecting noise

  • Many potential sources: process issues in semiconductors, thermal effects in resistive elements, and others.
  • In general, noise gets worse as:
    • Temperature increases
    • Resistance increases
    • Frequency range increases

📐 Signal-to-noise ratio (S/N)

Signal-to-noise ratio (S/N): a ratio between the nominal output signal level and the output noise level.

  • This ratio quantifies whether noise is low enough for a given application.
  • All other factors being equal, the higher the S/N, the better.
  • Example: if the signal level is much larger than the noise level, the S/N is high and noise is no longer a problem.
41

Miller's Theorem

6.5 Miller's Theorem.

🧭 Overview

🧠 One-sentence thesis

Miller's Theorem simplifies the analysis of inverting voltage amplifiers by replacing a bridging impedance between input and output with equivalent smaller parallel impedances at the input and output.

📌 Key points (3–5)

  • What Miller's Theorem does: converts a single impedance bridged between input and output into two separate equivalent impedances in parallel with the input and output.
  • Why it's useful: simplifies circuit analysis by turning a bridging element into standard input/output network components.
  • Key effect: the equivalent impedances are always smaller than the original bridging impedance; the input reduction is especially large at higher gains.
  • Common confusion: the theorem applies only to inverting voltage amplifiers (negative gain), not all amplifier types.
  • Typical cases: resistors and capacitors are the two most common bridging impedances; capacitors show a multiplicative effect at the input.

🔌 The bridging impedance problem

🔌 What a bridging impedance is

  • Some inverting voltage amplifier designs place an impedance Z between the input and output terminals.
  • This bridging configuration is used for different reasons, a prime example being shaping the frequency response of the amplifier.
  • The bridging element complicates analysis because it connects two different nodes with different voltages.

🎯 Goal of Miller's Theorem

Miller's Theorem: a technique to determine equivalent impedances that lie in parallel with the input and output of an inverting voltage amplifier.

  • The equivalents simply become part of the input and output networks around the amplifier.
  • This transforms a two-port bridging element into two one-port elements, making the circuit easier to analyze.

🧮 The Miller equivalent formulas

🧮 Input equivalent impedance

The input Miller equivalent impedance is defined as the impedance in parallel with the input that would draw the same amount of current as the original bridging impedance.

Derivation logic:

  • Current through the Miller impedance = (voltage across it) / Z
  • Voltage across it = V_in − V_out
  • Since the amplifier is inverting, V_out = −A_v × V_in (gain is negative)
  • So voltage across Z = V_in − (−A_v × V_in) = V_in × (|A_v| + 1)
  • Current = V_in × (|A_v| + 1) / Z
  • Dividing this current into V_in yields the equivalent impedance:

Z_in-miller = Z / (|A_v| + 1)

  • The input equivalent is the original impedance divided by (magnitude of gain plus one).
  • At higher gains, the reduction effect is very large.

🧮 Output equivalent impedance

A similar derivation yields:

Z_out-miller = Z × |A_v| / (|A_v| + 1)

  • The output equivalent is also smaller than the original bridging impedance, but not reduced as dramatically as the input equivalent.

📏 General rule

  • The Miller equivalent presents equivalent impedances that are less than the original bridging impedance.
  • The reduction is a function of the amplifier's voltage gain.
  • Example: if an amplifier has high gain (say |A_v| = 100) and a bridging impedance of 100 kΩ, the input equivalent would be approximately 100 kΩ / 101 ≈ 1 kΩ, a dramatic reduction.

🔧 Two typical cases

🔧 Resistor case

For a pure resistance R, perform a direct substitution for Z in the formulas:

  • Input equivalent resistance: R_in-miller = R / (|A_v| + 1)
  • Output equivalent resistance: R_out-miller = R × |A_v| / (|A_v| + 1)
  • The bridging resistor is replaced by two smaller resistors in parallel with the input and output.

🔧 Capacitor case

For a capacitor, substitute the capacitive reactance X_c for Z, recalling that C = 1 / (2π f X_c).

Key difference:

  • For the capacitor, there is a multiplicative effect at the input.
  • The effect of the original bridging capacitor on a high-gain amplifier is equivalent to a much larger input shunt capacitor.
  • Input equivalent capacitance: C_in-miller = C × (|A_v| + 1)
  • Output equivalent capacitance: C_out-miller = C × (|A_v| + 1) / |A_v|

Why this matters:

  • A small bridging capacitor can appear as a large capacitor at the input when gain is high.
  • This is important for frequency response analysis, as larger capacitance affects the amplifier's behavior at different frequencies.
  • Example: a 200 pF bridging capacitor with |A_v| = 30 becomes an input equivalent of 200 pF × 31 = 6200 pF = 6.2 nF at the input.

⚠️ Don't confuse

  • Resistors get smaller in both input and output equivalents.
  • Capacitors get larger at the input (multiplicative effect) but the output equivalent is close to the original value.
  • The theorem only applies to inverting voltage amplifiers; the gain must be negative for the formulas to work correctly.
42

BJT Small Signal Amplifiers

Chapter 7: BJT Small Signal Amplifiers

🧭 Overview

🧠 One-sentence thesis

Small signal BJT amplifier analysis focuses on voltage gain and input/output impedances using an AC model based on the dynamic resistance of the base-emitter junction, which can be derived from the device's operating point on its I-V curve.

📌 Key points (3–5)

  • Small signal definition: output signals well below clipping with power dissipation no more than a few hundred milliwatts for load or transistor.
  • Two analysis approaches: hybrid parameters (h-parameters like hfe, hie, hoe, hre) versus r' parameters; the r' approach is simpler and sufficient.
  • AC model basis: the collector-base uses a current source (iC = β iB), while the base-emitter junction requires modeling its AC dynamic resistance.
  • Dynamic resistance concept: found by taking the reciprocal of the slope of the tangent line at the quiescent bias point on the base-emitter I-V curve.
  • Common confusion: the dynamic resistance changes slightly as the signal swings, producing an average value that causes asymmetrical distortion.

🎯 Analysis techniques for BJT amplifiers

🔧 Hybrid parameters approach

Hybrid parameters: four parameters used to analyze BJT circuits—hfe (forward current gain, also called β), hie (input impedance), hoe (output admittance), and hre (reverse voltage gain).

  • The second subscript letter indicates the configuration (e.g., "e" in hfe means common emitter: input at base, output at collector, emitter at ground).
  • This method uses four distinct parameters to characterize the transistor's behavior.

🔧 r' parameters approach

  • Pronounced "r prime."
  • Sufficient for all analyses covered in this chapter.
  • Produces straightforward equations for circuit gain and input impedance using basic principles: Ohm's law, KVL (Kirchhoff's Voltage Law), and KCL (Kirchhoff's Current Law).
  • The excerpt focuses on this simpler system.

🔌 AC model construction

🔌 Collector-base region modeling

  • Represented with a current-controlled current source.
  • Uses AC instead of DC: iC = β iB.
  • This part of the model is straightforward, similar to the DC model.

🔌 Base-emitter junction modeling

  • The DC model used a simple 0.7 volt junction.
  • For AC analysis, must consider the AC resistance of the diode (dynamic resistance).
  • This is the trickier part of the AC model construction.

📈 Dynamic resistance derivation

📈 Operating point concept

  • The AC signal rides on top of the DC bias current.
  • The signal causes the operating point to trace back and forth along the base-emitter I-V curve.
  • For small signals, this sweep is very small (only a few percent of quiescent current) and can be approximated as a straight line segment.

📈 Slope and resistance relationship

Dynamic resistance: the reciprocal of the slope of the line tangent to the operating point (quiescent bias current IC) on the base-emitter I-V curve.

  • The slope of the line segment represents conductance.
  • Resistance is the reciprocal of conductance.
  • Therefore: dynamic resistance = 1 / (slope at quiescent point).
  • This slope is technically called the device's transconductance, denoted as gm.

⚠️ Resistance variation and distortion

  • The slope changes slightly as the signal swings:
    • Positive swing (above quiescent point): steeper slope → slightly lower resistance.
    • Negative swing (below quiescent point): shallower slope → slightly higher resistance.
  • Computing dynamic resistance assumes a straight line, giving an average value.
  • This variance in resistance causes asymmetrical distortion in the amplifier output.
  • Don't confuse: this is not a fixed resistance but an average that varies with signal swing.

🧮 Mathematical foundation

🧮 Shockley equation for BJT

The derivation starts with the Shockley equation (modified for BJT terminals):

IC = IS (e^(VBE·q/nkT) − 1)

Where:

  • IC = junction (collector) current
  • IS = reverse saturation current
  • VBE = voltage across base-emitter junction
  • q = electron charge (1.6×10^-19 coulombs)
  • n = quality factor (typically 1 to 2)
  • k = Boltzmann constant (1.38×10^-23 joules/kelvin)
  • T = temperature in kelvin

🧮 Simplifications at room temperature

  • At 300 kelvin (approximately 80°F), q/kT ≈ 38.6.
  • For any reasonable VBE value, the "−1" term is small enough to ignore.
  • The quality factor n is taken as 1 for this analysis.
  • These simplifications make the equation more practical for circuit analysis.
43

BJT Small Signal Amplifiers: Chapter Objectives and AC Model

7.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

This chapter teaches how to analyze BJT amplifiers using small signal AC models and r' parameters to determine voltage gain, input impedance, and output impedance for circuits operating well below clipping limits.

📌 Key points (3–5)

  • Small signal definition: output signals well below clipping with power dissipation no more than a few hundred milliwatts for load or transistor.
  • Two analysis approaches exist: hybrid parameters (h-parameters like hfe, hie, hoe, hre) versus r' parameters; this chapter focuses on r' parameters because they produce straightforward equations using Ohm's law, KVL, and KCL.
  • AC model differs from DC model: the collector-base region uses an AC current source (iC = β iB), but the base-emitter junction requires modeling the AC resistance of the diode, not just 0.7V.
  • Common confusion—dynamic vs static resistance: the AC signal rides on the DC bias current, so the dynamic resistance is the reciprocal of the slope at the operating point, not a fixed value.
  • Chapter scope: covers voltage amplifiers, voltage followers, localized feedback (swamping), multistage amplifiers, coupling methods, and Darlington pairs.

🎯 Learning objectives

🎯 What you will be able to do

After completing this chapter, you should be able to:

  • Determine the voltage gain, input impedance, and output impedance of simple BJT amplifiers.
  • Detail the functional differences between voltage amplifiers and voltage followers.
  • Explain the advantages and disadvantages of using localized feedback (swamping).
  • Determine the combined characteristics of multistage BJT amplifiers.
  • Detail the advantages and disadvantages of using direct coupling versus capacitor coupling in multistage amplifiers.
  • Explain the operation of the Darlington pair.

🔍 Small signal analysis definition

🔍 What qualifies as "small signal"

Small signal analysis: analysis of output signals that are well below the clipping limit and with power dissipation of no more than a few hundred milliwatts for either the load or transistor.

  • There is no specific universal definition separating small signal from large signal.
  • This chapter adopts the above working definition for practical purposes.
  • What is excluded: compliance, maximum load power, device dissipation—these are large-signal concerns.
  • Example: an amplifier operating at 50 mW load power with output voltage at 30% of clipping voltage qualifies as small signal.

🧮 Two analysis techniques

🧮 Hybrid parameters (h-parameters)

ParameterSymbolMeaning
Forward current gainhfeSimply called β
Input impedancehieImpedance looking into the base
Output admittancehoeAdmittance at the collector
Reverse voltage gainhreFeedback voltage ratio
  • The second subscript letter (e.g., "e" in hfe) indicates the configuration: common emitter (input at base, output at collector, emitter at common ground).

🧮 r' parameters (r-prime)

  • Pronounced "r prime."
  • Sufficient for all analyses in this chapter.
  • Produces straightforward equations for circuit gain, input impedance, etc., using only Ohm's law, KVL, and KCL.
  • Why this chapter uses r' parameters: simpler and more intuitive than hybrid parameters for the target applications.

🔧 Simplified AC model of the BJT

🔧 Collector-base region

  • Represented with a current-controlled current source: iC = β iB (AC version).
  • This is analogous to the DC model but now handles AC signals.

🔧 Base-emitter junction challenge

  • The DC model used a simple 0.7 volt junction.
  • For AC analysis, we must consider the AC resistance of the diode (dynamic resistance).
  • The AC signal rides on top of the DC bias current, not in isolation.

📐 Dynamic resistance derivation

📐 Concept: AC signal riding on DC bias

  • The AC signal causes the operating point to trace back and forth along the base-emitter I-V curve.
  • For small signals, this sweep is very small (perhaps only a few percent of the quiescent current) and can be approximated as a straight line segment.
  • The slope of this line segment represents conductance; the reciprocal of the slope represents resistance.

📐 How to find dynamic resistance

Dynamic resistance of the base-emitter junction: the reciprocal of the slope of the line tangent to the operating point (tangent to the quiescent bias current IC).

  • As the signal swings positive (above the quiescent point), the slope steepens slightly → dynamic resistance decreases slightly.
  • As the signal swings negative (below the quiescent point), the slope becomes more shallow → dynamic resistance increases slightly.
  • We compute an average value by assuming a straight line segment.
  • Don't confuse: this is not a fixed resistance; it varies slightly with signal swing, causing asymmetrical distortion (similar to the type shown in Chapter 6, Figure 6.6).

📐 Shockley equation foundation

The derivation begins with the Shockley equation (modified for BJT terminal names):

IC = IS (e^(VBE·q/(n·k·T)) − 1)

Where:

  • IC is the junction (collector) current
  • IS is the reverse saturation current
  • VBE is the voltage across the base-emitter junction
  • q is the charge on an electron, 1.6×10⁻¹⁹ coulombs
  • n is the quality factor (typically between 1 and 2)
  • k is the Boltzmann constant, 1.38×10⁻²³ joules/kelvin
  • T is the temperature in kelvin

Simplifications at 300 kelvin (about 80°F):

  • q/kT is approximately 38.6.
  • For any reasonable VBE, the "−1" term is small enough to ignore.
  • Take n as 1.

📐 Transconductance note

  • The slope value (conductance) is technically called the device's transconductance, denoted as gm.
  • This concept will reappear in later chapters.

⚠️ Distortion implication

⚠️ Source of asymmetrical distortion

  • The variance in dynamic resistance as the signal swings positive and negative produces asymmetrical distortion.
  • This is the same type of distortion illustrated in Chapter 6, Figure 6.6.
  • More details on this effect will be covered later in the chapter.
44

Introduction to Amplifier Frequency Limits

7.1 Introduction

🧭 Overview

🧠 One-sentence thesis

All amplifiers operate effectively only within a specific frequency range, bounded by lower and upper frequency limits that can be analyzed and designed using specialized AC models.

📌 Key points (3–5)

  • Universal frequency limits: Every amplifier has an upper frequency limit (f₂), but the lower limit (f₁) can be designed to reach DC (0 Hz) in direct-coupled amplifiers.
  • What the limits mean: f₁ and f₂ mark where the midband response drops by 3 dB (half power); also called "3 dB down frequencies" or "corner frequencies."
  • Analysis requires multiple models: A complete frequency analysis needs three separate equivalent circuits—one for midband, one for low frequencies, and one for high frequencies.
  • Common confusion: Assumptions valid at midband (e.g., coupling capacitors acting as short circuits) may not hold at frequency extremes.
  • Design capability: Engineers can both calculate existing frequency limits and modify amplifier designs to meet specific frequency requirements for BJT and FET circuits.

🎯 What frequency limits are

📏 Defining the operating range

Frequency limits: The range of frequencies over which an amplifier works effectively, bounded by lower limit f₁ and upper limit f₂.

  • These limits are not arbitrary cutoffs—they represent specific performance thresholds.
  • The concept was introduced earlier (Chapter 6) and is now expanded for detailed analysis.
  • Both BJT and FET amplifiers exhibit these limits, regardless of specific design.

🔻 The 3 dB down point

Corner frequencies (f₁ and f₂): The frequencies at which the midband response has fallen by three decibels, equivalent to half power.

  • Also called "3 dB down frequencies."
  • This is a standard engineering threshold for defining usable bandwidth.
  • Below f₁ or above f₂, the amplifier's performance degrades significantly.

🔄 Two types of frequency behavior

⬇️ Lower frequency limit (f₁)

  • Represents the lowest frequency at which the amplifier maintains adequate performance.
  • Special case: Some amplifiers can be designed with no lower frequency limit.
  • When f₁ = 0 Hz (DC), the amplifier is called "DC coupled" or "direct coupled."
  • Example: A direct-coupled amplifier can amplify signals all the way down to constant (non-changing) voltages.

⬆️ Upper frequency limit (f₂)

  • Represents the highest frequency at which the amplifier maintains adequate performance.
  • Universal constraint: Without exception, all amplifiers exhibit an upper frequency limit.
  • No amplifier can operate effectively at arbitrarily high frequencies.

🔬 Analysis approach

🧪 Multiple AC models needed

The excerpt emphasizes that analyzing frequency limits requires abandoning the single-model approach:

  • Midband model: The familiar model used for typical operating frequencies.
  • Low-frequency model: Specialized model where previous assumptions may break down.
  • High-frequency model: Separate model accounting for high-frequency effects.

⚠️ Why assumptions change

  • Midband assumption example: Coupling capacitors treated as short circuits.
  • At frequency extremes: This assumption "may no longer be valid."
  • Don't confuse: A component's behavior at midband does not predict its behavior at all frequencies.
  • A complete frequency analysis requires examining all three equivalent circuits separately.

🛠️ Design and analysis goals

📊 What engineers can do

CapabilityDescription
AnalyzeCompute the frequency limits (f₁ and f₂) of existing BJT and FET amplifiers
DesignCreate amplifiers to meet specific frequency criteria
ModifyAdjust existing designs to change frequency limits

🧩 Circuit components matter

  • The excerpt mentions that certain circuit components impact low-frequency performance.
  • Other components impact high-frequency performance.
  • Understanding which components affect which limits enables targeted design modifications.
45

7.2 Simplified AC Model of the BJT

7.2 Simplified AC Model of the BJT

🧭 Overview

🧠 One-sentence thesis

The simplified AC model of the BJT uses a dynamic resistance (r'e) set by the DC bias current to predict AC behavior, enabling three amplifier configurations—common emitter, common collector, and common base—each with distinct gain and phase characteristics.

📌 Key points (3–5)

  • Dynamic resistance concept: The AC model treats the base-emitter junction as a small-signal resistance (r'e) determined by the DC collector current, not the AC signal itself.
  • DC sets AC behavior: The DC bias current establishes r'e, so stable DC biasing is essential for stable AC performance.
  • Three amplifier topologies: Common emitter (voltage and current gain, inverts phase), common collector (current gain only, maintains phase), and common base (voltage gain only, maintains phase).
  • Common confusion: The AC signal causes small variations in r'e as it swings along the curved I-V characteristic, but the model assumes a straight-line average, which introduces asymmetrical distortion.
  • Temperature sensitivity: r'e decreases with increasing temperature because the reverse saturation current (I_S) varies with temperature, affecting thermal stability in power amplifiers.

🔬 Dynamic resistance fundamentals

🔬 What dynamic resistance represents

Dynamic (AC) resistance (r'e): the small-signal resistance of the base-emitter junction, calculated as the reciprocal of the slope of the I-V curve at the operating point.

  • The base-emitter junction has a curved exponential I-V relationship (Shockley equation).
  • At any operating point (quiescent bias current I_C), the curve has a local slope.
  • The AC model approximates this curved segment as a straight line with slope dI_C/dV_BE.
  • The dynamic resistance is the reciprocal: dV_BE/dI_C.
  • This is an average value because the actual slope changes slightly as the signal swings above and below the operating point.

📐 Deriving r'e from the Shockley equation

The derivation starts with the Shockley equation for a BJT:

  • I_C = I_S × (e raised to the power of (V_BE × q / (n × k × T)) minus 1)
  • Where I_C is collector current, I_S is reverse saturation current, V_BE is base-emitter voltage, q is electron charge (1.6E-19 coulombs), n is quality factor (1 to 2), k is Boltzmann constant (1.38E-23 joules/kelvin), T is temperature in kelvin.

At 300 kelvin (about 80°F):

  • The factor q/kT is approximately 38.6.
  • For reasonable V_BE values, the "minus 1" term is negligible.
  • Taking n = 1, the equation simplifies to: I_C = I_S × e raised to the power of (38.6 × V_BE).

Taking the first derivative with respect to V_BE:

  • dI_C/dV_BE = 38.6 × I_S × e raised to the power of (38.6 × V_BE).
  • Substituting the simplified Shockley equation back in: dI_C/dV_BE = 38.6 × I_C.

The dynamic resistance is the reciprocal:

  • dV_BE/dI_C = 25.9 millivolts / I_C.
  • Including bulk resistance, a good approximation is: r'e = 26 millivolts / I_C.

⚠️ Nonlinearity and distortion

  • The actual base-emitter curve is not a straight line; it is exponential.
  • As the AC signal swings positive (above the quiescent point), the slope steepens slightly, reducing dynamic resistance.
  • As the signal swings negative (below the quiescent point), the slope becomes more shallow, increasing resistance.
  • The model assumes a straight-line average, so these variations appear as asymmetrical distortion in the amplifier output.
  • Don't confuse: r'e is not constant during the signal swing; the model uses an average that ignores small changes.

🌡️ Temperature and bias stability

🌡️ Temperature dependence of r'e

  • The reverse saturation current I_S in the Shockley equation varies with temperature.
  • As temperature increases, I_S increases, which affects the derivative and thus r'e.
  • r'e decreases with increasing temperature.
  • This has important implications for thermal stability in higher-power amplifiers (discussed in later work).

🔒 DC bias sets AC stability

  • The DC collector current I_C directly determines r'e through the formula r'e = 26 mV / I_C.
  • The AC model's behavior depends on the DC operating point.
  • Stable DC bias is essential for stable AC performance.
  • Example: If the DC bias drifts due to temperature or component variation, r'e changes, altering the AC gain and distortion characteristics.
  • This is why stable bias circuits (from Chapter 5) are emphasized.

🧩 The simplified AC model

🧩 Model structure

The simplified AC model (Figure 7.2 in the excerpt) consists of:

  • r'e: the dynamic resistance of the base-emitter junction, set by DC bias current I_C.
  • Beta (β): the current gain, relating AC base current i_B to AC collector current i_C (i_C = β × i_B).
  • The AC collector current i_C is determined by the AC input current i_B, which is a function of the applied input signal.
  • In contrast, r'e is set by the DC bias current I_C.

🔍 Model limitations

  • This is a simplified model that does not include:
    • Junction capacitance effects.
    • Lead inductance.
    • Other high-frequency parasitics.
  • It is appropriate as a low to mid-band frequency model.
  • The AC input can produce small variations in r'e, which manifest as waveform distortion.

🔧 Three amplifier configurations

🔧 Common emitter

  • Input: applied to the base.
  • Output: taken at the collector.
  • Common terminal: emitter (at ground).
  • Characteristics:
    • Exhibits both voltage gain and current gain.
    • Inverts the phase of the signal (180-degree phase shift).
  • Example: A general-purpose voltage amplifier.

🔧 Common collector (emitter follower)

  • Input: applied to the base.
  • Output: taken at the emitter.
  • Common terminal: collector (at ground).
  • Characteristics:
    • Voltage gain is about unity (approximately 1).
    • Exhibits current gain.
    • Maintains the phase of the input signal (no inversion).
    • Also called an emitter follower or voltage follower.
  • Example: A buffer stage to provide current gain without voltage amplification.

🔧 Common base

  • Input: applied to the emitter.
  • Output: taken at the collector.
  • Common terminal: base (at ground).
  • Characteristics:
    • Exhibits voltage gain.
    • Current gain is unity at best (no current amplification).
    • Maintains the phase of the input signal (no inversion).
  • Example: High-frequency applications where input impedance needs to be low.

📊 Configuration comparison

ConfigurationInputOutputCommon terminalVoltage gainCurrent gainPhase
Common emitterBaseCollectorEmitterYesYesInverted
Common collectorBaseEmitterCollector~1 (unity)YesMaintained
Common baseEmitterCollectorBaseYes~1 (unity)Maintained

🔄 Flexibility in biasing

  • Each of the three topologies can be implemented using various DC bias techniques.
  • Examples: two-supply emitter bias, voltage divider bias.
  • Each can use either NPN or PNP transistors.

🎛️ Common emitter amplifier details

🎛️ Basic circuit structure

The common emitter amplifier (Figure 7.3) is based on a two-supply emitter bias circuit with these additions:

  • Input signal voltage V_in: the AC signal to be amplified.
  • Load R_L: the output load.
  • Coupling capacitors C_in and C_out: isolate the input and load from the DC bias.
    • For DC, these capacitors act as opens, creating isolation.
    • For AC, their reactances are much smaller than surrounding resistors at the signal frequency, so they appear as shorts and pass the AC signal.

🔋 Emitter resistor configuration

  • The single emitter resistor of the bias network is replaced by two resistors: R_E and R_SW, plus a bypass capacitor C_E.
  • For DC: C_E is open, so the effective emitter bias resistance is R_E + R_SW.
  • For AC: C_E behaves ideally as a short, so the AC emitter resistance falls to just R_SW.
  • R_SW is called a swamping resistor or emitter degeneration resistor.
  • Its primary purpose is to help control the voltage gain of the amplifier.

🔌 AC equivalent circuit

Using the AC transistor model and the Superposition Theorem, the AC equivalent circuit (Figure 7.4) is derived:

  1. All capacitors are shorted (they act as shorts for AC).
  2. DC sources are replaced with their ideal internal resistance (a short), placing those points at AC ground.
  3. The transistor is replaced with the AC model.
  4. Resistances are combined and renamed using lowercase r to distinguish from DC resistances (uppercase R).

Notation:

  • r_E: AC resistance from emitter to AC ground (corresponds to R_SW in the original schematic).
  • r_C: total resistance from collector to AC ground (corresponds to R_C in parallel with R_L; if unloaded, r_C = R_C).
  • r_B: corresponds to R_B in a simple bias; in voltage divider bias, it equals R_1 in parallel with R_2.

📈 Voltage gain definition

  • Voltage gain A_v is defined as the ratio of output voltage v_out to input voltage v_in: A_v = v_out / v_in.
  • Using Ohm's law and the AC equivalent circuit, the gain can be calculated (details continue in section 7.3, not fully included in this excerpt).
46

Common Emitter Amplifier

7.3 Common Emitter Amplifier

🧭 Overview

🧠 One-sentence thesis

The common emitter amplifier inverts the signal and provides both voltage and current gain, with the swamping resistor creating a trade-off between maximum gain and lower distortion plus higher input impedance.

📌 Key points (3–5)

  • Three transistor amplifier configurations: common emitter (voltage and current gain, inverts phase), common collector (unity voltage gain, current gain, maintains phase), and common base (voltage gain, unity current gain, maintains phase).
  • How coupling and bypass capacitors work: coupling capacitors isolate DC bias while passing AC signals; bypass capacitors short AC to ground while appearing open to DC.
  • Swamping resistor trade-off: larger swamping resistor reduces gain but decreases distortion and increases input impedance—a "quality versus quantity" choice.
  • Common confusion: the emitter resistor splits into two parts—one bypassed (for DC bias) and one not bypassed (the swamping resistor RSW that controls AC gain).
  • Voltage gain depends on resistance ratio: gain is approximately minus the ratio of collector resistance to total emitter resistance (r'e + rE), with the negative sign indicating phase inversion.

🔧 Three amplifier topologies

🔧 Common Emitter configuration

Common Emitter: input applied to the base, output taken at the collector, emitter terminal at common or ground point.

  • Exhibits both voltage gain and current gain.
  • Inverts the phase of the signal (180° shift).
  • Finds wide use as a general-purpose voltage amplifier.
  • Can be implemented with various DC bias techniques (two-supply emitter bias, voltage divider bias) and either NPN or PNP transistors.

🔧 Common Collector configuration

Common Collector: input applied to the base, output taken at the emitter, collector terminal at common or ground point.

  • Voltage gain is about unity (approximately 1).
  • Exhibits current gain.
  • Maintains the phase of the input signal (no inversion).
  • Also called an emitter follower or voltage follower.

🔧 Common Base configuration

Common Base: input applied to the emitter, output taken at the collector, base terminal at common or ground point.

  • Exhibits voltage gain.
  • Current gain is unity at best (approximately 1).
  • Maintains the phase of the input signal (no inversion).

🧩 Circuit components and their roles

🧩 Coupling capacitors (Cin and Cout)

Coupling capacitors: isolate the input signal source and load from the DC bias circuit.

  • Act as opens to DC, preventing DC bias alteration.
  • Chosen so their reactances are much smaller than surrounding resistors at the signal frequency.
  • Appear as shorts to AC signals, allowing the signal to pass through the amplifier.
  • Example: Cin isolates the input signal voltage Vin from affecting DC bias; Cout isolates the load RL.

🧩 Bypass capacitor (CE) and emitter resistor split

Bypass capacitor: placed in parallel with part of the emitter resistor to create different AC and DC resistances.

  • For DC: capacitor is open, so effective emitter bias resistance is RE + RSW.
  • For AC: capacitor behaves ideally as a short, so AC emitter resistance falls to just RSW.
  • The single bias resistor is replaced by a pair: RE and RSW.
  • Don't confuse: the total DC emitter resistance includes both resistors, but the AC emitter resistance (rE) is only the swamping resistor RSW.

🧩 Swamping (emitter degeneration) resistor RSW

Swamping resistor (RSW): the un-bypassed portion of the emitter resistor, used primarily to control voltage gain.

  • Also called an emitter degeneration resistor because it degrades (reduces) voltage gain.
  • Larger RSW means lower gain but better performance in other ways.
  • When RSW = 0 (emitter completely bypassed), gain is maximum but distortion is worst.
  • The resistor "swamps out" the variation in r'e, reducing distortion—the larger RSW is relative to r'e, the greater the distortion reduction.

📐 AC equivalent circuit analysis

📐 Building the AC equivalent

The AC equivalent circuit is derived using the Superposition Theorem and the AC transistor model:

  1. Short all capacitors (they act as shorts to AC).
  2. Replace DC sources with their ideal internal resistance (a short), placing those points at AC ground.
  3. Swap the transistor for its AC model.
  4. Combine and rename resistances where needed.

📐 Resistance naming conventions

Original schematicAC equivalentNotes
RSWrEAC resistance from emitter to AC ground
RC (unloaded) or RC ‖ RL (loaded)rCTotal resistance from collector to AC ground
RB (or R1 ‖ R2 in voltage divider)rBBase biasing resistance
  • Lower case r is used for AC resistance to avoid confusion with DC resistance (upper case R).

📊 Voltage gain characteristics

📊 Voltage gain formula

Voltage gain Av is defined as the ratio of vout to vin:

Av = – rC / (r'e + rE)

  • The negative sign indicates phase inversion (180° shift for sine waves).
  • Gain is essentially a ratio of collector to emitter resistances.
  • Maximum gain occurs when RSW = 0 (emitter completely bypassed), but this increases distortion.

📊 Phase inversion and solutions

  • The common emitter amplifier inverts the waveform top to bottom.
  • For sine waves, this is equivalent to a 180° phase shift.
  • In some applications this is a major issue; in others, not so much.
  • Solution if needed: use a second inverting gain amplifier in sequence (inverting the inversion).

📊 Effect of swamping on gain

  • Larger swamping resistor RSW → lower gain.
  • Gain depends entirely on r'e when RSW = 0, which increases distortion.
  • RSW being much larger than r'e effectively "swamps out" the variation in r'e and reduces distortion.
  • Example from the excerpt: swamped amplifier had gain of –7.1, while completely bypassed (RSW = 0) had gain of –218.2 (over 30 times greater).

🔌 Input and output impedance

🔌 Input impedance formula

Input impedance Zin is the ratio of vin to iin:

Zin(base) = β(r'e + rE)

Zin = rB ‖ Zin(base)

  • Both the swamping resistor and β play a role in setting input impedance.
  • Larger values of RSW and β produce larger input impedances.
  • Example from the excerpt: swamped circuit had Zin = 14.2 kΩ, while completely bypassed had Zin = 5.65 kΩ (less than half).

🔌 Output impedance

Output impedance Zout: the internal impedance of the equivalent source that drives the load.

Zout = r'C ‖ RC

  • Looking back into the amplifier from the load: Cout is shorted ideally and VCC is at AC ground.
  • This leaves RC in parallel with the transistor.
  • The transistor is modeled as a current source with ideal internal resistance approaching infinity; in reality r'C is likely around 100 kΩ.
  • In many circuits, RC is considerably smaller than r'C, therefore Zout ≈ RC.

⚖️ The swamping trade-off

⚖️ Quality versus quantity

The excerpt describes this as a "classic quality versus quantity trade-off: a large low quality gain versus a modest high quality gain."

AspectWith swamping (RSW > 0)Without swamping (RSW = 0)
Voltage gainLower (modest)Higher (maximum)
DistortionLower (better quality)Higher (worse quality)
Input impedanceHigher (more desirable)Lower (less desirable)
  • Swamping decreases voltage gain but reduces distortion and increases input impedance.
  • The latter two are generally desirable for a voltage amplifier.
  • A non-swamped amplifier has the largest gain but suffers from the worst distortion and a low input impedance.

⚖️ Achieving both high gain and low distortion

The excerpt poses the question: "How do we get both high gain and low distortion?"

  • One answer: use multiple low gain stages in cascade (in sequence).
  • This allows each stage to have low distortion while the overall system achieves high gain.

🧮 Calculating loaded versus unloaded gain

🧮 Two methods for voltage gain

The excerpt demonstrates two approaches when a load resistor is present:

Method 1 (direct loaded gain):

  • Include the load resistor as part of rC: rC = RC ‖ RL
  • Calculate Av = – rC / (r'e + rE)
  • Fastest method for a specific load value.

Method 2 (unloaded gain plus divider effect):

  • Find unloaded gain: Av(unloaded) = – RC / (r'e + rE)
  • Calculate voltage divider effect: Adivider = RL / (RL + RC)
  • Composite gain: Av = Av(unloaded) × Adivider
  • Preferred method when swapping out different load values.

🧮 Finding r'e from collector current

To calculate Zin and Av, you need r'e, which requires finding IC:

  1. Use KVL around the base-emitter loop.
  2. Approximate DC base voltage (near zero for two-supply emitter bias).
  3. Calculate IC = (|VEE| – VBE) / (RE + RSW)
  4. Calculate r'e = 26 mV / IC

Example from the excerpt: IC = 0.43 mA led to r'e = 60.5 Ω.

47

Common Collector Amplifier

7.4 Common Collector Amplifier

🧭 Overview

🧠 One-sentence thesis

The common collector amplifier (emitter follower) provides high input impedance, low output impedance, and unity voltage gain to prevent signal loss and match high-impedance sources to low-impedance loads.

📌 Key points (3–5)

  • Primary purpose: reduce impedance loading effects by matching high-impedance sources to low-impedance loads, not to amplify voltage.
  • Key characteristics: high input impedance, low output impedance, non-inverting voltage gain of approximately one.
  • Why "follower": the output voltage follows the input—same voltage level and in phase with the input.
  • Common confusion: this configuration does not produce voltage gain, but it does produce current gain and therefore power gain.
  • Typical applications: high-impedance input buffer stages or drivers for low-impedance loads like loudspeakers.

🔌 Circuit configuration and naming

🔌 What makes it "common collector"

  • The input signal is coupled into the base (like the common emitter amplifier).
  • The output signal is taken at the emitter instead of at the collector.
  • The collector is at AC common (ground), so no collector resistor is needed.
  • Example: In Figure 7.16, a two-supply emitter bias configuration shows the collector connected to AC common while the output comes from the emitter.

🏷️ Why "emitter follower" or "voltage follower"

Emitter follower / voltage follower: the output voltage follows the input voltage—at the same voltage level and in phase.

  • The name describes the behavior: output tracks the input without inversion.
  • "Voltage follower" is the more generic term used across different amplifier types.
  • Don't confuse: "follower" does not mean it copies the input perfectly; it means the gain is approximately one, preventing signal loss rather than amplifying.

⚡ Performance characteristics

⚡ Impedance properties

PropertyValuePurpose
Input impedanceHighMinimizes loading on the source
Output impedanceLowCan drive low-impedance loads effectively
Voltage gainApproximately 1 (non-inverting)Prevents signal loss, not amplification
  • The high input impedance means the amplifier draws minimal current from the source.
  • The low output impedance allows the amplifier to deliver current to demanding loads without voltage drop.

⚡ Gain characteristics

  • Voltage gain: approximately one (unity gain).
  • Current gain: yes, the amplifier does provide current amplification.
  • Power gain: yes, because power equals voltage times current, and current is amplified even though voltage is not.
  • The excerpt emphasizes: "While this configuration does not produce voltage gain, it does produce current gain, and therefore, power gain."

🎯 Applications and design thinking

🎯 Primary use cases

The excerpt identifies two main applications:

  1. High-impedance input buffer stages: prevents a high-impedance source from being loaded down by subsequent stages.
  2. Drivers for low-impedance loads: such as loudspeakers, which require significant current delivery.

🎯 Design philosophy

  • The excerpt advises: "Perhaps the best way to think about the follower is not that it gives a voltage gain of one, but that it will prevent signal loss."
  • This reframes the purpose: the goal is impedance transformation and signal preservation, not amplification.
  • Example: A high-impedance sensor output can be buffered by an emitter follower before being sent to a low-impedance transmission line, preventing signal degradation.

🔧 Analysis approach

🔧 AC emitter resistance

  • The AC emitter resistance, r_E, is determined by either the emitter bias resistor R_E or other circuit elements (the excerpt text cuts off here).
  • This parameter is critical for calculating input impedance and gain, similar to the common emitter amplifier analysis shown earlier in the excerpt.
48

Common Base Amplifier

7.5 Common Base Amplifier

🧭 Overview

🧠 One-sentence thesis

The common collector amplifier (emitter follower) provides unity voltage gain with high input impedance and low output impedance, making it ideal for impedance matching between high-impedance sources and low-impedance loads.

📌 Key points (3–5)

  • Primary function: The common collector acts as a voltage follower with gain ≈ 1, high input impedance, low output impedance, and non-inverting operation.
  • Why it matters: Prevents signal loss and reduces impedance loading effects, serving as a buffer or driver stage.
  • Key characteristic: Output voltage follows the input (same level, in phase) while providing current gain and power gain.
  • Common confusion: It doesn't amplify voltage but prevents voltage loss; the gain approaching unity is the desired goal, not a limitation.
  • Power supply considerations: Bypass capacitors and decoupling networks are essential to prevent noise, ripple, and oscillations from degrading signal quality.

🔌 Power Supply Quality Issues

⚡ AC grounding and bypass capacitors

  • Real DC power sources are not ideal: they may not present a perfect AC ground and may contain ripple or noise.
  • Non-ideal behavior can cause hum and oscillations that diminish output signal quality.

Power supply bypass capacitors: modest-sized capacitors (typically 1 μF or larger) placed physically close to active devices to minimize resistive and inductive effects of circuit board traces.

  • These capacitors ensure the power supply acts as a good AC ground at the device location.
  • Larger capacitors may be needed for high output power amplifiers.

🔇 Decoupling noise and ripple

  • Noise and ripple from the power supply can couple into the input signal and become amplified in the output.
  • This is especially problematic in voltage divider biasing circuits where the divider creates DC potential but also couples in undesirable AC components at the base terminal.
  • The base location is particularly sensitive because signals there get amplified.

Solution approach:

  • Ideal solution: use a high-quality regulated DC supply (but often impractical due to cost).
  • Practical solution: use an RC decoupling network with capacitor C_D and resistor R_3.
    • C_D creates low impedance at the divider junction, shunting noise/ripple to ground.
    • R_3 prevents shorting out the input signal while working in parallel with base input impedance.

🎯 Common Collector Configuration

🏗️ Circuit structure

The common collector amplifier differs from common emitter in key ways:

  • Input couples into the base (like common emitter).
  • Output signal is taken at the emitter instead of the collector.
  • Collector is at AC common, so no collector resistor is needed.
  • Often uses two-supply emitter bias.

🎭 Alternative names and purpose

Emitter follower or voltage follower: names reflecting that output voltage follows the input at the same level and in phase.

Key characteristics:

  • High input impedance
  • Low output impedance
  • Non-inverting voltage gain ≈ 1
  • Produces current gain and therefore power gain (even without voltage gain)

Primary applications:

  • High-Z input buffer stages
  • Drivers for low-impedance loads (e.g., loudspeakers)
  • Impedance matching between mismatched source and load

📐 Voltage Gain Analysis

📊 Gain equation derivation

The voltage gain formula is: A_v = r_E / (r'_e + r_E)

Where:

  • r_E is the AC emitter resistance (either the emitter bias resistor R_E alone, or R_E in parallel with load resistance R_L)
  • r'_e is the dynamic emitter resistance
  • Use R_E alone for unloaded gain; use R_E parallel R_L for loaded gain

Key insight: If r_E is much greater than r'_e, the gain approaches unity (1).

🎯 Design philosophy

  • The best way to think about the follower: it prevents signal loss rather than providing gain of one.
  • Signal distortion tends to be low because unity gain is the desired goal, not a compromise.
  • Output is in phase with input (non-inverting).

🔌 Impedance Characteristics

📥 Input impedance

The input impedance formulas are unchanged from common emitter configuration:

  • Z_in(base) = β × (r'_e + r_E)
  • Z_in = r_B parallel Z_in(base)

Where r_B is typically the base biasing resistor (R_B in two-supply emitter bias, or R_1 parallel R_2 for voltage divider bias).

📤 Output impedance

Output impedance derivation differs considerably from common emitter:

Z_out = R_E parallel Z_out(emitter)

Where Z_out(emitter) = r'_e + Z_B(equivalent)

And Z_B(equivalent) = (r_B parallel r_gen) / β

Final formula: Z_out = R_E parallel [r'_e + (r_B parallel r_gen) / β]

Important note: In many instances the emitter bias resistor is large enough to ignore in the parallel combination.

Why this matters: The division by β in the base network equivalent resistance significantly reduces the output impedance, achieving the low output impedance characteristic needed for driving low-impedance loads.

49

Multi-Stage Amplifiers

7.6 Multi-Stage Amplifiers

🧭 Overview

🧠 One-sentence thesis

Cascading multiple amplifier stages multiplies their individual gains to achieve much higher system gain than any single stage can provide, with each stage serving as the load for the preceding stage.

📌 Key points (3–5)

  • How stages connect: Each stage's input impedance becomes the load resistance for the previous stage; gains multiply together.
  • System impedance rules: System input impedance equals the first stage's input impedance; system output impedance equals the last stage's output impedance.
  • Direct coupling advantage: Eliminates coupling capacitors between stages, reducing component count and enabling DC amplification with no lower frequency limit.
  • Common confusion: The source only "sees" the first stage because it only delivers current to that stage, not to subsequent stages.
  • Gain multiplication power: Even modest per-stage gains (e.g., three stages at 10× each) produce very high system gains (1000×).

🔗 Cascading fundamentals

🔗 How stages interact

  • In a multi-stage amplifier, stages are connected in series (cascaded).
  • Each stage acts as the load for the stage before it.
  • The input impedance Z_in of stage N becomes the load resistance R_L of stage N-1.
  • Example: A two-stage amplifier where stage 1 drives stage 2 means stage 2's input impedance is stage 1's load.

🧮 Gain calculation

  • Individual stage gains multiply together to produce the system gain.
  • If stage 1 has gain A_v1 = 10 and stage 2 has gain A_v2 = 10, system gain = 10 × 10 = 100.
  • The excerpt emphasizes: "three swamped common emitter stages with voltage gains of just 10 each would produce a system voltage gain of 1000."
  • This multiplication allows high overall gain even when individual stages are heavily swamped (to reduce distortion).

🎯 System impedance determination

Input impedance:

  • The system input impedance is the input impedance of the first stage only.
  • The source drives only the first stage and "sees" only that stage's impedance.
  • Later stages do not affect system input impedance.

Output impedance:

  • The system output impedance is the output impedance of the last stage only.
  • Only the final stage directly drives the load.
ParameterDetermined byWhy
System Z_inFirst stage Z_inSource connects only to first stage
System Z_outLast stage Z_outLoad connects only to last stage
System gainProduct of all stage gainsSignal passes through all stages

🔌 Capacitive coupling

🔌 Inter-stage coupling capacitors

Inter-stage coupling capacitor: A capacitor placed between stages to pass AC signals while blocking DC voltages.

  • The coupling capacitor (C_inter in the excerpt's example) prevents DC potential at one stage's collector from interfering with the bias of the next stage.
  • For DC analysis, each stage is treated as a separate circuit.
  • For AC analysis, the capacitor acts as a short circuit, allowing signal to pass.

🧩 Two-stage example breakdown

The excerpt describes a two-stage amplifier (Figure 7.30):

  • Stage 1: Non-swamped common emitter with two-supply emitter bias.
  • Stage 2: Swamped common emitter with voltage divider bias.
  • Coupling: C_inter blocks DC between stages.

Analysis approach:

  1. DC analysis: Treat as two independent circuits.
  2. AC analysis for stage 1: Its load is R_1 || R_2 || Z_in-base2 (the parallel combination of stage 2's biasing resistors and input impedance).
  3. AC analysis for stage 2: Proceeds normally; its gain multiplies stage 1's gain.
  4. System input impedance: R_B || Z_in-base1 (stage 1's input impedance).

⚡ Direct coupling

⚡ What direct coupling means

Direct coupling: A technique that allows DC to flow from stage to stage, eliminating coupling capacitors.

  • With direct coupling, there is no lower frequency limit—the amplifier can amplify DC signals.
  • Fewer components are needed (no coupling capacitors, potentially fewer biasing resistors).
  • Requires careful design so that one stage's DC output properly biases the next stage.

🛠️ How direct coupling works

The excerpt provides a two-stage direct-coupled example (Figure 7.31):

Stage 1:

  • Swamped common emitter using two-supply emitter bias.
  • Uses a Darlington pair for high input impedance.
  • Base current is so low that the DC drop on R_B is negligible, so the input coupling capacitor can be eliminated.

Stage 2:

  • The DC collector voltage from stage 1 is applied directly to the base of stage 2.
  • This DC voltage sets up the bias for stage 2 via its emitter resistors (same principle as voltage divider bias, but the base voltage comes from the previous stage instead of a divider network).
  • Uses a PNP transistor (opposite polarity to stage 1's NPN).

Output coupling elimination:

  • By using a PNP device and a bipolar power supply, the designer can set resistor values so that the DC collector voltage of stage 2 equals 0 V.
  • If there's no DC voltage at the output, there's nothing to block, so no output coupling capacitor is needed.

🎁 Benefits of direct coupling

  • Component reduction: Eliminates coupling capacitors and can eliminate voltage divider resistors.
  • DC amplification: No lower frequency limit; can amplify DC signals.
  • Performance: Can achieve higher performance with fewer parts.
  • Caveat: The excerpt notes that if emitter bypass capacitors are used, DC gain will be less than AC gain, so the amplifier is not "fully DC coupled" in that sense.

🔀 Configuration mixing

🔀 Combining different amplifier types

  • Multi-stage designs can mix different AC configurations.
  • Example from the excerpt: "a common collector follower for the first stage that drives a common emitter voltage amplifier."
  • Can also mix NPN and PNP devices (as in the direct-coupled example).
  • Different biasing types may be used in different stages.

🎯 Strategic stage selection

Each configuration has strengths:

  • Common emitter: High voltage gain, inverts signal.
  • Common collector (follower): High input impedance, low output impedance, unity gain, non-inverting.
  • Common base: High voltage gain, low input impedance, non-inverting.

Example strategy: Use a follower as the first stage to present high input impedance to the source, then use a common emitter stage for voltage gain.


Don't confuse: The system gain is the product of individual stage gains, not the sum. A common error is to add gains instead of multiplying them.

50

BJT Class A Power Amplifiers

Chapter 8: BJT Class A Power Amplifiers

🧭 Overview

🧠 One-sentence thesis

Class A amplifiers allow collector current to flow for the entire 360° cycle and can be analyzed using AC load lines to determine maximum output swing, but they suffer from poor efficiency (maximum 25%) because they draw full power continuously regardless of signal presence.

📌 Key points (3–5)

  • What defines Class A: collector current flows for the entire 360° of the signal cycle without interruption.
  • AC load line purpose: plots all possible transistor voltage-current pairs to determine maximum signal swing (compliance) before clipping occurs.
  • Centered Q-point advantage: placing the Q-point at the center of the AC load line maximizes undistorted output swing; off-center causes earlier clipping on one side.
  • Common confusion: transistor dissipation is highest at idle (no signal) and drops to half at full output—turning up the volume actually cools the transistor.
  • Efficiency limitation: maximum theoretical efficiency is only 25% because the supply must be at least twice V_CEQ and the amplifier always draws full current.

🔌 Class A operation fundamentals

🔌 What Class A means

Class A operation: signal current in the collector flows 360° out of the cycle—it flows for the entire cycle without interruption.

  • This is the defining characteristic that separates Class A from other amplifier classes.
  • All voltage amplifiers from the previous chapter are Class A designs.
  • A single output device can amplify the entire input waveform.
  • Example: if you apply a sine wave input, collector current varies continuously around the Q-point but never stops flowing.

🔄 Comparison with other classes

The excerpt mentions several amplifier classes:

ClassCurrent flowTypical useEfficiency
A360° (continuous)Audio, low power stagesLow (≤25%)
B180° (half cycle)Power amplifiersHigher
CLess than 180°High power telecomHighest
DDiscontinuous (switching)Modern audioVery high
  • Class designation indicates operational principle, not quality or fidelity.
  • As class letter increases, designs become more complicated but more efficient.
  • Don't confuse: "class" refers to how the transistor operates, not the amplifier's sound quality.

📊 AC load line analysis

📊 What the AC load line shows

The AC load line plots all possible combinations of transistor voltage (v_CE) and current (i_C) during signal operation.

  • Similar to DC load line but usually has a steeper slope.
  • Must share the Q-point with the DC load line.
  • Has two endpoints: cutoff voltage v_CE(cutoff) and saturation current i_C(sat).
  • The slope difference occurs because AC resistance is typically less than DC resistance due to capacitor bypassing and loading.

📐 Load line endpoints

The AC load line endpoints are calculated from the Q-point and AC resistances:

Saturation current (maximum current):

  • i_C(sat) = I_CQ + V_CEQ divided by (r_E + r_C)
  • This is the current when the transistor voltage drops to zero.

Cutoff voltage (maximum voltage):

  • v_CE(cutoff) = V_CEQ + I_CQ times (r_E + r_C)
  • This is the voltage when collector current falls to zero.

Where r_E and r_C are the AC emitter and collector resistances (use zero if not present in the circuit).

🎯 Q-point positioning effects

Q-point closer to saturation (Figure 8.3):

  • Output clips when trying to swing toward zero voltage.
  • Maximum unclipped swing is limited to V_CEQ.
  • Saturation clipping distorts the waveform severely.

Q-point closer to cutoff (Figure 8.4):

  • Output clips when trying to swing toward maximum current.
  • Maximum unclipped swing is limited to I_CQ times (r_E + r_C).
  • Cutoff clipping distorts the waveform severely.

Centered Q-point (Figure 8.5):

  • Provides maximum undistorted output swing.
  • Peak voltage swing equals V_CEQ.
  • Peak current swing equals I_CQ.
  • This is the optimal configuration for maximum output.

Don't confuse: clipping on one side versus the other—both are distortion, but centering the Q-point prevents either from occurring prematurely.

🧮 Compliance calculation

Compliance: the maximum undistorted peak output voltage swing.

General rule: Peak compliance is the smaller of V_CEQ or I_CQ times (r_E + r_C).

  • For voltage followers and unswamped amplifiers, load compliance approximately equals transistor compliance.
  • For heavily swamped amplifiers, compliance is reduced by voltage division between load and swamping resistors.
  • Example: an amplifier with voltage gain of 4 loses about 20% of maximum swing due to swamping.

Centered Q-point condition: V_CEQ divided by I_CQ equals (r_E + r_C).

⚡ Power and efficiency analysis

⚡ Maximum load power

Once compliance is known, maximum load power is calculated using RMS values:

  • Convert peak compliance to RMS by multiplying by 0.707 (or dividing by square root of 2).
  • P_load(max) = (Compliance_RMS squared) divided by R_L.
  • Important: use the actual load resistance R_L, not the parallel combination with biasing resistors.
  • Using the parallel combination would incorrectly include power dissipated in biasing resistors.

🔥 Transistor power dissipation

The transistor's power dissipation behaves counterintuitively:

At idle (no signal):

  • P_DQ = V_CEQ times I_CQ
  • This is the maximum transistor dissipation.

At full load:

  • Average dissipation drops to P_DQ divided by 2.
  • The transistor dissipates half the idle power.
  • The other half of P_DQ goes to the load.

Why this happens:

  • Class A amplifiers always draw the same current from the supply (I_CQ).
  • Total supplied power remains constant regardless of signal level.
  • As signal increases, power shifts from transistor to load.
  • At maximum swing, power is split equally between transistor and load.

Don't confuse: higher volume means cooler transistor in Class A—to keep the output transistor cool, turn the volume up, not down.

📉 Efficiency limitations

Class A amplifiers have inherently poor efficiency:

Maximum theoretical efficiency: 25%

Why so low:

  • Best case load power is P_DQ divided by 2.
  • Power supply must be at least twice V_CEQ to cover peak-to-peak swing.
  • Supplied DC power is at least 2 times V_CEQ times I_CQ = 2 times P_DQ.
  • Efficiency = (P_DQ/2) divided by (2 × P_DQ) = 25%.

Actual efficiency is often much worse:

  • Non-centered Q-points reduce efficiency further.
  • AC and DC load lines rarely identical, requiring larger supply voltage.
  • Example 8.1 showed only 5.9% efficiency due to off-center Q-point.

When Class A makes sense:

  • Early stages of multi-stage amplifiers (very low power requirements).
  • Applications where simplicity outweighs efficiency concerns.
  • When output power requirements are small enough that waste heat is manageable.

🔊 Loudspeaker loads

🔊 Dynamic loudspeaker construction

The excerpt describes the most common type: the dynamic loudspeaker.

Key components:

  • Voice coil (H): magnet wire wound around a former, connected to amplifier via lead wires.
  • Permanent magnet (E): creates strong fixed magnetic field (ceramic, alnico, or rare earth).
  • Diaphragm (F): attached to voice coil, pushes air to create sound.
  • Suspension (B) and spider (D): allow voice coil to move freely.
  • Frame (A): holds everything together.

How it works:

  • Current from amplifier flows through voice coil.
  • Voice coil creates its own magnetic field.
  • This field aids or opposes the permanent magnet's field depending on current direction.
  • Resulting force moves the coil and attached diaphragm.
  • Larger current creates stronger field, greater movement, and higher sound pressure.

Design limitations:

  • Very difficult for one driver to cover full audio spectrum (20 Hz to 20 kHz).
  • Specialized drivers used: woofers (low frequency), tweeters (high frequency), midranges (middle frequencies).
  • Conversion efficiency is terrible: only 1% to 2% of electrical power becomes acoustic output.
  • Most applied power simply heats the voice coil.

🌀 Complex impedance characteristics

Loudspeakers do not present a simple resistive load:

Nominal versus actual impedance:

  • Loudspeakers rated with nominal impedance (typically 8Ω home, 4Ω automotive).
  • Actual impedance varies significantly with frequency.
  • Testing with power resistors only provides coarse approximation.

Electrical model (Figure 8.13):

  • Includes voice coil resistance (R_VC) and inductance (L_VC).
  • Additional components represent mechanical properties (suspension losses, etc.).
  • Three parallel elements create resonant peak at low frequencies (f_S, free-air resonance).
  • Series inductance causes impedance to rise with frequency.

Real-world behavior (Figure 8.14):

  • Nominal 8Ω speaker can exceed 30Ω at resonance.
  • Can drop below 7Ω at some frequencies.
  • Phase angle can reach 40° capacitive or inductive.
  • Complete speaker systems (multiple drivers plus crossovers) have even more complex impedance plots.

⚠️ Impact on amplifier operation

Increased current demand:

  • Where impedance drops below nominal, more current needed for given voltage.
  • Transistors must handle higher peak currents than simple resistive analysis suggests.

Phase shift effects (Figure 8.15):

  • Reactive load causes phase shift between voltage and current.
  • Peak current-voltage product can be twice that of resistive load.
  • This combination may exceed transistor's safe operating area.
  • Transistors may need higher ratings than calculated for ideal resistive load.

AC load line with complex impedance (Figure 8.16):

  • Resistive load produces straight-line load line (green).
  • Complex impedance produces elliptical load line (red).
  • As signal increases, operation traces ellipse around Q-point.
  • Impedance angle changes ellipse aspect ratio (0° = straight line, 90° = circle).
  • Loudspeaker impedance changes with frequency, so load line shape changes with frequency.
  • Some operating regions may fall outside transistor's safe operating area.

Don't confuse: nominal impedance rating with actual impedance—the real impedance varies dramatically with frequency and can be much higher or lower than the nominal value.

🌡️ Thermal management

🌡️ Power transistor characteristics

Power transistors use different packaging than small-signal devices:

TO-3 case example (2N3055):

  • All-metal case instead of plastic TO-92.
  • Only two leads shown (base and emitter).
  • Entire body is the collector for maximum heat transfer contact.
  • Maximum ratings: 115W at 25°C case temperature, 15A collector current, 60V collector-emitter voltage.
  • Cannot withstand maximum current and voltage simultaneously.

Performance differences from small-signal:

  • Beta (current gain) considerably lower, can drop below 20 at high currents.
  • I_C(sat) tends to be larger, can exceed 0.5V.
  • Safe operating area limits combinations of V_CE and I_C.
  • Safe zone extends further for short pulses versus continuous operation.

📉 Power derating

Transistor power capability decreases as temperature rises:

Derating curve (Figure 8.18):

  • 2N3055 rated for 115W only at case temperatures of 25°C or lower.
  • At 100°C, can only dissipate about 65W.

Calculation formula: P_D = P_25 minus D times (T_case minus 25°C)

Where:

  • P_D is power dissipation at new case temperature.
  • P_25 is power dissipation at 25°C.
  • D is derating factor (watts per degree Celsius).
  • T_case is new case temperature.

Example: 2N3055 at 75°C with derating factor 0.657 W/°C yields 82.1W capability (down from 115W at 25°C).

🧊 Heat sink fundamentals

Heat sink: a metal device attached to the power transistor to efficiently move heat away from it.

Construction and mounting:

  • Typically aluminum with array of fins to increase surface area.
  • Designed for specific case styles (TO-3, TO-220, TO-202, etc.).
  • Requires insulation spacers and special hardware to maintain electrical isolation.
  • Mica sheet and plastic washers/bushings prevent heat sink from becoming electrically live.

Best practices:

  • Always use heat sink grease or thermally conductive pad between device and sink.
  • Excessive grease decreases performance—use appropriate amount.
  • Mount fins vertically for optimum natural convective cooling.
  • Do not overcrowd or obstruct devices using heat sinks.
  • Do not block airflow, especially directly above and below.
  • Consider forced convection (fan) for high thermal demands.

🔬 Thermal resistance model

Thermal resistance (θ): measure of how easy it is to transfer heat energy from one part to another; units are Celsius degrees per watt.

Thermal circuit analogy:

  • Temperature is analogous to voltage.
  • Thermal power dissipation is analogous to current.
  • Thermal resistance is analogous to electrical resistance.

Basic equation: P_D = ΔT divided by θ_total (thermal version of Ohm's law).

Thermal path (Figures 8.21 and 8.22):

  • Junction temperature (T_j) created by transistor current times voltage.
  • θ_jc: thermal resistance from junction to case.
  • θ_cs: thermal resistance from case to heat sink (via mounting interface).
  • θ_sa: thermal resistance from heat sink to ambient air.
  • All three resistances in series from junction to ambient.

Heat sink selection formula: P_D = (T_j minus T_a) divided by (θ_jc + θ_cs + θ_sa)

Rearranged to find required heat sink rating: θ_sa = [(T_j minus T_a) divided by P_D] minus θ_jc minus θ_cs

Known values:

  • T_j and θ_jc given by semiconductor manufacturer.
  • T_a determined experimentally (usually higher than room temperature due to localized warming).
  • θ_cs determined from standard graphs (Figure 8.23) based on case style and insulation.

Example (Example 8.3):

  • Device: T_j(max) = 175°C, TO-3 case, θ_jc = 1.5 C°/W.
  • Conditions: 15W dissipation, 40°C ambient, 0.002 mica insulator with grease.
  • From graph: θ_cs ≈ 0.35 C°/W.
  • Required: θ_sa = (175 - 40)/15 - 1.5 - 0.35 = 7.15 C°/W maximum.
  • At 40W dissipation, required θ_sa drops to 1.53 C°/W, requiring forced air cooling or much larger heat sink.

Don't confuse: lower thermal resistance is better—it means heat transfers more easily, keeping the junction cooler.

51

BJT Class A Power Amplifiers

8.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

Class A amplifiers allow signal current to flow continuously for the entire cycle, and their maximum output swing and power are determined by the AC load line, which shares the Q point with the DC load line but typically has a steeper slope due to lower AC resistance.

📌 Key points (3–5)

  • Class A definition: signal current in the collector flows 360° (the entire cycle) without interruption.
  • AC vs DC load lines: both share the Q point, but the AC load line is usually steeper because AC resistance is lower (due to loading and capacitor bypassing).
  • AC load line endpoints: cutoff voltage and saturation current are determined by the Q point and AC resistances (rC + rE).
  • Common confusion: AC load line vs DC load line—the AC line reflects the signal path (with bypassed components and AC loads), while the DC line reflects biasing; they intersect only at the Q point.
  • Why it matters: the AC load line determines compliance (maximum output voltage swing), maximum load power, efficiency, and device dissipation requirements.

🔌 Amplifier classes and Class A definition

🔌 What amplifier class means

  • The class of an amplifier indicates its fundamental operational principle, not its fidelity or quality.
  • As the class letter increases (A → B → C → D), designs become more complicated but also more efficient.
  • Common classes for audio and linear applications: A, B, and D.
  • Class C is used mainly for high-power telecommunications; classes G and H are variations of class B.

⚡ Class A operation

Class A: signal current in the collector flows 360° out of the cycle—it flows for the entire cycle without interruption.

  • All amplifiers presented in the prior chapter (Chapter 7) are class A designs.
  • In contrast:
    • Class B: collector current flows for just 180°.
    • Class D: collector current is discontinuous; the transistor is used as a switch.
  • Don't confuse: class A means continuous current flow, not necessarily higher quality; it is a design principle, not a performance metric.

📐 AC load line construction

📐 Why the AC load line is needed

  • To determine the maximum load voltage swing (compliance), we analyze the AC equivalent of the amplifier.
  • The AC equivalent includes both AC collector resistance (rC) and AC emitter resistance (rE).
  • It applies to swamped or unswamped common emitter amplifiers and emitter followers.
  • If a resistance is not used (e.g., rC in a follower), substitute zero for it.

📏 AC vs DC load lines

FeatureDC load lineAC load line
PurposeAnalyzes biasing circuitsDetermines signal swing and compliance
SlopeLess steep (higher DC resistance)Steeper (lower AC resistance due to loading and bypassing)
EndpointsVCE(cutoff), IC(sat)vCE(cutoff), iC(sat)
Shared pointQ pointQ point
  • Both load lines must share the Q point (the no-signal operating point).
  • The AC load line slope is usually steeper because AC resistance tends to be less than DC resistance.
  • Consequently:
    • vCE(cutoff) (AC) tends to be smaller than VCE(cutoff) (DC).
    • iC(sat) (AC) tends to be larger than IC(sat) (DC).

🧮 AC load line endpoint formulas

The AC equivalent circuit starts with no-signal values: collector current ICQ and transistor voltage VCEQ.

🔺 Saturation current iC(sat)

  • As the input signal grows, collector current iC increases.
  • This increases voltage drops across rE and rC (by Ohm's law).
  • By Kirchhoff's voltage law (KVL), vCE must decrease.
  • The collector current can only increase until vCE drops to 0 V.
  • Maximum increase in current: VCEQ divided by (rC + rE).

Formula:

iC(sat) = ICQ + VCEQ / (rE + rC)

🔻 Cutoff voltage vCE(cutoff)

  • The transistor starts with VCEQ and ICQ.
  • The largest vCE increase occurs if current falls to zero.
  • Then, all potential originally developed across rE and rC by ICQ must be absorbed by the transistor.

Formula:

vCE(cutoff) = VCEQ + ICQ (rE + rC)

Example: If the Q point is set with VCEQ = 6 V, ICQ = 10 mA, rE = 100 Ω, and rC = 400 Ω, then:

  • iC(sat) = 10 mA + 6 V / (100 Ω + 400 Ω) = 10 mA + 12 mA = 22 mA.
  • vCE(cutoff) = 6 V + 10 mA × (100 Ω + 400 Ω) = 6 V + 5 V = 11 V.

🎯 Q point positioning

🎯 Three possible configurations

The Q point can be positioned in three ways on the AC load line:

  1. Q point closer to saturation.
  2. Q point closer to cutoff.
  3. Q point centered on the AC load line.

📉 Q point closer to saturation

  • The excerpt mentions plotting input voltage (in red) and corresponding collector current and collector-emitter voltage (in blue).
  • When the Q point is closer to saturation, the available swing toward saturation is limited.
  • This configuration restricts the maximum positive excursion of the output signal.
  • Don't confuse: a Q point closer to saturation does not mean higher power; it means less headroom for positive signal swings.

(Note: The excerpt ends mid-sentence; further details on Q point positioning are not provided.)

🔧 Analysis goals for Class A amplifiers

🔧 What we focus on

  • Compliance: maximum output voltage swing.
  • Maximum load power: the largest power that can be delivered to the load.
  • Device dissipation requirements: how much power the transistor must handle.
  • Amplifier efficiency: ratio of useful output power to total power consumed.

🔧 What we ignore (for power analysis)

  • Input impedance.
  • Voltage gain.
  • r'e (the small-signal emitter resistance)—considered simply as a source of distortion.
  • Most power amplifiers are configured as voltage followers, so the focus is on output swing and power, not gain.
52

Class A Amplifier Operation and Load Line Analysis

8.1 Introduction

🧭 Overview

🧠 One-sentence thesis

Class A amplifiers achieve maximum undistorted output voltage swing (compliance) when the Q point is centered on the AC load line, allowing equal excursions toward both saturation and cutoff.

📌 Key points (3–5)

  • Class A definition: signal current in the collector flows continuously for the entire 360° cycle without interruption.
  • AC vs DC load lines: both share the Q point, but the AC load line is typically steeper due to lower AC resistance from loading and capacitor bypassing.
  • Q point positioning matters: placing the Q point closer to saturation or cutoff causes clipping on one side; centering it maximizes undistorted swing.
  • Common confusion: the AC load line endpoints (v_CE(cutoff) and i_C(sat)) differ from the DC endpoints (V_CE(cutoff) and I_C(sat)) because AC resistance ≠ DC resistance.
  • Compliance rule: peak compliance is the smaller of V_CEQ or I_CQ(r_E + r_C), determining the maximum undistorted output voltage.

🔌 Class A amplifier fundamentals

🔌 What defines class A operation

Class A: signal current in the collector flows 360° out of the cycle—it flows for the entire cycle without interruption.

  • All amplifiers presented in the prior chapter are class A designs.
  • The transistor conducts continuously throughout the entire waveform cycle.
  • Contrast with other classes:
    • Class B: i_C flows for just 180°.
    • Class D: i_C is discontinuous; the transistor is used as a switch.
    • Class C: largely relegated to high power telecommunications.

🧩 AC equivalent circuit

  • The generic AC equivalent includes both AC collector resistance (r_C) and AC emitter resistance (r_E).
  • This structure applies to:
    • Swamped or unswamped common emitter amplifiers.
    • Emitter followers (set r_C = 0 if not used).
  • Voltage polarities and current direction are shown for a positive input voltage.
  • The circuit operates around a no-signal current I_CQ and no-signal transistor voltage V_CEQ.

📈 AC load line construction

📈 How AC and DC load lines relate

FeatureDC load lineAC load line
Cutoff voltageV_CE(cutoff)v_CE(cutoff) (typically smaller)
Saturation currentI_C(sat)i_C(sat) (typically larger)
Shared pointQ pointQ point
SlopeFlatterSteeper (due to lower AC resistance)
  • Both load lines must share the Q point in common.
  • The AC load line slope is usually steeper because AC resistance tends to be less than DC resistance due to loading and capacitor bypassing.
  • Consequently, v_CE(cutoff) tends to be smaller than V_CE(cutoff) and i_C(sat) tends to be larger than I_C(sat).

🔢 AC load line endpoint formulas

Saturation current endpoint:

  • As input signal grows, i_C increases, increasing voltage drops across r_E and r_C (Ohm's law).
  • This forces v_CE to decrease (KVL).
  • Collector current can only increase until v_CE drops to 0 V.
  • Maximum increase is V_CEQ / (r_C + r_E).
  • Formula: i_C(sat) = I_CQ + V_CEQ / (r_E + r_C)

Cutoff voltage endpoint:

  • The transistor starts with V_CEQ and I_CQ.
  • Largest v_CE increase occurs if current falls to zero.
  • Then all potential originally developed across r_E and r_C by I_CQ must be absorbed by the transistor.
  • Formula: v_CE(cutoff) = V_CEQ + I_CQ(r_E + r_C)

⚖️ Q point positioning and clipping

⚖️ Q point closer to saturation

  • As input signal increases, output is limited at zero for v_CE and at i_C(sat) for i_C.
  • Both collector current and collector-emitter voltage waveforms are severely clipped and distorted.
  • Largest unclipped peak voltage swing: V_CEQ.
  • Largest peak current swing: i_C(sat) − I_CQ, or equivalently V_CEQ / (r_E + r_C).
  • Example: if the input voltage continues to increase beyond the saturation limit, the output waveforms are clipped at the top, producing gross distortion.

⚖️ Q point closer to cutoff

  • Shifting the Q point toward cutoff solves the saturation clipping problem but creates a new problem: cutoff clipping.
  • Largest unclipped peak voltage swing: v_CE(cutoff) − V_CEQ, or alternately I_CQ(r_E + r_C).
  • Largest peak current swing: I_CQ.
  • The waveform is clipped on the cutoff side—it doesn't matter which side is clipped; either way it's gross distortion.

🎯 Centered Q point (optimal)

  • A centered Q point on the AC load line produces the largest unclipped voltage swing.
  • Largest unclipped peak voltage swing: V_CEQ.
  • Largest unclipped peak current swing: I_CQ.
  • Centering condition: V_CEQ / I_CQ = r_E + r_C.
  • Don't confuse: centering on the AC load line is different from centering on the DC load line; the AC line is steeper, so the condition depends on AC resistances.

📊 Compliance and maximum power

📊 Peak compliance calculation

Peak compliance: the maximum undistorted output voltage swing, determined by the smaller of V_CEQ or I_CQ(r_E + r_C).

  • Most times the maximum load voltage (compliance) equals the maximum transistor voltage.
  • This is the case in voltage followers and unswamped amplifiers.
  • Noticeable reduction occurs only with very heavily swamped amplifiers, where compliance is reduced by the voltage divider between load and swamping resistors.
  • Example: a swamped amplifier with voltage gain of 4 would lose about 20% of maximum swing; swamping has to be very heavy (resulting in very low gains) before appreciable signal is lost.

⚡ Maximum load power

  • Power is determined using RMS values, so peak compliance must be divided by square root of 2 (or multiplied by 0.707).
  • Formula: P_load(max) = (Compliance_RMS)² / R_L
  • Important: use the load resistance value R_L, not the total AC effective value r_L (which is R_L in parallel with a biasing resistor).
  • If r_L were used, the calculation would include power in the load plus power in the biasing resistor, which is incorrect.

🔥 Transistor power dissipation

  • Under no-signal conditions, the transistor operates statically at the Q point.
  • Quiescent power dissipation: P_DQ = V_CEQ × I_CQ
  • At full load for a centered Q point:
    • v_CE = V_CEQ(1 − sin 2πft)
    • i_C = I_CQ(1 + sin 2πft)
  • The excerpt notes that maximum transistor power dissipation does not occur at maximum load power (contrary to intuition), but the full derivation is cut off.
53

Amplifier Classes: Q-Point Positioning and Class A Efficiency

8.2 Amplifier Classes

🧭 Overview

🧠 One-sentence thesis

Centering the Q-point on the AC load line maximizes unclipped output swing in amplifiers, but class A designs remain inherently inefficient—dissipating maximum power at idle and achieving at best 25% efficiency—making them practical only for low-power applications despite their simplicity.

📌 Key points (3–5)

  • Q-point position determines clipping type: shifting the Q-point toward saturation causes saturation clipping; shifting toward cutoff causes cutoff clipping; centering maximizes unclipped swing.
  • Peak compliance rule: the maximum unclipped output swing is the smaller of V_CEQ or I_CQ(r_E + r_C).
  • Class A power paradox: the transistor dissipates more power at idle (P_DQ) than at full load (P_DQ/2), because power shifts to the load as signal amplitude increases.
  • Common confusion—efficiency vs. dissipation: class A amplifiers always draw full power from the supply regardless of signal presence, yielding a theoretical maximum efficiency of only 25%.
  • When to use class A: suitable for early stages of multi-stage amplifiers where load power is very small, because increased complexity of more efficient designs is not cost-effective.

🎚️ Q-Point positioning and clipping

📉 Q-point closer to saturation

When the Q-point is shifted toward saturation:

  • The output signal is limited at zero for v_CE and at i_C(sat) for i_C.
  • The waveforms are severely clipped and distorted on the saturation side.
  • Largest unclipped swings:
    • Peak voltage swing: V_CEQ
    • Peak current swing: i_C(sat) − I_CQ, or equivalently V_CEQ / (r_E + r_C)

Example: If the Q-point is too close to saturation, increasing the input signal causes the output to flatten at the saturation limit before reaching the cutoff limit.

📈 Q-point closer to cutoff

When the Q-point is shifted toward cutoff:

  • Cutoff clipping occurs instead of saturation clipping.
  • Largest unclipped swings:
    • Peak voltage swing: v_CE(cutoff) − V_CEQ, or alternately I_CQ(r_E + r_C)
    • Peak current swing: I_CQ

The excerpt emphasizes: "It doesn't really matter which side has been clipped, either way it's gross distortion."

⚖️ Centered Q-point (optimal)

To produce the largest unclipped voltage swing, the Q-point must be centered on the AC load line.

When centered:

  • Largest unclipped peak voltage swing: V_CEQ
  • Largest unclipped peak current swing: I_CQ
  • Centering condition: V_CEQ / I_CQ = r_E + r_C

Don't confuse: Maximum transistor voltage vs. maximum load voltage. Most times they are equal (voltage followers, unswamped amplifiers), but heavily swamped amplifiers reduce compliance due to voltage division between load and swamping resistors. Example: A swamped amplifier with voltage gain of 4 loses about 20% of maximum swing.

📏 Compliance and maximum load power

📐 Peak compliance formula

Peak compliance is the smaller of V_CEQ or I_CQ(r_E + r_C).

  • This determines the maximum unclipped output swing.
  • The smaller value limits the swing in one direction.

⚡ Maximum load power calculation

Once compliance is known:

  1. Convert peak compliance to RMS: divide by √2 (or multiply by 0.707).
  2. Apply power law: P_load(max) = (Compliance_RMS)² / R_L

Important: Use the load resistance R_L, not the total AC effective value r_L (which includes parallel biasing resistors). Using r_L would calculate power in the load plus power in the biasing resistor.

Example: For compliance = 3 V peak and R_L = 32 Ω:

  • Compliance_RMS = 0.707 × 3 V = 2.12 V
  • P_load(max) = (2.12 V)² / 32 Ω = 141 mW

🔥 Transistor power dissipation in class A

🔋 Quiescent (no-signal) dissipation

Under no-signal conditions, the transistor operates statically at the Q-point:

  • P_DQ = V_CEQ × I_CQ

This is the maximum transistor dissipation.

📊 Full-load dissipation

At full load with a centered Q-point:

  • v_CE = V_CEQ(1 − sin 2πft)
  • i_C = I_CQ(1 + sin 2πft)
  • Instantaneous power: P_D = v_CE × i_C = V_CEQ × I_CQ × (1 − sin² 2πft)
  • Simplifies to: P_D = V_CEQ × I_CQ × (0.5 + 0.5 cos 4πft)
  • This equals: P_D = P_DQ/2 + (P_DQ/2) cos 4πft

The first term is a fixed offset; the second is a sinusoid at twice the signal frequency with peak amplitude equal to the offset. Average over time is simply the offset value: P_DQ/2.

🌡️ The class A cooling paradox

The transistor only dissipates half the power at full load that it dissipates under idle conditions.

Why this makes sense:

  • The class A amplifier always draws the same power from the DC supplies, regardless of load signal size.
  • Without clipping, average current remains I_CQ.
  • As signal amplitude increases, more power shifts from the transistor to the load.
  • At maximum load swing, both transistor and load dissipate P_DQ/2.

Counterintuitive result: To keep the output transistor of a class A amplifier cool, turn the volume up, not down.

⚙️ Class A efficiency analysis

📉 Maximum theoretical efficiency

Best-case scenario (centered Q-point):

  • Maximum load power: P_DQ/2
  • Power supply must be at least twice V_CEQ (to cover peak-to-peak swing)
  • DC power: P_DC = 2 × V_CEQ × I_CQ = 2 × P_DQ
  • Efficiency: η = P_out / P_in = (P_DQ/2) / (2 × P_DQ) = 25%

This represents the maximum efficiency for an RC-coupled class A amplifier. Actual efficiency may be considerably less depending on biasing.

⚠️ The Achilles heel of class A

CharacteristicClass A behaviorConsequence
Power drawFull power from supply regardless of signal presenceWasteful
Best-case efficiencyOnly 25%At best, only one quarter of supply power becomes useful load power
Transistor dissipationAt least twice delivered load powerRequires robust heat management

✅ When class A is justified

Despite inefficiency, class A is useful when:

  • Large output powers are not needed.
  • Early stages of multi-stage amplifiers: Load power is very small (just power to the following stage), so increased complexity of more efficient designs is not warranted or cost-effective.
  • Simplicity is valued: Class A is a relatively simple design.

🧮 Worked example analysis

🔢 Example circuit parameters

Given circuit (Figure 8.7):

  • V_EE = 15 V, V_BE = 0.7 V, R_E = 120 Ω
  • V_CEQ = 5.7 V (by inspection)
  • Load resistance: 32 Ω

Step 1—Calculate I_CQ:

  • I_CQ = (|V_EE| − V_BE) / R_E = (15 V − 0.7 V) / 120 Ω = 119 mA

Step 2—Calculate AC cutoff voltage:

  • v_CE(cutoff) = V_CEQ + I_CQ(r_C + r_E)
  • r_C + r_E = 0 + (120 Ω || 32 Ω) = 25.3 Ω
  • v_CE(cutoff) = 5.7 V + 119 mA × 25.3 Ω = 5.7 V + 3 V = 8.7 V

Step 3—Determine compliance:

  • V_CEQ = 5.7 V
  • I_CQ(r_C + r_E) = 3 V
  • Compliance = smaller value = 3 V peak

📈 Power and efficiency results

Maximum load power:

  • P_load(max) = (0.707 × 3 V)² / 32 Ω = 141 mW
  • Not much for a loudspeaker, but fair for headphones.

Worst-case transistor dissipation:

  • P_D(max) = P_DQ = I_CQ × V_CEQ = 119 mA × 5.7 V = 678 mW

Supplied DC power:

  • P_DC = I_CQ × (V_CC − V_EE) = 119 mA × 20 V = 2.38 W

Efficiency:

  • η = P_load(max) / P_DC = 141 mW / 2.38 W = 5.9%

This is much worse than the theoretical 25% best case, due at least in part to the Q-point not being centered on the AC load line.

Transistor ratings:

  • Breakdown voltage (BV_CEO) should be at least v_CE(cutoff) = 8.7 V
  • Maximum current rating should be at least i_C(sat) = 119 mA + 5.7 V / 25.3 Ω = 344 mA

💻 Computer simulation insights

🎛️ Emitter follower simulation (Darlington pair)

Circuit parameters:

  • V_EE = 10 V, V_BE = 1.4 V (Darlington), R_E = 330 Ω, Load = 50 Ω
  • I_CQ = (10 V − 1.4 V) / 330 Ω = 26 mA
  • V_CEQ = 6.4 V (emitter is −1.4 V, collectors tied to V_CC = +5 V)
  • v_CE(cutoff) − V_CEQ = 26 mA × (330 Ω || 50 Ω) = 26 mA × 43.4 Ω = 1.13 V

Q-point is not centered: closer to cutoff.

  • Cutoff clipping expected around 1.1 V
  • Saturation clipping expected around 6 V
  • More room for current to swing up to saturation than down to zero

Simulation results (two volt peak input):

  • Negative portion of load voltage clips at approximately 1.1 V as expected.
  • Input not large enough to cause saturation clipping.
  • Voltage gain approximately 0.95 (very close to unity, as expected for a follower).

🔄 Voltage amplifier simulation

Modified circuit to produce voltage amplifier with gain ≈ 1:

  • Load moved to collector
  • 330 Ω biasing resistor added (same AC load impedance)
  • V_CC raised by 10 V to maintain similar V_CEQ
  • Original 330 Ω emitter resistor split: 287 Ω and 43 Ω (same I_CQ, unity gain)

Expected behavior: Clipping at approximately 1.1 V on the positive portion (waveforms flipped vertically compared to follower).

🔍 Clamping action observation

When input level is increased to see clipping on the other half:

  • At first appears never to clip.
  • Careful examination reveals clamping action (presented in Chapter 3).
  • Waveform shifts.
  • Peak-to-peak value close to v_CE(cutoff).
  • Slightly less due to V_CE(sat) not being 0 V, particularly for Darlington pairs.

🔊 Loudspeaker fundamentals

🧲 Dynamic loudspeaker construction

The most common form of loudspeaker is the dynamic loudspeaker.

All dynamic loudspeakers share common elements regardless of size or acoustic capability. Key components:

  • Voice coil (H): coil of magnet wire wound around a former (G), typically aluminum or high-temperature material; may be single-layer edge-wound ribbon or several layers of round wire; diameter ranges from a fraction of an inch to several inches.
  • Diaphragm (F): attached to the voice coil.
  • Suspension: outer edge suspension (B) and inner spider (D) freely suspend the voice coil.
  • Permanent magnet (E): creates strong fixed magnetic field; commonly ceramic, alnico, or rare earth construction.
  • Frame (A): voice coil ends connect to flexible lead wires (C) terminating on the frame—where the amplifier connects.

⚡ Operating principle

Based on magnetic repulsion and attraction:

  1. Current from amplifier flows through voice coil.
  2. Voice coil creates its own magnetic field.
  3. This field either aids or opposes the fixed field from the permanent magnet, depending on current direction.
  4. Resulting force causes coil to move within the fixed field.
  5. Diaphragm moves with coil, pushing air and creating sound.
  6. Larger current → stronger field → greater movement → larger sound pressure.

Historical note: This fundamental design has changed little since its invention in the 1920s. Modern materials (magnets, suspension, diaphragm) have improved, but the operational principle remains the same.

🎵 Driver specialization

It is very difficult to create a driver that can cover the full audio spectrum of 20 Hz to 20 kHz while achieving sufficient listening volume at low distortion.

Drivers are often designed for limited portions of the audio spectrum:

Driver typeFrequency rangeCommon name
Low frequencyBassWoofers
Mid frequencyMiddle rangeMidranges
High frequencyTrebleTweeters
54

Class A Operation and Load Lines

8.3 Class A Operation and Load Lines

🧭 Overview

🧠 One-sentence thesis

Class A amplifiers exhibit unique power dissipation characteristics that are significantly affected by reactive loads such as loudspeakers, requiring careful transistor selection to stay within safe operating areas.

📌 Key points (3–5)

  • Clipping behavior: Class A amplifiers clip asymmetrically, and increasing input level can cause waveform clamping and shifting rather than symmetric clipping on both halves.
  • Reactive loads vs resistive loads: Loudspeakers present complex impedances with phase shifts that can double peak transistor power dissipation compared to purely resistive loads.
  • AC load lines with complex impedances: The load line becomes an ellipse instead of a straight line when the load has a phase angle, creating new operating regions that may exceed safe operating area.
  • Common confusion: Nominal loudspeaker impedance (e.g., 8 Ω) is not constant—actual impedance varies greatly with frequency, ranging from below nominal to many times higher, with significant phase angles.
  • Power derating: Power transistors can only dissipate their rated maximum power at low case temperatures; capability decreases substantially at higher temperatures.

🔊 Loudspeaker characteristics

🔊 Dynamic loudspeaker construction

Dynamic loudspeaker: a speaker that uses magnetic repulsion and attraction to convert electrical signals into sound.

  • Core component is the voice coil (magnet wire wound around a former), connected to flexible lead wires that terminate on the frame where the amplifier connects.
  • The voice coil is fixed to a diaphragm and suspended by an outer edge suspension and an inner spider.
  • A powerful permanent magnet (ceramic, alnico, or rare earth) creates a fixed magnetic field.
  • When amplifier current flows through the coil, it creates its own magnetic field that either aids or opposes the fixed field, causing the coil and diaphragm to move and push air to create sound.
  • Larger current → stronger field → greater movement → larger sound pressure.
  • Example: The fundamental design has changed little since the 1920s; modern improvements are mainly in materials, not operational principle.

🎵 Driver types and efficiency

  • Frequency coverage: It is very difficult to create a single driver covering the full audio spectrum (20 Hz to 20 kHz) at sufficient volume and low distortion.
  • Specialized drivers:
    • Low frequency drivers = woofers
    • High frequency drivers = tweeters
    • Middle range drivers = midranges (formerly called squawkers)
  • Low efficiency: Virtually all direct radiating dynamic loudspeaker systems suffer from low conversion efficiency—only about 1% to 2% of applied electrical power becomes useful acoustic output; the vast majority simply heats the voice coil.

🧲 Complex impedance behavior

Nominal impedance: a reference value (commonly 8 Ω for home use, 4 Ω for automotive) that does not represent the true impedance at all frequencies.

  • Why it's complex: Electrical and mechanical characteristics combine to create an equivalent circuit with resistive, inductive, and capacitive elements.
  • The electrical model includes:
    • Voice coil resistance and inductance (R_VC and L_VC)
    • Electrical equivalents of mechanical properties (e.g., suspension losses)
    • Three parallel elements that create a resonant peak
    • Series inductance that causes impedance to rise with frequency
  • Resonant peak: Typically occurs at the lower end of the spectrum at the free-air resonance frequency (f_S); for a nominal 8 Ω woofer, peak impedance can exceed 30 Ω.
  • Frequency variation: At some frequencies, impedance can drop below 7 Ω; at others, it can be many times the nominal value.
  • Phase angle: Can be upwards of 40° capacitive or inductive, depending on frequency.
  • Don't confuse: A large power resistor is only a coarse approximation of a real loudspeaker; it does not capture the frequency-dependent impedance and phase behavior.

⚡ Impact of reactive loads on amplifier operation

⚡ Increased power dissipation with phase shift

  • Resistive load baseline: With a purely resistive load, the transistor exhibits a certain power dissipation pattern.
  • Reactive load effect: Adding a noticeable phase shift to simulate a partly reactive load causes the peak current-voltage product to become twice the value seen with a purely resistive load.
  • Safe operating area risk: This combination might lay outside the safe operating area of the transistor.
  • Design implication: Transistors may need to be rated higher than the values computed for an idealized resistive load.
  • Example: If a resistive load calculation suggests a certain transistor rating, the actual reactive loudspeaker load may require a transistor with double the power handling capability.

📐 AC load line with complex impedance

  • Resistive load line: A straight line from the Q point to the axes; as signal amplitude increases, the operating point swings along this line until it maxes out at the two axes.
  • Complex impedance load line: The straight line becomes an ellipse around the Q point.
    • Signal starts at zero amplitude (just the Q point).
    • As signal increases, it traces out an ellipse around the Q point.
    • Further increases create larger ellipses until the maximum swing just touches the axes (full compliance).
  • Aspect ratio changes: If the impedance angle changes, the aspect ratio of the ellipse changes.
    • Larger angle → more open ellipse
    • 0° (purely resistive) → collapsed ellipse (straight line)
    • 90° (purely reactive) → fully open ellipse (circle)
  • Frequency-dependent behavior: As input frequency sweeps from low to high, the load line wavers back and forth between straight lines and various elliptical shapes.
  • Critical concern: Some new operating regions (where the ellipse curve is above and to the right of the resistive load line) may go outside the safe operating area of the transistor.

🔧 Class A amplifier clipping and clamping

🔧 Asymmetric clipping behavior

  • Expected clipping: With modified circuit values (resistors of 287 Ω and 43 Ω) to achieve unity voltage gain, clipping is expected at approximately 1.1 volts on the positive portion.
  • Strange behavior at higher input: If the input level is increased to attempt to see clipping on the other half of the waveform, it initially appears as though it never clips.
  • Clamping action: Careful examination reveals that the circuits exhibit a certain amount of clamping action (a concept presented in Chapter 3), which causes the waveform to shift.
  • Peak-to-peak value: Inspecting the peak-to-peak value shows it will be close to the value of v_CE(cutoff), slightly less because V_CE(sat) is not 0 V (particularly for a Darlington pair).
  • Don't confuse: Increasing input level does not produce symmetric clipping on both halves; instead, clamping shifts the waveform.

🔥 Power transistor ratings and derating

🔥 Maximum ratings and case design

  • Example device: The 2N3055 NPN power transistor has:
    • Maximum power dissipation: 115 W at case temperature of 25°C
    • Maximum collector current: 15 A
    • Maximum collector-emitter voltage: 60 V
  • Important limitation: The device cannot withstand maximum current and voltage simultaneously.
  • Case design (TO-3): Only two leads are shown (emitter and base); the entire body of the device is the collector.
    • Reason: The device will most likely be attached to a metal heat sink to help dissipate generated heat.
    • Greater contact area → more effective heat flow.
  • Lower beta: Power transistors have considerably lower β than small signal devices; β might fall to less than 20 for very high currents.
  • Higher saturation current: I_C(sat) tends to be larger for higher power transistors, upwards of half a volt.

🌡️ Power derating with temperature

Power derating curve: a graph showing how maximum power dissipation capability decreases as case temperature increases.

  • Rated power limitation: Although the 2N3055 is rated for 115 watts, that is only true at case temperatures of 25°C or lower.
  • Temperature effect: At higher temperatures, power dissipation capability decreases.
  • Example: At 100°C, the 2N3055 can only dissipate about 65 watts.
  • Calculation formula: P_D = P_25 − D × (T_case − 25°C), where D is the derating factor.
  • Don't confuse: The maximum power rating on the datasheet is not a constant; it applies only at or below 25°C case temperature.

📊 Safe operating area

  • Safe operating area (SOA): A plot showing the combination of V_CE and I_C that must fall within a defined zone (lower-left region).
  • Pulse vs continuous: The safe zone extends out further if the current/voltage combination results from a short pulse rather than a continuous condition.
  • Design constraint: The operating point must remain within the SOA boundary to prevent transistor damage.
Load typeLoad line shapePeak power dissipationDesign challenge
Purely resistiveStraight lineBaselineSimple calculation
Complex (reactive)EllipseUp to 2× baselineMay exceed SOA; requires higher-rated transistors
55

Loudspeakers and Power Transistors

8.4 Loudspeakers

🧭 Overview

🧠 One-sentence thesis

Complex loudspeaker impedance causes the AC load line to trace elliptical paths that can push transistors outside their safe operating area, requiring careful attention to power ratings, thermal management, and heat sink design.

📌 Key points (3–5)

  • Complex impedance effect: Loudspeakers create elliptical AC load lines (not straight lines) because their impedance has both magnitude and phase angle that varies with frequency.
  • Safe operating area: The combination of collector-emitter voltage and collector current must stay within specified limits, which change based on whether operation is continuous or pulsed.
  • Power derating: Power transistors can only dissipate their rated power at 25°C; at higher case temperatures, maximum power dissipation decreases linearly.
  • Common confusion: The ellipse vs. line distinction—purely resistive loads (0° phase) create straight load lines, purely reactive loads (90° phase) create circles, and real loudspeakers create ellipses that change shape with frequency.
  • Thermal management: Heat sinks transfer heat away from power transistors using increased surface area (fins), with thermal resistance analogous to electrical resistance.

🔊 Complex impedance and load lines

🔊 How loudspeaker impedance affects the load line

  • With a purely resistive load, the AC load line is a straight line passing through the Q point.
  • Real loudspeakers have complex impedance (resistance + reactance), causing the load line to become an ellipse around the Q point.
  • As signal amplitude increases:
    • Resistive case: traces along the straight green line
    • Complex impedance case: traces progressively larger ellipses in red

📐 Phase angle and ellipse shape

The impedance phase angle determines the ellipse aspect ratio:

Phase angleLoad line shapeDescription
Straight linePurely resistive (collapsed ellipse)
Between 0° and 90°EllipseMixed resistance and reactance
90°CirclePurely reactive (fully open ellipse)

⚠️ Frequency-dependent behavior

  • Loudspeaker phase angle changes with frequency.
  • As frequency sweeps from low to high, the load line "wavers back and forth between straight lines and various elliptical shapes."
  • Critical issue: Some elliptical operating regions extend above and to the right of the resistive load line, potentially exceeding the transistor's safe operating area.

Example: At certain frequencies, the ellipse may push the transistor into voltage/current combinations that damage it, even though a purely resistive load at the same power level would be safe.

🔌 Power transistor specifications

🔌 The 2N3055 NPN power transistor

Key maximum ratings:

  • Power dissipation: 115 W at 25°C case temperature
  • Collector current: 15 A maximum
  • Collector-emitter voltage: 60 V maximum
  • Important: The device cannot handle maximum current and voltage simultaneously.

🏗️ TO-3 case design

  • Uses all-metal TO-3 case (not plastic TO-92 used for small signals).
  • Only two leads shown: emitter and base.
  • The entire metal body serves as the collector.
  • Reason: Maximum contact area with heat sink for efficient heat dissipation.

📉 Performance characteristics

Power transistors differ from small-signal devices:

  • Beta (β): Considerably lower than small-signal transistors; can fall below 20 at very high currents.
  • Saturation current (I_C(sat)): Tends to be larger; can be upwards of half a volt.

🛡️ Safe operating area (SOA)

Safe operating area: the zone of V_CE and I_C combinations that the transistor can handle without damage.

  • The combination of collector-emitter voltage and collector current must fall within the lower-left zone of the SOA plot.
  • Key distinction: The safe zone extends further for short pulses than for continuous operation.
  • Don't confuse: A transistor may handle higher power briefly (pulsed) than it can sustain continuously.

🌡️ Power derating and thermal limits

🌡️ Temperature-dependent power rating

Power derating: the reduction in maximum power dissipation capability as case temperature rises above 25°C.

The 115 W rating only applies at 25°C or lower. At higher temperatures, capability decreases.

Example from the excerpt: At 100°C, the 2N3055 can only dissipate about 65 watts (not 115 W).

🧮 Derating calculation

Formula in words:

  • Power dissipation at new temperature = Power at 25°C minus [Derating factor times (Case temperature minus 25°C)]

Where:

  • P_D = power dissipation at the new case temperature
  • P_25 = power dissipation at 25°C (115 W for 2N3055)
  • D = derating factor in watts per degree Celsius (0.657 W/°C for 2N3055)
  • T_case = new case temperature

Example calculation from excerpt: At 75°C, power dissipation = 115 W minus 0.657 W/°C times (75°C minus 25°C) = 82.1 W.

🧊 Heat sink design and usage

🧊 Purpose and construction

Heat sink: a metal device attached to a power transistor to efficiently move heat away from the transistor.

  • Typically made of aluminum.
  • Feature an array of fins to increase surface area.
  • Designed to mount specific case styles (TO-3 "can", TO-220, TO-202 power tabs).

🔧 Mounting requirements

  • Special mounting hardware and insulation spacers maintain electrical isolation between transistor and heat sink.
  • Prevents the heat sink from becoming electrically live.
  • Usually uses: mica sheet, plastic washers and bushings for mounting screws (or nylon screws for small heat sinks).

✅ Best practices for heat sink installation

The excerpt lists these rules:

  1. Always use heat sink grease or thermally conductive pad between heat sink and device to increase thermal transfer; however, excessive quantities decrease performance.
  2. Mount fins vertically for optimum natural convective cooling.
  3. Do not overcrowd or obstruct devices using heat sinks.
  4. Do not block air flow around heat sinks—especially directly above and below items relying on natural convection.
  5. Consider forced convection (a small fan) if thermal demands are particularly high.

🌡️ Thermal resistance model

Thermal resistance (θ): denotes how easy it is to transfer heat energy from one mechanical part to another; units are Celsius degrees per watt.

The thermal circuit equivalent model:

  • Temperature is analogous to voltage.
  • Thermal power dissipation is analogous to current.
  • Thermal resistance is analogous to electrical resistance.

Formula in words: Power dissipated = Temperature differential divided by total thermal resistance.

This is described as "a thermal version of Ohm's law."

56

Power Transistor Data Sheet Interpretation

8.5 Power Transistor Data Sheet Interpretation

🧭 Overview

🧠 One-sentence thesis

Understanding power derating, thermal resistance models, and heat sink selection enables engineers to safely operate power transistors at elevated temperatures without exceeding junction temperature limits.

📌 Key points (3–5)

  • Power derating: as case temperature rises above 25°C, maximum allowable power dissipation decreases linearly according to a derating factor.
  • Heat sinks transfer heat: metal devices with fins increase surface area to efficiently move heat from the transistor to ambient air, preventing thermal failure.
  • Thermal resistance model: temperature differences and thermal resistances (θ) form an electrical-circuit analogy where power dissipation acts like current and temperature acts like voltage.
  • Common confusion: the transistor runs hottest when there is no signal in class A operation, not during signal amplification, because power shifts to the load during signal operation.
  • Heat sink selection: given junction temperature limit, ambient temperature, power dissipation, and device thermal resistances, calculate the maximum acceptable heat sink thermal resistance (θ_sa).

🌡️ Power derating fundamentals

🌡️ What power derating means

Power derating: the reduction in maximum allowable power dissipation as the transistor case temperature increases above 25°C.

  • At 25°C, the transistor has a rated maximum power dissipation (P_25).
  • As case temperature (T_case) rises, the device's ability to dissipate heat is compromised.
  • The relationship is linear: power dissipation decreases by a fixed derating factor (D, in watts per degree Celsius) for each degree above 25°C.

📐 Derating calculation

The excerpt provides the formula in words:

  • P_D = P_25 − D × (T_case − 25°C)
  • P_D is the new allowable power dissipation at the elevated case temperature.
  • D is the derating factor found on the data sheet (e.g., 0.657 W/C° for the 2N3055).

Example: For a 2N3055 at 75°C:

  • P_25 = 115 W
  • D = 0.657 W/C°
  • P_D = 115 W − 0.657 W/C° × (75°C − 25°C) = 82.1 W
  • The graph method gives "a little over 80 watts," confirming the calculation.

🔥 Why derating matters

  • The issue with power transistors is always heat.
  • Internal power dissipation heats the transistor; if junction temperature exceeds the maximum (T_j(max)), the device fails.
  • Derating ensures safe operation by limiting power dissipation as temperature rises.

🧊 Heat sinks and mounting

🧊 What heat sinks do

Heat sink: a metal device (typically aluminum with fins) attached to the power transistor to efficiently move heat away from the device by increasing surface area.

  • Heat sinks are designed for specific case styles: TO-3 "can," TO-220, TO-202 power tabs.
  • They transfer heat from the transistor to the surrounding air more efficiently than the transistor alone.

🔧 Mounting and insulation

  • Special hardware maintains electrical isolation between transistor and heat sink (the heat sink must not be electrically live).
  • Typical insulation: mica sheet, plastic washers and bushings for mounting screws, or nylon screws for small heat sinks.
  • Always use heat sink grease or thermally conductive pad between the heat sink and device to increase thermal transfer.
  • Don't confuse: excessive grease decreases performance; use only enough to fill air gaps.

📋 Best practices for heat sink use

RuleReason
Mount fins verticallyOptimum natural convective cooling
Don't overcrowd devicesObstructions block heat dissipation
Don't block airflowEspecially above/below items relying on natural convection
Use forced convection if neededHigh thermal demands may require a small fan directed at the heat sink

⚡ Thermal resistance model

⚡ Thermal circuit analogy

Thermal resistance (θ): a measure of how easy it is to transfer heat energy from one mechanical part to another, with units of Celsius degrees per watt (C°/W).

  • Temperature is analogous to voltage.
  • Thermal power dissipation is analogous to current.
  • The model creates a "thermal circuit equivalent" to analyze heat flow.

🔗 Thermal resistance chain

The excerpt describes four temperature points and three thermal resistances in series:

SymbolMeaningNotes
T_jJunction temperatureCreated by transistor current × voltage; has a maximum limit
T_cCase temperatureJunction heats the case
T_sHeat sink temperatureCase heats the heat sink
T_aAmbient air temperatureHeat sink transfers energy to surrounding air
θ_jcJunction-to-case thermal resistanceSet by device manufacturer; no user control
θ_csCase-to-sink thermal resistanceFunction of case style and insulation material; some user control
θ_saSink-to-ambient thermal resistanceHeat sink specification; user has great control by choosing the heat sink

📊 Thermal Ohm's law

The excerpt gives the formula in words:

  • P_D = ΔT / θ_total
  • P_D is power dissipated by the semiconductor device (watts).
  • ΔT is the temperature differential.
  • θ_total is the sum of thermal resistances.

More useful form:

  • P_D = (T_j − T_a) / (θ_jc + θ_cs + θ_sa)
  • This is a thermal version of Ohm's law.

🧮 How the model works

  • Ground represents absolute zero temperature.
  • A voltage source of T_a (ambient temperature) is connected to ground and the heat sink.
  • The three thermal resistances are in series.
  • A current source (set by present power dissipation) drives the circuit.
  • Key insight: if power dissipation is high, the "voltage drops" (temperature rises) across thermal resistances are high, creating high junction temperature.
  • Because T_j has a maximum limit, higher power dissipation requires lower thermal resistances.

🛠️ Heat sink selection process

🛠️ Known and unknown quantities

Normally known:

  • Power dissipation (P_D)
  • Junction temperature limit (T_j(max)) — from device data sheet
  • Ambient temperature (T_a) — measured experimentally (tends to be higher than room temperature due to localized warming)
  • θ_jc — from device data sheet
  • θ_cs — from standard graphs based on case style and insulation material

Unknown (to be determined):

  • θ_sa — the required heat sink thermal resistance

📈 Using θ_cs graphs

  • The excerpt references Figure 8.23 for TO-3 and TO-220 cases.
  • TO-3 cases have generally lower θ_cs values than TO-220, making them suitable for higher power devices.
  • The TO-3 case also makes it easier for manufacturers to reduce θ_jc.
  • Different curves apply depending on whether thermal grease is used and the type of insulator (e.g., 0.002 mica).

🧪 Example calculation (15 W dissipation)

Given:

  • T_j(max) = 175°C
  • TO-3 case style
  • θ_jc = 1.5 C°/W
  • P_D = 15 W
  • T_a = 40°C
  • Heat sink grease and 0.002 mica insulator

Steps:

  1. Find θ_cs from graph (Curve 3, "With Thermal Grease"): approximately 0.35 C°/W
  2. Rearrange the formula: θ_sa = (T_j − T_a) / P_D − θ_jc − θ_cs
  3. θ_sa = (175°C − 40°C) / 15 W − 1.5 C°/W − 0.35 C°/W = 7.15 C°/W
  4. This is the maximum acceptable value for the heat sink's thermal resistance.
  5. The heat sink pictured will likely be sufficient without forced air cooling.

Note: Using heat sink grease gives an extra 0.8 C°/W or so of margin.

🧪 High-power example (40 W dissipation)

Same conditions but P_D = 40 W:

  • θ_sa = (175°C − 40°C) / 40 W − 1.5 C°/W − 0.35 C°/W = 1.53 C°/W
  • This requires a much lower thermal resistance.
  • Options:
    • Add forced air cooling of at least 700 feet/minute to the same heat sink.
    • Find a more thermally efficient (probably much larger) heat sink for natural convection alone.

⚠️ Design trade-offs

  • Lower θ_sa values require larger heat sinks or forced air cooling.
  • Natural convection is simpler but limits power dissipation.
  • Forced convection (fans) increases complexity and cost but allows higher power in smaller packages.

🔄 Class A operation thermal behavior

🔄 Counterintuitive thermal characteristic

Don't confuse: In class A amplifiers, the transistor runs hottest when there is no signal, not during signal amplification.

Why this happens:

  • Class A operation means collector current flows for 360° of the cycle.
  • The amplifier draws full current from the power supply regardless of whether a signal is present.
  • With no signal, all power is dissipated within the transistor.
  • With an applied signal, some power formerly dissipated within the transistor is shifted to the load.
  • Therefore, signal application actually reduces transistor dissipation.

🔄 Efficiency implications

  • Maximum theoretical efficiency of class A is only 25%.
  • Low efficiency means most power becomes heat.
  • This makes thermal management critical for class A power amplifiers.
57

Heat Sinks

8.6 Heat Sinks

🧭 Overview

🧠 One-sentence thesis

Heat sinks efficiently transfer heat from transistors to the surrounding air, with their effectiveness measured by thermal resistance (θ), and forced air cooling is added when natural convection cannot handle high power dissipation.

📌 Key points (3–5)

  • Purpose of heat sinks: move heat from the transistor's internal structure to the surrounding air efficiently.
  • Thermal resistance (θ): measures the thermal effectiveness of a heat sink; lower θ means better heat transfer.
  • Power dissipation determines cooling needs: higher power dissipation requires either lower thermal resistance or forced air cooling.
  • Common confusion: the same heat sink may work with natural convection at low power but require forced air (e.g., 700 feet/minute) at higher power levels.
  • Design trade-offs: you can use forced air with a smaller heat sink or choose a larger, more thermally efficient heat sink for natural convection alone.

🌡️ Thermal resistance fundamentals

🌡️ What thermal resistance measures

Thermal resistance (θ): the measure of a heat sink's effectiveness at transferring heat from the transistor to the surrounding air.

  • Expressed in units of degrees Celsius per watt (C°/W).
  • Lower θ values indicate better heat transfer capability.
  • The excerpt emphasizes that θ quantifies thermal effectiveness, not just physical size.

🔽 Why lower is better

  • A lower thermal resistance means the heat sink can transfer more heat for the same temperature difference.
  • Example: A heat sink with θ = 1.53 C°/W transfers heat less efficiently than one with θ = 0.35 C°/W.
  • Don't confuse: thermal resistance is not the same as electrical resistance, though both use similar mathematical relationships.

🔧 Calculating heat sink requirements

🔧 The thermal resistance chain

The excerpt shows a calculation using multiple thermal resistances in series:

  • θ_jc: junction-to-case thermal resistance (internal to the transistor)
  • θ_cs: case-to-sink thermal resistance (interface between transistor and heat sink)
  • θ_sa: sink-to-ambient thermal resistance (heat sink to surrounding air)

The total thermal path is: P_D = (T_j - T_a) / (θ_jc + θ_cs + θ_sa)

Where:

  • P_D = power dissipation
  • T_j = junction temperature
  • T_a = ambient temperature

📐 Worked example comparison

The excerpt provides two scenarios with the same transistor but different power levels:

Power dissipationRequired θ_saCooling method needed
Lower power (not specified)Higher θ_sa acceptableNatural convection (under 200 ft/min)
40 W1.53 C°/WForced air ≥700 ft/min OR larger heat sink
  • With T_j = 175°C, T_a = 40°C, θ_jc = 1.5 C°/W, θ_cs = 0.35 C°/W
  • Solving for θ_sa: θ_sa = (175 - 40)/40 - 1.5 - 0.35 = 1.53 C°/W

🔄 Design alternatives

When the calculated θ_sa is too low for natural convection:

  1. Add forced air cooling: increase airflow (e.g., to 700 feet/minute) to effectively lower θ_sa
  2. Use a larger heat sink: find a more thermally efficient heat sink with lower natural convection θ_sa
  3. Trade-off: forced air adds noise and complexity; larger heat sinks add size and cost

💨 Natural convection vs forced air cooling

💨 When forced air becomes necessary

  • Natural convection: air flow under 200 feet/minute (passive cooling)
  • Forced air: active cooling with fans, measured in feet/minute
  • The excerpt shows that at 40 W dissipation, at least 700 feet/minute is needed if using the original heat sink

🔥 High power applications

  • For high power applications where significant heat is generated, heat sinks are augmented with forced air cooling
  • This is a practical necessity, not just an optimization
  • Example: The same heat sink that works passively at low power may be completely inadequate at 40 W without forced air

⚠️ Don't confuse

  • The heat sink itself doesn't change, but its effective thermal resistance (θ_sa) improves with forced air
  • Natural convection is simpler and quieter but only works for lower power dissipation levels
58

BJT Class B Power Amplifiers

Chapter 9: BJT Class B Power Amplifiers

🧭 Overview

🧠 One-sentence thesis

Class B amplifiers achieve much greater power efficiency than class A by using two transistors that idle near cutoff and draw current only as needed, though this comes at the cost of added complexity and unique distortion challenges.

📌 Key points (3–5)

  • What class B operation means: AC collector current flows for only 180° out of the cycle, requiring two devices to amplify the entire signal.
  • Why class B is more efficient: the Q point sits at cutoff instead of midway on the AC load line, so transistors idle at zero (or very low) current rather than half-maximum, drawing power only when needed.
  • Trade-off: improved efficiency comes with added complexity—two transistors, trickier biasing, and a unique form of distortion not present in class A.
  • Common confusion: class A idles at half-maximum current even with no signal; class B idles near zero and runs cool at low output power.
  • Key auxiliary circuits: current mirror, Sziklai pair, V_BE multiplier, and overload protection improve performance and address biasing challenges.

⚡ Why class B exists: the efficiency problem

🔋 Class A inefficiency

  • Class A amplifiers transform at best only 25% of DC input power into useful load power.
  • The optimal Q point for class A is midway along the AC load line, allowing equal swing toward saturation and cutoff.
  • The waste: even with no signal present, the transistor still pulls half of the maximum current continuously.
  • This continuous current draw means constant power dissipation and heat, regardless of output level.

💡 Class B solution

Class B operation: AC collector current flows for 180° out of the cycle.

  • The Q point is pushed down to sit right at cutoff on the AC load line.
  • This means I_CQ (quiescent collector current) is at or near zero.
  • Key advantage: transistors draw current only as needed, not continuously.
  • Result: class B designs run relatively cool at idle and at low output power, and a much larger percentage of applied DC power can be turned into useful AC output to the load.

🔧 Class B configuration basics

🔀 Two-device requirement

  • Because each transistor conducts for only 180° (half the cycle), two devices are needed to amplify the entire signal.
  • One device handles the positive half-wave, the other handles the negative half-wave.
  • The two waveform halves must be "stitched together," which can be a problem area.

⚙️ Added complexity

The excerpt lists several complications compared to class A:

  • Biasing: trickier than the straightforward class A biasing circuits.
  • Distortion: class B amplifiers suffer from a unique form of distortion that class A amplifiers do not (the chapter will discuss "notch distortion" and mitigation methods).
  • Circuit design: requires modifications and auxiliary sub-circuits to improve performance.

🛠️ Auxiliary circuits and techniques

🔄 Current mirror

  • Mentioned as one of the auxiliary device configurations used to improve performance.
  • The chapter will explain its operation.

🔗 Sziklai pair

  • Another auxiliary configuration to be covered.
  • Used in class B designs to address performance challenges.

🔢 V_BE multiplier

  • A sub-circuit that helps with biasing.
  • The chapter will explain its operation and use.

🛡️ Overload protection

  • Methods to protect the output devices from overload.
  • Important because class B stages can be vulnerable to damage under certain conditions.

🔊 Output stage topologies

The chapter will outline:

  • Fully complementary output stages utilizing direct coupled driver stages.
  • Quasi complementary output stages utilizing direct coupled driver stages.

📐 Design focus and scope

🎯 Common collector configuration

  • As with class A amplifiers, the common collector (voltage follower) configuration is the most widely used for class B circuits.
  • The chapter focuses on followers, not on voltage amplifiers.

📊 Performance metrics

The chapter will cover how to determine:

  • AC load lines for class B amplifier stages.
  • Compliance (maximum output voltage swing).
  • Maximum load power.
  • Efficiency.
  • Required device ratings.

⚖️ Class A vs class B comparison

AspectClass AClass B
EfficiencyAt best 25%Much greater (specific value to be covered)
Q point locationMidway on AC load lineAt cutoff on AC load line
Idle currentHalf of maximumZero or very small nominal value
Power drawContinuous, even with no signalOnly as needed
Heat at idle/low powerRuns hotRuns cool
Number of transistorsOne per stageTwo per stage (for full cycle)
Biasing complexityStraightforwardTrickier
Unique distortionNone mentionedNotch distortion

⚠️ Don't confuse

  • Class A idle behavior: "no signal" does not mean "no current"—the transistor still draws half-maximum current continuously.
  • Class B idle behavior: "no signal" means near-zero current draw, which is why it runs cool and is more efficient.
  • The 180° conduction per device in class B is not a limitation but the design principle that enables efficiency; the two devices together cover the full 360° cycle.
59

BJT Class B Power Amplifiers

9.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

Class B amplifiers achieve far greater power efficiency than class A by biasing two complementary transistors at cutoff so each conducts for only 180° of the waveform cycle, though this requires careful design to avoid notch distortion and thermal runaway.

📌 Key points (3–5)

  • What class B operation means: each transistor conducts for 180° (half the cycle), requiring two devices (NPN and PNP) to amplify the entire signal; idle current is near zero, unlike class A.
  • Efficiency advantage: theoretical maximum efficiency is 78.5% (versus 25% for class A); worst-case transistor dissipation is only one-fifth of maximum load power.
  • Notch distortion problem: pure class B creates a "dead zone" near zero-crossing because transistors don't turn on until ±0.7 V; solved by class AB biasing (slightly "on") using diodes or V_BE multipliers.
  • Common confusion—thermal runaway vs. current hogging: in parallel output transistors, one device can get hotter, conduct more current, get even hotter, and destroy itself; prevented by small emitter resistors (local negative feedback).
  • Why it matters: class B is the mainstay of linear high-power amplifiers (especially audio) because it runs cool at idle, draws current only as needed, and delivers much more power per watt of DC supply.

🔋 Core class B operation

🔋 What class B means

Class B operation: AC collector current flows for 180° out of the cycle.

  • Each transistor is biased right at cutoff on the AC load line, so I_CQ = 0 A and virtually no power is drawn at idle.
  • When the input swings positive, the NPN transistor turns on and sources current into the load; when the input swings negative, the NPN clips and the PNP transistor sinks current from the load.
  • This is called a push-pull configuration: one transistor pushes, the other pulls.
  • Don't confuse: class A idles at half maximum current (Q point midway on the load line); class B idles at zero current (Q point at cutoff).

🔍 AC vs. DC load lines

  • The AC load line for class B runs from cutoff (V_CEQ = half the total supply) up toward saturation as the transistor conducts.
  • The DC load line goes straight up (vertical) because there is no collector resistor to limit DC current—saturation current is theoretically infinite.
  • Danger: without current limiting, a shorted load or bias error can destroy the transistors instantly.

Example: If V_CC = +15 V and V_EE = −15 V, then V_CEQ = 15 V for each transistor; the AC load line starts at 15 V and swings toward 0 V as collector current increases.

⚠️ Notch distortion and class AB

⚠️ The dead-zone problem

  • Pure class B creates notch distortion (also called crossover distortion): the output waveform has a flat spot or "notch" near zero-crossing.
  • Why it happens: the input signal must exceed ±0.7 V to turn on the transistors; anything between −0.7 V and +0.7 V is a dead zone that the amplifier does not respond to.
  • This distortion is worst for small signals and gets relatively better as signal level increases (opposite of most nonlinearities).

🛠️ Class AB solution

Class AB operation: biasing the transistors slightly "on" (small idle current) so only a very small input signal is needed to turn them on fully.

  • The conduction angle is slightly more than 180° (hence "AB").
  • Biasing methods:
    • Simple voltage divider with resistors (R₃ and R₄ set to drop ~0.7 V each)—but resistors don't match the exponential I-V curve of PN junctions well, so stability is poor.
    • Diode biasing: use signal diodes in place of resistors; diodes mimic the base-emitter characteristic much better, so bias is more stable and devices track temperature changes.
    • V_BE multiplier (covered later): adjustable bias using a transistor and resistor divider; allows precise tuning of idle current.

Don't confuse: the diodes are already forward-biased by the DC supply, so the AC signal sees the diode's dynamic resistance, not an open circuit.

📊 Class B performance metrics

📊 Compliance and load power

  • Peak compliance = V_CEQ = 0.5 × (total DC supply).
  • For a bipolar supply (e.g., +V_CC and −V_EE), V_CEQ equals one side of the supply.
  • Maximum load power: P_load(max) = (Compliance_RMS)² / R_L = (0.707 × V_CEQ)² / R_L.

Example: V_CC = +15 V, V_EE = −15 V, R_L = 8 Ω → V_CEQ = 15 V peak = 10.6 V RMS → P_load(max) = (10.6)² / 8 ≈ 14 W.

🔥 Transistor power dissipation

  • Worst-case dissipation occurs at about 40% of maximum load power (not at maximum load power).
  • At that point, each transistor dissipates approximately P_load(max) / 5 (20% of max load power).
  • At maximum load power, transistor dissipation is only about 13.7% of P_load(max).
  • Why: when i_C is maximum, v_CE is near zero, so power is low; worst case is when i_C peaks at ~64% of max and v_CE is still ~36% of max, so their product (power) is highest.
ConditionLoad powerTransistor dissipation (each)
Worst case~40% of P_load(max)~20% of P_load(max)
Maximum load power100% of P_load(max)~14% of P_load(max)

⚡ Efficiency

  • Maximum theoretical efficiency = π/4 ≈ 78.5% (at maximum load power).
  • This is over three times the 25% maximum efficiency of class A.
  • Class B draws current dynamically (only as needed), so it runs cool at idle and low output levels.

🔌 Device ratings

  • BV_CEO must exceed the total DC supply (when one transistor is off, it can see the entire supply voltage).
  • Maximum collector current i_C(sat) = V_CEQ / r_L (the compliance divided by the load resistance).

🏗️ Practical circuit techniques

🏗️ Current mirror for biasing

Current mirror: a circuit that uses a diode and transistor to "mirror" a programmed current; the emitter current tracks the diode current because V_BE = V_D.

  • A resistor R sets the diode current I_D ≈ (V_supply − 0.7) / R.
  • The transistor's base-emitter is in parallel with the diode, so V_BE = V_D.
  • If the transistor's I-V curve matches the diode's, then I_E ≈ I_D.
  • Why it's better than resistors: diodes and transistors track temperature changes together; much more stable than using a resistor to bias a PN junction.

🔗 Direct coupled driver

  • A direct coupled driver is a class A common-emitter stage whose "collector resistor" is replaced by the class B output stage.
  • Eliminates three components: collector resistor, interstage coupling capacitor, and the lower base biasing resistor of the output stage.
  • Raises the effective load resistance for the driver, increasing voltage gain.
  • Biasing rule: the DC voltage across the biasing resistor (R₃ in the excerpt) must equal approximately V_CC − 0.7 V to ensure the output sits at 0 VDC.

Example: If V_CC = 20 V and R₃ = 560 Ω, then I_CQ1 = (20 − 0.7) / 560 ≈ 34.5 mA; this sets the voltage across the emitter resistor R₄, which determines the divider ratio for R₁ and R₂.

🔧 High current gain configurations

Darlington pair: two transistors in cascade; the first transistor's collector current drives the second's base, so total β = β₁ × β₂.

  • Requires compensation for four base-emitter drops (two per side).

Sziklai pair (composite pair): an NPN-PNP combination that acts like a high-β PNP.

  • The input transistor (PNP) drives its collector current into the base of the output transistor (NPN).
  • Only one V_BE drop to compensate (versus two for Darlington).
  • Quasi-complementary output: uses a Darlington NPN on one side and a Sziklai pair on the other; allows both power devices to be identical NPN models.

Don't confuse: Darlington has two drops per side; Sziklai has one drop per side.

🛡️ Active current limiting

  • Problem: no DC current limit in basic class B; a shorted load can destroy transistors.
  • Solution: insert a small resistor R_E in the emitter path and place a protection transistor Q₂ across it.
  • Under normal operation, voltage across R_E is less than 0.5 V, so Q₂ is off.
  • If load current exceeds the safe limit, voltage across R_E reaches 0.7 V, turning on Q₂; Q₂ diverts base current away from the output transistor, limiting load current to ~0.7 / R_E.
  • Once the fault is removed, Q₂ disengages and normal operation resumes (unlike a fuse, which must be replaced).

🔁 Current sharing and thermal runaway

Thermal runaway (current hogging): when multiple transistors are in parallel, one device gets hotter, conducts more current (because r'_e decreases with temperature), gets even hotter, and eventually destroys itself.

  • Solution: add small resistors to the emitters of each paralleled transistor (local negative feedback).
  • If one transistor tries to grab more current, the voltage drop across its emitter resistor increases, reducing its V_BE and compensating for the initial current increase.
  • Don't confuse: this is not the same as the small R_E used for current limiting; these are even smaller resistors used to balance current among parallel devices.

🎛️ V_BE multiplier

V_BE multiplier: a transistor and resistor divider that generates a voltage equal to a multiple of V_BE; allows adjustable bias for class AB operation.

  • If R₁ is three times R₂, the total voltage from point A to point B is four times V_BE.
  • By replacing one resistor with a potentiometer, the bias voltage becomes adjustable, allowing precise tuning of idle current to minimize notch distortion.
  • A capacitor C_B shunts the multiplier to ensure it acts as a short for AC signals.

🌉 Bridged output

  • Bridged configuration: two identical amplifiers drive a floating load from both sides; one amplifier gets the normal signal, the other gets an inverted copy.
  • As the left output goes positive, the right output goes negative by the same amount, so the load sees twice the voltage.
  • Power varies as voltage squared, so doubling voltage quadruples load power.
  • Downsides: doubles the output circuitry; the load is not grounded (both terminals must be wired to the amplifier, not one to chassis ground).

Example: A car amplifier with V_CC = +12 V normally delivers 2.25 W into 8 Ω; bridged, it delivers 9 W.

📐 Design and analysis summary

📐 Key formulas

QuantityFormulaNotes
V_CEQ0.5 × (total DC supply)Each transistor sees half the supply
Compliance (peak)V_CEQSame as V_CEQ for class B
i_C(sat)V_CEQ / r_LMaximum AC current
P_load(max)(Compliance_RMS)² / r_LMaximum load power
P_D(worst)P_load(max) / 5Worst-case transistor dissipation (each)
Efficiency (max)π / 4 ≈ 78.5%At maximum load power
BV_CEOTotal DC supplyWhen one transistor is off

📐 Design checklist

  1. Set V_CEQ = half the total supply.
  2. Calculate compliance and P_load(max) from V_CEQ and r_L.
  3. Size transistors for BV_CEO ≥ total supply and I_C(max) ≥ V_CEQ / r_L.
  4. Size heat sinks for P_D(worst) = P_load(max) / 5 (per transistor).
  5. Choose biasing method: diodes for simplicity, V_BE multiplier for adjustability.
  6. If using direct coupled driver, ensure DC voltage across biasing resistor ≈ V_CC − 0.7 V.
  7. Add current limiting if load can be shorted.
  8. Add emitter resistors if using parallel output transistors (current sharing).
  9. Add power supply bypass capacitors (large electrolytics in parallel with smaller poly caps for high-frequency performance).

Don't confuse: The driver stage (class A) may clip before the output stage (class B); perform an AC load line analysis on the driver to verify full swing capability.

60

BJT Class B Power Amplifiers – Introduction

9.1 Introduction

🧭 Overview

🧠 One-sentence thesis

Class B amplifiers achieve much greater power efficiency than class A by positioning the Q point at cutoff so that each transistor handles only half the waveform and draws current only when needed, though this introduces added complexity and unique distortion challenges.

📌 Key points (3–5)

  • Core efficiency advantage: Class B draws current only as needed and idles near zero, unlike class A which continuously draws half-maximum current even at idle.
  • Why two transistors are required: Each device conducts for only 180° of the cycle (one half-wave), so a mirror-image pair is needed to amplify the entire signal.
  • Trade-off for efficiency: Increased circuit complexity, trickier biasing, and a unique form of distortion (notch distortion) that class A does not suffer from.
  • Common confusion: Class A vs class B Q point placement—class A centers the Q point midway on the AC load line; class B places it right at cutoff.
  • Why class A is inefficient: At best, class A converts only 25% of DC input power into useful AC load power because the transistor idles at half-maximum current.

⚡ Why class B improves efficiency

⚡ The class A idle problem

  • In class A operation, the optimal Q point sits midway along the AC load line to allow equal swing toward saturation and cutoff.
  • The inefficiency: Even with no input signal, the transistor still pulls half of the maximum current continuously.
  • The excerpt compares this to a car engine running at 3000 RPM while sitting motionless at a red light—wasteful.

🔋 Class B solution: Q point at cutoff

Class B operation: AC collector current flows for 180° out of the cycle.

  • The Q point is pushed down to sit right at cutoff on the AC load line.
  • At idle: I_CQ = 0 A, so virtually no power is drawn from the supply.
  • Under signal: Current is drawn only as needed, so the amplifier runs cool at idle and low output power.
  • Result: A much larger percentage of applied DC power is converted into useful AC output to the load, and power dissipation requirements for transistors are lowered.

🔀 How class B handles the full waveform

🔀 Single transistor limitation

  • Placing the Q point at cutoff means the transistor immediately clips the negative portion of the waveform (for an NPN device).
  • When the input swings positive, collector current increases and voltage develops across the load.
  • When the input swings negative, the transistor turns off—no collector current, no load voltage, and v_CE stays at cutoff.
  • The action resembles half-wave rectification: only the positive half is reproduced.

🪞 Mirror-image pair requirement

  • A PNP version of the circuit does the exact opposite: reproduces the negative portion and clips the positive portion.
  • Two devices needed: One NPN and one PNP to cover both halves of the cycle.
  • Challenge: How to "stitch together" the two waveform halves without introducing problems—this is a potential trouble area.

⚙️ Trade-offs and complexity

⚙️ Added circuit complexity

  • Unlike class A, class B requires two transistors for linear operation.
  • Biasing is trickier and requires modifications to the straightforward class A biasing circuits.
  • The excerpt mentions auxiliary device configurations and sub-circuits to improve performance: current mirror, Sziklai pair, V_BE multiplier, and overload protection circuitry.

⚠️ Unique distortion issue

  • Class B amplifiers suffer from a unique form of distortion (notch distortion) that class A amplifiers do not experience.
  • This distortion arises from how the two waveform halves are joined together.

🎯 Common configuration

  • As with class A, the common collector (voltage follower) configuration is the most widely used for class B circuits.
  • The chapter focuses on followers rather than voltage amplifiers.

📊 Class A vs class B comparison

AspectClass AClass B
Q point locationMidway on AC load lineAt cutoff
Idle currentHalf of maximumZero (or nominal)
Efficiency (max)25%Much higher
Number of transistorsOneTwo (NPN + PNP pair)
Distortion typeStandardNotch distortion
Circuit complexitySimpler biasingMore complex, requires special sub-circuits
Power drawContinuous full powerDraws current only as needed
61

9.2 The Class B Configuration

9.2 The Class B Configuration

🧭 Overview

🧠 One-sentence thesis

Class B amplifiers achieve much higher efficiency than class A by biasing each transistor at cutoff so that two complementary devices handle opposite half-cycles of the waveform, though this introduces unique distortion challenges at the zero-crossing point.

📌 Key points (3–5)

  • What class B means: each transistor conducts for only 180° (half) of the AC cycle, requiring two devices (NPN and PNP) to reproduce the full waveform.
  • Why class B matters: efficiency jumps from 25% (class A) to a theoretical maximum of 78.5% because transistors idle at nearly zero current instead of half-maximum.
  • Crossover distortion problem: a dead zone between ±0.7 V causes notch distortion because neither transistor conducts until the input exceeds the base-emitter threshold.
  • Class AB solution: adding a small bias voltage (via diodes or current mirrors) keeps transistors slightly on, eliminating the dead zone while preserving most of the efficiency gain.
  • Common confusion: maximum transistor heating occurs at only ~40% of maximum load power (not at full power), because that's when the product of collector current and collector-emitter voltage peaks.

⚡ Why class B exists: the efficiency problem

⚡ Class A inefficiency

  • In class A, the Q point sits midway on the AC load line so the signal can swing equally toward saturation and cutoff.
  • The waste: even with no signal present, the transistor draws half of maximum current continuously.
  • Analogy from the excerpt: "How much sense would it make to have an engine with a 6000 RPM maximum run at 3000 RPM when you're sitting motionless at a red light?"

⚡ Class B solution: biasing at cutoff

  • The Q point is pushed down to cutoff on the AC load line, so idle current (I_CQ) is 0 A and virtually no power is drawn at rest.
  • Trade-off: locating the Q point at cutoff means the transistor immediately clips the negative half of the waveform.
  • Consequence: a mirror-image circuit (complementary PNP) is needed to produce the missing half.

🔄 How class B operation works

🔄 The push-pull principle

  • Two transistors (one NPN, one PNP) have their emitters tied together and connected to the load.
  • The NPN collector goes to the positive supply; the PNP collector goes to ground (or negative supply in a bipolar configuration).
  • Sourcing and sinking:
    • When the input swings positive, the NPN turns on and pushes current into the load (sourcing).
    • When the input swings negative, the PNP turns on and pulls current from the load (sinking).
  • With no signal, both transistors are off because their bases are tied together and no external potential is applied, so both base-emitter voltages are zero.

🔄 AC load line behavior

  • As the input signal swings positive, the operating point slides up the load line toward saturation, increasing collector current and creating a load voltage that follows the input.
  • When the input tries to swing negative, there is no place else to go on the load line—the negative portion is clipped.
  • The complementary transistor handles the opposite half in the same manner.

❌ Crossover distortion and the class AB fix

❌ The dead zone problem

Crossover distortion (also called notch distortion): a gross form of distortion that occurs because the input signal will not truly turn on the NPN transistor until it exceeds approximately 0.7 V, or drops below −0.7 V for the PNP side.

  • The region between ±0.7 V is a dead zone that the amplifier does not respond to.
  • The amplifier "rips out" anything between ±0.7 V, creating a flat spot near the zero-crossing points.
  • Particularly nasty: this distortion hits small signals worse than large signals, unlike most nonlinearities that worsen as signal level increases.

❌ Class AB operation

Class AB operation: providing a small idle current so that the transistors are almost on, requiring only a very small input signal to turn on the devices and slightly increasing the conduction angle beyond 180°.

  • Diode bias solution: inserting diodes (typically two in series) between the bases creates a small forward bias (~0.7 V each).
  • The diodes provide a much better match for the base-emitter junction than resistors because they track temperature changes similarly.
  • How the signal passes through: the diodes are already forward-biased by the DC supply, so the AC signal sees the dynamic resistance of the diode, not an open circuit.
  • Don't confuse: the diode's dynamic resistance does fluctuate with signal level, introducing slight distortion, but this is orders of magnitude smaller than the notch distortion it eliminates.

❌ Current mirror for better bias stability

Current mirror: a circuit that uses a diode in series with a resistor to set a reference current, which is then mirrored by a transistor whose base-emitter junction is in parallel with the diode.

  • The voltage across the resistor R equals the supply voltage minus the diode drop (approximately V − 0.7 V), setting up a current I_R.
  • If base current is negligible, this same current flows through the diode as I_D, establishing a specific diode voltage.
  • Because the diode is in parallel with the base-emitter junction, V_BE = V_D, so the emitter current mirrors the diode current.
  • Why it's better: using a diode to match a PN junction is far more stable than using a resistor (linear I-V curve) to match the exponential I-V curve of a transistor.
  • In integrated circuits, device matching is easy; for discrete components, any signal diode provides a much better match than a resistor.

📐 Class B circuit maximums and ratings

📐 Voltage and compliance

  • Because the two transistors split the available supply, V_CEQ always equals half of the total supply.
  • Since devices are biased at cutoff: v_CE(cutoff) = V_CEQ = 0.5 × Total DC Supply (Equation 9.1).
  • Peak compliance equals V_CEQ, which is the same as one side of a bipolar supply (Equation 9.2).
  • Breakdown voltage: the off-state transistor can see the entire power supply when the opposite transistor is fully conducting, so BV_CEO = Total DC Supply (Equation 9.3).

📐 Current and load power

  • Saturation current is dictated by compliance and load: i_C(sat) = Compliance_peak / r_L (Equation 9.4).
  • Maximum load power: P_load(max) = Compliance_RMS² / r_L (Equation 9.5).
  • Critical warning: there is nothing in the collector-emitter line to limit DC current—the DC load line goes straight up with no saturation current value short of infinity.
  • If bias is not careful or the load is accidentally shorted, huge currents can destroy the transistors.

🔥 Transistor power dissipation (the tricky part)

🔥 Why idle dissipation is not the worst case

  • Idle (Q point) power dissipation is very low: P_DQ = I_CQ × V_CEQ, where I_CQ is typically a few percent of i_C(sat).
  • Unlike class A, P_DQ does not represent the worst case for class B.

🔥 Deriving worst-case dissipation

  • During the conduction phase, collector current appears as the positive half of a sine wave, peaking at i_C(sat) (or V_CEQ / r_L).
  • Collector-emitter voltage starts at V_CEQ and swings down to zero as a negative half sine.
  • Introducing a coefficient k (0 ≤ k ≤ 1) representing the percentage of maximum current:
    • i_C = k × (V_CEQ / r_L) × sin(2πft) (Equation 9.6)
    • v_CE = V_CEQ × (1 − k × sin(2πft)) (Equation 9.7)
  • Power dissipation P_D = i_C × v_CE; after integration and averaging over one full cycle:
    • P_D = 2 × P_load(max) × (k/π − k²/4) (Equation 9.8)
  • Taking the derivative and setting it to zero: worst case occurs at k = 2/π ≈ 63.7% of the load line.
  • Substituting back: P_D(worst) ≈ P_load(max) / 5 (Equation 9.9).

🔥 Key insight

  • When the load receives about 40% of its maximum power (since 0.637² ≈ 0.4), the transistors are at their hottest, dissipating approximately 20% of maximum load power (about half the power delivered to the load at that point).
  • Example: if a class B amplifier is rated for 100 W maximum load power, the transistors get hottest at 40 W load power, with each transistor dissipating 20 W.
  • At maximum load power (k = 1), transistor dissipation drops to 13.7% of maximum load power.
  • Why this happens: at maximum load power, when i_C is maximum, v_CE is 0, so instantaneous power is 0; at the worst case, collector current peaks at ~64% but v_CE only drops to ~36%, resulting in a higher average power dissipation.

📊 Efficiency and practical considerations

📊 Maximum theoretical efficiency

  • Efficiency η = P_out / P_in = P_load(max) / P_DC.
  • At P_load(max), the DC supply delivers full voltage 2V_CEQ and peak current i_C(sat) = V_CEQ / r_L.
  • Average current over a half-cycle: i_C(avg) = (1/π) × (V_CEQ / r_L).
  • Supplied power: P_DC = 2V_CEQ × i_C(avg) = (2/π) × (V_CEQ² / r_L) = (4/π) × P_load(max).
  • Therefore: η_max = π/4 ≈ 78.5% (over three times the 25% maximum of class A).

📊 Reactive loads and real-world limits

  • Reactive load issues from class A still apply: loads with complex impedance may be harder to drive than ideal resistive loads.
  • Transistors may need to be over-rated above P_load(max) / 5 to handle reactive loads safely.
  • The AC load line for a class B amplifier with a reactive load appears as an ellipse cut in half horizontally through the Q point.

🔗 Direct coupled driver stage

🔗 Why direct coupling

  • The basic class B circuits offer current gain and power gain but not voltage gain.
  • Coupling capacitors can connect preceding voltage gain stages, but a more effective method is the direct coupled driver.

🔗 How it works

  • A class A common emitter amplifier (e.g., using voltage divider bias) is combined with a class B follower.
  • The follower is positioned where the common emitter stage's collector resistor would normally be.
  • Components eliminated: the collector resistor, the interstage coupling capacitor, and the lower base biasing resistor of the class B output stage.
  • Removing the resistors raises the effective load resistance for the first stage, producing higher voltage gain from the driver.

🔗 Biasing the direct coupled driver

  • Critical rule: the DC voltage across the bias resistor (R₃ in the excerpt's example) must equal approximately V_CC − 0.7 V.
  • If this is not the case, the class B stage will not be symmetrical (V_CEQ2 ≠ V_CEQ3), and the final output will not sit at 0 VDC as required.
  • Knowing R₃ and its voltage determines its current, which flows into the driver transistor as I_CQ1.
  • From I_CQ1, calculate the voltage across the emitter resistor (R₄) and determine the appropriate divider ratio for the base biasing resistors.
  • Practical tip: due to component tolerances, one biasing resistor should be a potentiometer (or a resistor/pot combo) to "tweak" the final output to 0 VDC without being too sensitive.

🔗 System voltage gain

  • The follower has a gain of approximately one, so overall gain depends on the first stage common emitter amplifier.
  • The effective load at the collector of the driver is the bias resistor in parallel with a single Z_in(base) (only one output transistor is on at any given time; the off transistor appears as high impedance).
  • If the driver stage is swamped, the gain is approximately −r_L / r_E.
  • Important limitation: the class B stage is a follower with voltage gain of one, so if the driver stage can't produce the full swing, the output stage can't either. The driver stage may clip before the output stage.

🔗 Driver polarity

  • The direct coupled driver does not have to be NPN; a PNP can be used instead, simply shifted to the upper section rather than the lower section.

🛡️ Protection and refinements

🛡️ The current limiting problem

  • There is nothing in the basic class B amplifier to limit current, so an accidental load short causes transistor current to spike to destructive levels.
  • Fuse limitations: fuses are not fast-response devices; rating them low enough to catch fast transients may cause them to blow on moderately loud sustained low-frequency content. Replaceable fuses risk consumer misuse (wrong value, wire, foil, etc.).

🛡️ Active current limiter concept

  • A small resistor (R_E) is inserted into the emitter current path, with resistance small enough that the voltage across it is normally less than 0.5 V.
  • A transistor (Q₂) is placed across this resistor; under normal operation, Q₂ is off and does not affect the circuit.
  • If load current exceeds the safe limit, the voltage drop across R_E reaches 0.7 V, turning on Q₂.
  • Q₂ then conducts current away from the base of the main output transistor, limiting the output current.

🛡️ Simulation insights

  • Transient analysis confirms voltage gain is approximately unity (input and output waves nearly coincident except for clipping).
  • With diode bias, clipping occurs prematurely (around 8.5 V for a ±10 V supply) because the biasing diodes become reverse-biased when the input gets close to the power supply, preventing the signal from reaching the base.
  • Without biasing diodes, the output suffers obvious notch distortion, but compliance improves (clipping occurs just under the power rails).
  • Trade-off: a little bit of compliance is lost due to the diodes, but this is much better than getting notch distortion.
62

Extensions and Refinements of Class B Amplifiers

9.3 Extensions and Refinements

🧭 Overview

🧠 One-sentence thesis

Class B amplifiers can be enhanced with protection circuits, high-gain configurations, current-sharing schemes, and flexible biasing to improve reliability, power handling, and performance while managing thermal and electrical challenges.

📌 Key points (3–5)

  • Active current limiting protects output transistors from damage during overload or short-circuit conditions without permanent fuses.
  • High current gain configurations (Darlington and Sziklai pairs) address insufficient β in power transistors and reduce the number of bias compensation junctions needed.
  • Current sharing with thermal feedback prevents thermal runaway when multiple output transistors operate in parallel for high-power applications.
  • Common confusion: thermal runaway vs. normal heating—positive thermal feedback causes one paralleled device to "hog" current and self-destruct; local negative feedback (emitter resistors) breaks this loop.
  • Bridging doubles load voltage by driving both terminals of a floating load with inverted signals, quadrupling power output without increasing supply voltage.

🛡️ Protection and limiting

🛡️ Active current limiter operation

  • A small resistor (R_E) is inserted in the emitter path of the output transistor.
  • Under normal operation, voltage across R_E is below 0.7 V, so the protection transistor (Q2) remains off.
  • When load current exceeds safe limits, the voltage across R_E reaches 0.7 V, turning on Q2.
  • Q2 then diverts base current away from the main output transistor, capping load current at approximately 0.7 V / R_E.
  • Key advantage over fuses: once the fault is removed, the circuit automatically resumes normal operation without replacement.

🔧 Why fuses are problematic

  • Voice and music waveforms are highly dynamic, making proper fuse rating difficult.
  • Low-rated fuses may blow on loud sustained low-frequency content.
  • High-rated fuses may not protect against fast transients.
  • User-replaceable fuses risk incorrect values or improvised substitutes (wire, foil, bolts), leading to transistor damage.

🔋 High current gain configurations

🔋 Darlington pairs

  • Two transistors are cascaded: a drive transistor feeds its collector current into the base of the main output device.
  • Total current gain equals β₁ × β₂, providing much higher gain than a single transistor.
  • Trade-off: four base-emitter junctions (two per side) require four compensating diodes for proper bias.
  • Both transistors need high BV_CEO ratings, but the drive transistor handles less current and dissipates less power.

🔄 Sziklai pair (composite pair)

A dual-transistor configuration where the input transistor drives its collector current into the base of an opposite-type output transistor, achieving Darlington-like gain with only one V_BE drop.

  • Also called a composite pair; named after engineer George Sziklai.
  • For a composite PNP: the input is PNP, but the main power device is an NPN.
  • Key advantage: only one base-emitter junction to compensate (versus two for Darlington).
  • Quasi-complementary output: uses a Darlington NPN and a Sziklai PNP, allowing identical power device models (both NPN).
ConfigurationJunctions to compensateMain power device typesCurrent gain
Darlington2 per side (4 total)NPN + NPN, PNP + PNPβ₁ × β₂
Sziklai1 per side (2 total)NPN for both sidesβ₁ × β₂
Quasi-complementary3 total (2 + 1)Both NPNβ₁ × β₂

🔥 Current sharing and thermal management

🔥 Thermal runaway mechanism

  • Bipolar transistors have a positive temperature coefficient of transconductance: r'_e decreases as temperature rises.
  • A hotter device conducts more current → dissipates more power → gets hotter → conducts even more current.
  • In parallel configurations: devices are never perfectly matched, so one device tends to get hotter and "hog" all the current until it destroys itself.
  • This is a positive thermal feedback loop leading to transistor self-destruction.

🛠️ Local negative feedback solution

  • Small resistors are added to the emitters of paralleled output transistors.
  • All paralleled devices share a common base voltage from the drive transistor.
  • If one transistor starts grabbing more current, the voltage drop across its emitter resistor increases.
  • This forces a reduction in that transistor's base-emitter voltage, compensating for the initial current increase.
  • Don't confuse with: swamping resistors or collector feedback bias—same principle (local negative feedback) but different application context.

Example: Three NPN transistors in parallel for high-power output. Without emitter resistors, the hottest device would eventually carry all the load current and fail. With small emitter resistors, current distribution remains balanced even as temperatures vary.

🎛️ Flexible biasing and compensation

🎛️ V_BE multiplier

  • Replaces simple diode chains with an adjustable bias voltage generator.
  • Uses a transistor with two resistors: R₁ in series with R₂, where R₂ is in parallel with V_BE.
  • The voltage from point A to point B equals V_BE × (1 + R₁/R₂).
  • Key advantage: any voltage ratio can be set, and using a potentiometer makes bias adjustable for optimal crossover distortion reduction.
  • A capacitor (C_B) shunts the multiplier to ensure it acts as an AC short.

Example: To generate four V_BE drops, make R₁ three times as large as R₂. The total voltage will be 4 × V_BE, matching the requirements of a Darlington output stage.

📡 Miller compensation capacitor

  • A capacitor (C_M) uses the Miller Effect to create a much larger equivalent input capacitance.
  • Appears in parallel with the input of the driver transistor, creating a lag network.
  • Allows the designer to tailor the high-frequency response of the amplifier.

🌉 Bridging for increased power

🌉 How bridging works

  • Two identical amplifiers drive opposite ends of a floating (non-grounded) load.
  • The second amplifier receives an inverted copy of the input signal.
  • As the left output goes positive, the right output goes negative by the same amount.
  • The load sees twice the voltage it would from a single amplifier.
  • Power increase: voltage doubles → power quadruples (power varies as voltage squared).

🚗 Practical considerations

  • Benefit: in automotive systems (12 VDC supply), bridging can increase output from 2.25 watts to 9 watts into an 8 Ω load.
  • Downside 1: requires doubling of output circuitry (four transistors form an H-bridge).
  • Downside 2: the load is not grounded—both terminals must be wired, so single-wire chassis-ground loudspeaker systems cannot be used without rewiring.
  • Risk: connecting one loudspeaker terminal to chassis ground shorts out one side of the bridge, potentially damaging output transistors.
ConfigurationLoad voltageLoad powerCircuit complexityLoad grounding
Single-endedV_supplyPStandardOne terminal grounded
Bridged2 × V_supply4 × PDoubled (H-bridge)Floating (both wired)

🔌 Power supply bypassing

🔌 Why bypassing matters

  • DC power supplies must present good AC grounds, but PC board layouts and wiring constraints make this difficult.
  • Power amplifiers require larger, higher-quality bypass capacitors than small-signal amplifiers.
  • Challenge: large, high-quality capacitors (e.g., 10 μF polypropylene) are expensive compared to aluminum electrolytics.

🔧 Practical bypass strategy

  • Aluminum electrolytics have higher leakage, higher ESR, and poor high-frequency behavior (impedance increases due to inductive effects).
  • Solution: parallel a large aluminum electrolytic with a much smaller polyester or polypropylene capacitor.
  • The electrolytic provides low reactance at lower frequencies; the poly capacitor shunts it at higher frequencies.
  • Result: nearly as effective as a single large high-quality capacitor but much less expensive.

Example: A 1000 μF aluminum electrolytic in parallel with a 1 μF polypropylene capacitor provides good bypassing from low to high frequencies at a fraction of the cost of a 1000 μF polypropylene capacitor.

63

Junction Field Effect Transistors (JFETs)

Chapter 10: Junction Field Effect Transistors (JFETs)

🧭 Overview

🧠 One-sentence thesis

The JFET is a voltage-controlled current source that uses a reverse-biased junction to control current flow through a channel, fundamentally differing from the current-controlled BJT and requiring different biasing approaches.

📌 Key points

  • What a JFET is: a field effect transistor with a channel controlled by a gate junction, available as N-channel or P-channel types.
  • Voltage-controlled vs current-controlled: JFETs are modeled as voltage-controlled current sources, while BJTs are current-controlled current sources.
  • Reverse-biased operation: JFETs achieve current control via a reverse-biased gate-source junction, unlike BJTs which use a forward-biased base-emitter junction.
  • Common confusion: JFETs and BJTs are not interchangeable—their biasing circuits are incompatible, and each has different characteristics suited to different applications.
  • Complementary use: combining BJTs and FETs can produce superior performance compared to using either device alone.

🔧 JFET Structure and Terminals

🏗️ Physical structure

Channel: the main portion of the device through which current flows, built upon a substrate of oppositely doped material.

  • An N-channel device has an N-type channel built on a P-type substrate.
  • Two terminals attach to opposing ends of the channel: the source and the drain.
  • A region of opposite material type is embedded within the channel, with a lead called the gate attached to it.

📍 Terminal correspondence

The JFET terminals roughly correspond to BJT terminals:

  • Drain ≈ collector
  • Source ≈ emitter
  • Gate ≈ base

Important: This correspondence is not perfect; the devices operate on different principles.

⚖️ Symmetry considerations

  • Some JFETs are designed symmetrically, allowing drain and source terminals to be swapped with no change in operation.
  • This is not true for all devices—check the specific device design.
  • Under normal operation: drain current equals source current (I_D = I_S).

⚡ Operating Principles

🔌 Channel resistance

  • For small drain-source voltages, the channel exhibits resistance.
  • This resistance depends on:
    • Doping level
    • Physical layout of the device

📈 Current behavior with increasing voltage

When a positive voltage V_DD is applied to the drain (with gate supply at zero):

  1. Resistive region: Initially, increasing drain-source voltage causes proportional increase in channel current—the channel acts like a resistor.
  2. Saturation region: At some point, current saturates and no further increases occur despite further increases in V_DD and V_DS—the device behaves as a constant current source.
  3. Breakdown region: If drain-source voltage increases too much, breakdown occurs and current begins to increase rapidly.

Pinch-off voltage (V_p): the drain-source voltage where the transition from resistive to saturation region occurs.

🎛️ Gate voltage control mechanism

When the gate supply is made more negative:

  • The gate-source PN junction becomes reverse-biased.
  • A larger depletion region forms and widens into the channel.
  • The widened depletion region restricts current flow, causing:
    • Current saturation to occur sooner
    • Lower current levels overall
  • The more negative V_GG becomes, the lower I_D becomes.
  • Eventually, sufficient negative V_GG can reduce current to very low levels.

Key insight: Current control is achieved through the size of the depletion region, not through forward-biasing a junction.

🔄 Comparison with BJTs

🆚 Fundamental differences

AspectJFETBJT
Control modelVoltage-controlled current sourceCurrent-controlled current source
Junction biasReverse-biased gate-source junctionForward-biased base-emitter junction
Control mechanismDepletion region sizeBase current
Biasing compatibilityIncompatible with BJT circuitsIncompatible with JFET circuits

🔀 Device substitution

  • Cannot swap: One device cannot be swapped out for the other due to incompatible biasing schemes.
  • Different characteristics: JFETs and BJTs have different performance characteristics.
  • Application-specific: Each type lends itself to applications where the other might be wanting.
  • Don't confuse: "Better or worse" is not the right framework—they are suited to different applications.

🤝 Combined use

The excerpt emphasizes that judicious use of a combination of BJTs and FETs can produce superior performance compared to either device used alone, suggesting complementary strengths rather than competition.

📚 FET Types and Variants

🌳 FET family tree

Field effect transistors (FETs) are available in two broad types:

  • JFET: Junction FET (covered in this chapter)
  • MOSFET: Metal Oxide Semiconductor FET (covered in subsequent chapters)

🔀 Channel types

Like BJTs have NPN and PNP variants, FETs come in two "flavors":

  • N-channel type
  • P-channel type

The excerpt focuses on N-channel JFETs for explanation purposes.

64

Junction Field Effect Transistors (JFETs)

10.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

The JFET is a voltage-controlled semiconductor device that differs fundamentally from the BJT by using a reverse-biased junction to control current, offering unique characteristics that make it suitable for applications where BJTs may be inadequate.

📌 Key points (3–5)

  • What a JFET is: A field effect transistor that serves as an alternative to the BJT, available in two types (JFET and MOSFET) and two "flavors" (N-channel and P-channel).
  • How it differs from BJTs: JFETs are voltage-controlled current sources (vs. BJT's current-controlled), use reverse-biased junctions (vs. BJT's forward-biased base-emitter), and exhibit square-law behavior (vs. BJT's logarithmic).
  • Key operating parameters: Maximum current IDSS occurs at VGS = 0V; pinch-off voltage VP marks transition to constant current; VGS(off) is the turn-off voltage.
  • Common confusion: JFET biasing circuits are incompatible with BJT schemes—one device cannot simply be swapped for the other.
  • Chapter scope: Covers internal structure, theory of operation, DC biasing techniques, and comparison with BJTs; small signal amplifiers are covered in the next chapter.

🔬 JFET structure and operation

🧱 Internal structure

Channel: The main conduction path in a JFET, built on a substrate of oppositely doped material.

Source and drain: Terminals attached to opposing ends of the channel (roughly analogous to BJT's emitter and collector).

Gate: A terminal attached to a region of opposite material type embedded within the channel (roughly analogous to BJT's base).

  • Some devices are symmetric—drain and source can be swapped with no change in operation.
  • For small drain-source voltages, the channel exhibits resistance dependent on doping level and physical layout.
  • Under normal operation, ID equals IS.

⚡ How current control works

The JFET controls current through depletion region manipulation:

  • At VGS = 0V: Maximum current flows; as VDS increases from zero, current initially rises proportionally (ohmic region), then saturates at the pinch-off voltage VP (constant current region).
  • Negative VGS: Reverse-biases the gate-source PN junction, widening the depletion region into the channel and restricting current flow.
  • More negative VGS: Further reduces ID until reaching VGS(off), where drain current drops to virtually zero (small leakage current ID(off) remains).
  • Positive VGS: Must be avoided—forward-biases the PN junction and loses current control.

Example: Think of it like a water valve—turning VGS more negative is like turning off the spigot and decreasing the flow.

Don't confuse: The JFET's current control operates entirely in the second quadrant; the largest drain current flows when VGS = 0V (called IDSS), unlike BJTs where forward bias increases current.

📊 Operating regions

RegionConditionBehaviorNotes
Ohmic (triode)VDS < VPProportional ID increase with VDSChannel acts like resistor
Constant current (saturation)VDS > VPID levels out, independent of VDSNormal amplifier operation
BreakdownVDS > BV(DGS)Rapid current increaseAvoid this region

📐 JFET characteristic equations

🧮 Square-law relationship

The fundamental equation relating drain current to gate-source voltage (valid for constant current region, VDS > VP):

ID = IDSS × (1 − VGS / VGS(off))²

Where:

  • VGS is the gate-source voltage (VGS(off) ≤ VGS ≤ 0)
  • ID is the drain current
  • IDSS is the maximum current (at VGS = 0V)
  • VGS(off) is the turn-off voltage

This is a square-law device characteristic (portion of a parabola), much more gradual in slope than a BJT's logarithmic characteristic—important implications for voltage gain and distortion.

📈 Transconductance

Transconductance (gm or gfs): The slope of the characteristic curve plotting output current versus input voltage, measured in siemens.

Maximum transconductance (at VGS = 0V):

  • gm0 = −2 × IDSS / VGS(off)

Transconductance at any operating point:

  • gm = gm0 × (1 − VGS / VGS(off))

Key insight: Transconductance is a linear function of VGS, unlike the square-law current relationship.

Useful derived relationships:

  • gm / gm0 = √(ID / IDSS)
  • ID / IDSS = (gm / gm0)²

📉 Parameter variation

Important practical consideration: VGS(off) and IDSS vary significantly between devices of the same model.

Example from J111 series:

  • VGS(off) can range from −3V to −10V
  • IDSS can range from 2mA to 20mA
  • Generally, the most negative VGS(off) values associate with the largest IDSS values

🔌 DC biasing model

🎛️ Basic DC model

The JFET DC model consists of:

  • A voltage-controlled current source ID in the drain
  • Very large resistance RGS between gate and source (reverse-biased PN junction, ideally infinite)

Key consequence: Gate current IG ≈ 0 in most practical circuits, therefore ID = IS.

🔧 Biasing topologies overview

Four main biasing schemes, ranked by Q-point stability (least to most stable):

  1. Constant voltage bias (least stable)
  2. Self bias (modest stability)
  3. Combination bias (improved stability)
  4. Constant current bias (most stable)

🔋 Constant voltage bias

⚙️ Circuit operation

  • Simplest form: uses only a couple of resistors and power sources
  • VGG source drops across RG and gate-source junction
  • Since IG ≈ 0, the drop across RG ≈ 0, so VGS = VGG

Analysis steps:

  1. VGS = VGG (because IG ≈ 0)
  2. Find ID using the characteristic equation with IDSS and VGS(off)
  3. Calculate VDS from KVL: VDS = VDD − ID × RD

⚠️ Stability issues

  • Does not exhibit a stable Q-point
  • Wide variation in drain current with device parameter changes
  • Reminiscent of simple base bias for BJTs

Don't confuse: While computation is simple, this bias is unsuitable for applications requiring stable Q-points; use only when Q-point stability is not critical.

Advantage: Minimum component count.

🔄 Self bias (automatic bias)

🔁 How it works

Self bias: Uses drain current to create a voltage drop that sets up the gate-source voltage—the circuit "biases itself."

Key relationship: VGS = −ID × RS

  • RG connects directly to ground, so VG ≈ 0V
  • Magnitude of VGS equals voltage across RS
  • Definition is self-referential, requiring iterative or graphical solution

📊 Analysis methods

Graphical method: Use the self bias curve plotting normalized ID/IDSS versus gm0 × RS.

Steps:

  1. Calculate gm0 = −2 × IDSS / VGS(off)
  2. Find gm0 × RS
  3. Read normalized current ratio from curve
  4. Multiply by IDSS to get ID

Iterative approximation:

  1. Guess VGS (typically half of VGS(off))
  2. Calculate ID from characteristic equation
  3. Calculate ID from Ohm's law: ID = −VGS / RS
  4. If results don't agree, adjust VGS estimate and repeat until convergence

🛡️ Stability mechanism

Self bias gains stability through feedback:

  • If ID increases → larger voltage drop across RS
  • Larger VRS means VGS grows more negative
  • More negative VGS reduces ID, opposing the initial change

Similar in function to BJT collector feedback bias.

Design tip: For a target ID or VGS, rearrange to find required RS: RS = −VGS / ID

⚡ Combination bias (source bias)

🔌 Enhanced stability

Based on self bias but adds a negative power supply connected to RS, enhancing stability of ID, VDS, and gm.

Key difference: The source power supply increases voltage drop across RS:

  • VRS = ID × RS = |VGS| + |VSS|

If VSS >> VGS, then ID ≈ VSS / RS (very stable approximation).

📐 Swamping factor

The analytical solution introduces a "swamping factor" k:

  • k = VSS / VGS(off)
  • When k = 0: reverts to self bias formula
  • When k is very large: ID ≈ VSS / RS

Analysis uses combination bias curves for different k values (k = 2, 3, 4), plotting normalized ID/IDSS versus gm0 × RS.

🎯 Flexibility and stability

  • Two variables (RS and VSS) provide design flexibility
  • Can be realized using positive voltage divider at gate with RS returned to ground
  • Variation in ID is reduced compared to self bias (at expense of VGS variation)
  • For large VSS with correspondingly large RS, bias line becomes nearly horizontal, indicating very stable Q-point

🎚️ Constant current bias

🔒 Maximum stability

Uses a BJT to establish a very stable current source for the JFET.

Circuit configuration:

  • NPN BJT for N-channel JFET (PNP for P-channel, flipped top-to-bottom)
  • BJT configured as two-supply emitter bias
  • Base tied directly to ground (emitter at about −0.7 VDC)
  • Remainder of VEE drops across RE, establishing emitter current
  • BJT collector connects to JFET source: IS ≈ IE

📝 Analysis procedure

Straightforward computation without graphical aids:

  1. Examine BJT emitter loop: IE = (|VEE| − 0.7V) / RE
  2. Once IE is found, IS and ID are known
  3. Find remaining component potentials using Ohm's law and KVL

Requirement: IE should not be programmed larger than IDSS.

🔍 Characteristics

  • Most stable bias for Q-point (fixed ID)
  • Does not guarantee most stable voltage gain
  • Widest VGS variation of all biasing circuits when JFET is changed
  • If VGS is needed, rearrange characteristic equation: VGS = VGS(off) × (1 − √(ID / IDSS))

Don't confuse: Fixed ID provides Q-point stability but not necessarily optimal gain stability—combination bias might achieve that goal more easily.

📚 Data sheet interpretation

📄 Reading JFET specifications

Example from J111 series N-channel JFET:

Absolute maximum ratings:

  • Maximum drain-gate and gate-source voltages: 35V
  • Maximum power dissipation: 625 mW
  • Typical of small signal devices

Electrical characteristics:

  • Large VGS(off) variation: J111 ranges from −3V to −10V minimum/maximum
  • J112 and J113 exhibit even wider min/max ratios
  • Larger VGS(off) ranges associate with larger IDSS maximums

Performance curves show:

  • Drain curve families (ohmic region detail)
  • Characteristic curves (may be rotated around vertical axis)
  • Large IDSS values tend to pair with large VGS(off) values
  • Thermal variations are significant: as temperature increases, characteristic curve becomes less steep
  • Transconductance vs VGS plots show temperature effects and non-perfect linearity

🔬 Lab measurement technique

Simple method to measure IDSS and VGS(off):

Measuring IDSS:

  1. Ground gate and source terminals (forces VGS = 0V)
  2. Insert ammeter between VDD and drain
  3. Set VDD higher than VP (+15 VDC generally sufficient)
  4. Ammeter reading is IDSS

Measuring VGS(off):

  1. Leave ammeter in drain
  2. Connect gate to adjustable negative power supply
  3. Turn supply more negative until ammeter reads zero (< 1% of IDSS)
  4. Voltage source equals VGS(off)

🔄 BJT vs JFET comparison

🆚 Fundamental differences

AspectBJTJFET
Control typeCurrent-controlled current sourceVoltage-controlled current source
Junction biasForward-biased base-emitterReverse-biased gate-source
CharacteristicLogarithmic (Shockley equation)Square-law (parabolic)
Input currentSignificant base currentNegligible gate current (IG ≈ 0)
Biasing compatibilityIncompatible with JFET schemesIncompatible with BJT schemes

🎯 When to use each

Neither is universally "better"—they have different characteristics:

JFET advantages:

  • Much higher input impedance potential
  • Lower noise
  • Better high frequency performance

BJT advantages:

  • Higher voltage gain potential
  • More predictable parameters

Best approach: Judicious combination of BJTs and FETs can produce superior performance compared to either device used alone, with each playing to its strengths.

65

JFET Small Signal Amplifiers

10.1 Introduction

🧭 Overview

🧠 One-sentence thesis

JFET amplifiers offer higher input impedance and lower noise than BJT amplifiers, though with less voltage gain potential, and can be combined with BJTs to leverage the strengths of both device types.

📌 Key points (3–5)

  • Core parameter: transconductance (g_m) is the key AC parameter for JFETs, comparable to r'_e for BJTs.
  • Trade-offs vs BJT: JFETs have less voltage gain potential but offer much higher input impedance, lower noise, and better high-frequency performance.
  • Common source amplifier: analogous to the common emitter amplifier; inverts the signal and can use swamping to reduce distortion at the expense of gain.
  • Dual-mode operation: JFETs work as amplifiers in the active region and as voltage-controlled resistors or analog switches in the ohmic region.
  • Common confusion: the AC model shown is simplified for low frequencies only—at higher frequencies, device capacitances (C_GS, C_DG, C_DS) significantly affect performance.

🔧 JFET vs BJT comparison

🔧 Performance trade-offs

CharacteristicJFETBJT
Voltage gain potentialLowerHigher
Input impedanceMuch higherLower
NoiseLowerHigher
High-frequency performanceBetterWorse
  • Both types invert the signal in voltage amplifier configurations.
  • Both can use swamping to lower distortion at the expense of voltage gain.
  • Combining JFETs and BJTs allows each to play to its strengths, potentially outperforming single-device designs.

🎯 Key AC parameter

Transconductance (g_m): the key parameter for JFET AC analysis, comparable in importance to r'_e in a BJT.

  • In fact, 1/r'_e is g_m for a BJT.
  • g_m controls the voltage-controlled current source in the drain.

🧩 Simplified AC model

🧩 Model structure

  • The JFET AC model is essentially the same as the DC model.
  • It contains a voltage-controlled current source in the drain.
  • The reverse-biased gate-source junction appears as a very large resistance, r_GS.

⚠️ Frequency limitations

  • Important: this model is suitable only for low frequencies.
  • At higher frequencies, device capacitances play a major role and cannot be ignored.
  • Three device capacitances (not shown in the simplified model) shunt each pair of terminals: C_GS, C_DG, and C_DS.

📊 Lumped capacitances on data sheets

  • C_iss: capacitance looking into the gate with source and drain shorted to ground; C_iss = C_GS + C_DG.
  • C_rss: capacitance seen from the drain with gate and source shorted to ground; C_rss = C_DS + C_DG.
  • These capacitances can have a sizable impact on amplifier characteristics such as input impedance (Z_in).

🎚️ Common source amplifier

🎚️ Circuit analogy

Common source amplifier: analogous to the common emitter amplifier.

  • Input signal is presented to the gate terminal.
  • Output is taken from the drain.
  • The circuit can include a swamping resistor, r_S, in the source.

📈 Voltage gain derivation

  • Voltage gain A_v is the ratio of output voltage (v_out) to input voltage (v_in).
  • Starting from the fundamental definition and expressing voltages in Ohm's law terms:
    • A_v = v_out / v_in = v_D / v_G
    • A_v = (−i_D · r_L) / (i_D · r_S + v_GS)
    • A_v = (−g_m · v_GS · r_L) / (g_m · v_GS · r_S + v_GS)
    • Final form: A_v = −(g_m · r_L) / (g_m · r_S + 1)

🔄 Swamping resistor role

  • If there is no swamping resistor (r_S = 0), the gain simplifies to −g_m · r_L.
  • The swamping resistor r_S plays the same role as in the BJT:
    • Helps stabilize the gain.
    • Reduces distortion.
    • Does so at the expense of voltage gain.

🔌 Input impedance

  • The input impedance of the amplifier will be r_G (the excerpt cuts off here, but the model shows r_GS is very large, contributing to high input impedance).

🔀 Ohmic region applications

🔀 Non-amplifier uses

  • JFETs can be used in their ohmic region, not just as amplifiers or followers.
  • In this mode, the device no longer behaves as a constant current source.

🎛️ Voltage-controlled resistor

  • The channel resistance becomes a function of the gate-source voltage.
  • Can be used as a control element within a voltage divider.
  • Has the capability of changing resistance value much faster than a mechanical potentiometer.

🔌 Analog switches

  • JFETs in the ohmic region can function as analog switches.
  • Example: an organization could use a JFET to rapidly switch signal paths without mechanical wear.
66

10.2 JFET Internals

10.2 JFET Internals

🧭 Overview

🧠 One-sentence thesis

The JFET amplifier uses transconductance (g_m) as its key AC parameter and offers higher input impedance and lower noise than BJT amplifiers, though with less voltage gain potential.

📌 Key points (3–5)

  • Key AC parameter: transconductance (g_m) plays the same role in JFETs as r'_e does in BJTs (in fact, 1/r'_e is g_m for a BJT).
  • JFET vs BJT trade-offs: JFETs have less voltage gain potential but offer much higher input impedance, lower noise, and better high-frequency performance.
  • Swamping still applies: like BJT amplifiers, JFETs can use swamping resistors to lower distortion at the expense of voltage gain.
  • Common confusion—frequency effects: the simplified AC model is suitable only for low frequencies; at higher frequencies, device capacitances (C_GS, C_DG, C_DS) significantly affect amplifier characteristics like input impedance.
  • Multiple operating modes: JFETs work as amplifiers/followers (constant current source mode) and also in the ohmic region as voltage-controlled resistors or analog switches.

🔧 JFET characteristics and comparisons

🔧 JFET vs BJT performance trade-offs

CharacteristicJFETBJT
Voltage gain potentialLowerHigher
Input impedanceMuch higher (megohms)Lower
NoiseLowerHigher
High-frequency performanceBetterWorse
Signal inversionYes (voltage amplifier inverts)Yes (common emitter inverts)
  • Both device types can be combined in designs, each playing to their strengths, potentially outperforming single-device-type designs.
  • Example: A designer might use a JFET input stage for high impedance and low noise, followed by a BJT stage for higher gain.

⚙️ Operating modes

Amplifier/follower mode (constant current source):

  • The channel acts as a constant current source.
  • Used for voltage amplifiers and voltage followers.
  • Swamping remains possible to reduce distortion.

Ohmic region mode:

  • The device no longer behaves as a constant current source.
  • Channel resistance becomes a function of gate-source voltage.
  • Applications: voltage-controlled resistors and analog switches.
  • Advantage: can change resistance value much faster than a mechanical potentiometer (used as a control element within a voltage divider).

🔌 Simplified AC model

🔌 Model structure

The AC model of the JFET is essentially the same as the DC model, featuring a voltage-controlled current source in the drain and a very large resistance (r_GS) at the reverse-biased junction.

  • The model shown in Figure 11.1 includes:
    • Voltage-controlled current source at the drain
    • Very large resistance r_GS (well into the megohms)
    • Gate-source voltage v_GS controlling the drain current

⚠️ Low-frequency limitation

Important: This model is suitable only for low frequencies.

At higher frequencies, three device capacitances (not shown in the basic model) shunt each pair of terminals:

  • C_GS: gate-to-source capacitance
  • C_DG: drain-to-gate capacitance
  • C_DS: drain-to-source capacitance

Lumped capacitances (often given on data sheets):

  • C_iss (input capacitance): capacitance looking into the gate with source and drain shorted to ground = C_GS + C_DG
  • C_rss (reverse transfer capacitance): capacitance from drain with gate and source shorted to ground = C_DS + C_DG

These capacitances can have a sizable impact on amplifier characteristics such as input impedance.

🎯 Transconductance (g_m)

Transconductance (g_m) is the key AC parameter for JFETs, roughly of equal importance to r'_e in a BJT.

  • In fact, 1/r'_e is g_m for a BJT.
  • This parameter determines the strength of the voltage-to-current conversion in the device.

🔊 Common source amplifier

🔊 Circuit analogy

The common source amplifier is analogous to the common emitter amplifier.

  • Input signal is presented to the gate terminal.
  • Output is taken from the drain.
  • The circuit can include a swamping resistor r_S in the source.

📐 Voltage gain

For swamped amplifier:

  • Voltage gain A_v = - (g_m × r_L) / (g_m × r_S + 1)
  • The negative sign indicates signal inversion.

For non-swamped amplifier (no swamping resistor):

  • The first portion of the denominator drops out.
  • Gain simplifies to: A_v = - g_m × r_L

Role of swamping resistor r_S:

  • Plays the same role as in BJT amplifiers.
  • Helps stabilize the gain and reduce distortion.
  • Does so at the expense of voltage gain.

Example: If g_m = 5.35 mS and r_L = 2 kΩ with no swamping, A_v = -5.35 mS × 2 kΩ = -10.7 (inverted, gain of 10.7).

🔌 Input impedance

Basic relationship:

  • Input impedance Z_in = r_G || Z_in(gate)
  • For non-swamped case: Z_in(gate) = r_GS
  • At low frequencies, r_GS is very large (well into the megohms).
  • In most practical circuits, r_G is much lower, so: Z_in ≈ r_G

For swamped amplifiers:

  • Theoretically, Z_in(gate) will be higher than r_GS.
  • This is a moot point—in either case, it's relatively easy to obtain high input impedance.
  • Much easier than for typical single-device BJT amplifiers.

⚠️ High-frequency input impedance effects

Don't become complacent: assuming r_G simply sets input impedance is a mistake at higher frequencies.

  • With impedances this high, junction capacitance cannot be ignored.
  • Example: A typical general-purpose device has C_iss ≈ 5 to 10 pF.
  • This capacitance appears in parallel with r_G.
  • At 100 kHz ultrasonic signals: capacitive reactance X_C ≈ 160 kΩ
    • High compared to typical BJT circuits
    • But less than the R_G values commonly used for biasing
  • At higher frequencies: X_C decreases with frequency, making the situation worse.
  • The Miller effect (not detailed here) makes the situation "even worser."

📤 Output impedance

Determining output impedance:

  • From the vantage point of R_L, looking back into the amplifier: see R_D in parallel with the impedance at the drain.
  • At the drain: the current source i_D has very high internal impedance (hundreds of kΩ).
  • This is very high compared to typical R_D values.
  • Therefore: Z_out ≈ R_D

🔄 Bias compatibility

All DC bias forms from the previous chapter can be used, with some limitations:

Bias typeSwamping possible?Notes
Constant voltage biasNoDoes not use a source resistor
Self biasYesIncludes source resistor; R_S may need to be split and partially bypassed
Combination biasYesIncludes source resistor; R_S may need to be split and partially bypassed
Constant current biasNot typicalWould require additional work to fit in a new R_S; more typically, just bypass the current source with a capacitor for non-swamped amplifier

🧮 Design calculations

🧮 Finding transconductance

To calculate voltage gain, first determine g_m:

  1. Find drain current I_D (depends on bias circuit)
  2. Find maximum transconductance g_m0:
    • g_m0 = -2 × I_DSS / V_GS(off)
  3. Find actual transconductance g_m:
    • g_m / g_m0 = square root of (I_D / I_DSS)
    • Therefore: g_m = g_m0 × square root of (I_D / I_DSS)

Example (from Figure 11.4, constant current bias):

  • Given: I_DSS = 15 mA, V_GS(off) = -3 V
  • I_D = (|V_EE| - 0.7 V) / R_E = (5 V - 0.7 V) / 1 kΩ = 4.3 mA
  • g_m0 = -(-30 mA) / (-3 V) = 10 mS
  • g_m = 10 mS × square root of (4.3 mA / 15 mA) = 5.35 mS

🔍 Input impedance by inspection

For the unswamped common source amplifier with constant current bias:

  • Z_in = Z_in(gate) || R_G
  • Z_in ≈ 10 MΩ (when R_G = 10 MΩ and r_GS is much larger)
67

JFET Data Sheet Interpretation

10.3 JFET Data Sheet Interpretation

🧭 Overview

🧠 One-sentence thesis

JFET amplifiers offer very high input impedance compared to BJT circuits, but capacitance effects at high frequencies and swamping trade-offs must be carefully considered in practical designs.

📌 Key points (3–5)

  • Input impedance advantage: JFETs achieve much higher input impedance than single-device BJT amplifiers, primarily set by the gate resistor r_G.
  • Capacitance limitation: At high frequencies, input capacitance (C_iss, typically 5–10 pF) creates a parallel reactance that can reduce effective input impedance below the gate resistor value.
  • Output impedance approximation: Output impedance is approximately equal to the drain biasing resistor R_D because the drain current source has very high internal impedance (hundreds of kΩ).
  • Swamping trade-off: Adding an unbypassed source resistor (swamping) reduces gain but improves distortion and symmetry in the output waveform.
  • Common confusion: Don't assume r_G alone sets input impedance—junction capacitance becomes significant at ultrasonic and higher frequencies, and the Miller effect worsens the situation further.

🔌 Input impedance characteristics

🔌 How input impedance is determined

Input impedance Z_in equals the gate resistor r_G in parallel with the impedance looking into the gate terminal, Z_in(gate).

  • For non-swamped circuits, Z_in(gate) is r_GS (gate-to-source resistance).
  • At low frequencies, r_GS is very large (well into the megohms).
  • In most practical circuits, r_G is much lower than r_GS, so the approximation simplifies to: Z_in ≈ r_G.
  • This makes achieving high input impedance relatively easy compared to typical single-device BJT amplifiers.

⚡ Capacitance effects at high frequencies

  • The problem: With very high impedances, junction capacitance cannot be ignored.
  • Typical values: General-purpose devices have total input capacitance C_iss around 5 to 10 pF.
  • This capacitance appears in parallel with r_G.

Example: For ultrasonic signals at 100 kHz, the capacitive reactance X_C would be as low as 160 kΩ—high compared to typical BJT circuits but less than the r_G values commonly used for biasing.

  • Frequency dependence: At higher frequencies, X_C decreases further, worsening the situation.
  • Additional factor: The Miller effect makes the problem even worse (the excerpt humorously notes this makes it "even worser").
  • Don't confuse: The high DC input impedance does not guarantee high impedance at all frequencies—capacitance dominates at higher frequencies.

🔋 Output impedance characteristics

🔋 How output impedance is approximated

  • From the load's perspective looking back into the amplifier, the load sees R_D in parallel with the impedance at the drain.
  • At the drain, there is a current source i_D with very high internal impedance (hundreds of kΩ) compared to typical R_D values.
  • Approximation: Z_out ≈ R_D

🔧 Bias configuration compatibility

All DC bias forms from the previous chapter can be used, with some limitations:

Bias typeSwamping possible?Notes
Constant voltage biasNoDoes not use a source resistor
Self biasYesIncludes source resistor; R_S may need to be split and partially bypassed
Combination biasYesIncludes source resistor; R_S may need to be split and partially bypassed
Constant current biasNot typicallyWould require additional work to fit in a new R_S; more commonly the current source is just bypassed with a capacitor for non-swamped operation

📐 Voltage gain calculations

📐 Gain formula for unswamped amplifiers

  • General formula: A_v = negative g_m times r_L divided by (g_m times r_S plus 1)
  • For unswamped case (r_S = 0): The formula reduces to A_v = negative g_m times r_L

🔢 Calculation procedure

The examples demonstrate a systematic approach:

  1. Find drain current I_D from the bias configuration
  2. Calculate maximum transconductance g_m0: negative 2 times I_DSS divided by V_GS(off)
  3. Find actual transconductance g_m: g_m0 times the square root of (I_D divided by I_DSS)
  4. Apply gain formula with the appropriate load resistance r_L

Example from constant current bias: With I_DSS = 15 mA, V_GS(off) = −3 V, and I_D = 4.3 mA, the calculation yields g_m0 = 10 mS, then g_m = 5.35 mS, resulting in A_v = −8.02.

Shortcut for self bias: Instead of finding I_D directly, use the self-bias graph with the normalized value g_m0 times R_S to find the normalized drain current, then calculate g_m from that ratio.

🎚️ Swamping effects on performance

🎚️ What swamping does

  • Implementation: Split the source resistor into two parts; bypass only the larger portion, leaving a small unbypassed resistor (the swamping resistor).
  • Gain reduction: The unbypassed portion appears as r_S in the gain formula, reducing overall gain.
  • Distortion improvement: Swamping improves output waveform symmetry and reduces total harmonic distortion (THD).

📊 Simulation comparison

The excerpt describes a simulation with R_S = 1 kΩ:

Non-swamped version:

  • Full 1 kΩ bypassed
  • Calculated gain: A_v = −12.3
  • Simulated gain: approximately −11.75 (about 5% low)
  • Distortion visible: positive peaks broader than negative peaks and don't reach the same magnitude

Swamped version:

  • Split into 200 Ω (unbypassed) and 800 Ω (bypassed)
  • Calculated gain: A_v = −4.96
  • Simulated gain: −4.85 (just a few percent low)
  • Better symmetry in output waveform
  • Input signal raised to 240 mV peak (from 100 mV) to maintain similar output amplitude

Don't confuse: Swamping does not change the DC bias (I_D and g_m remain the same), but it does change the AC gain by introducing negative feedback through the unbypassed source resistor.

68

JFET and MOSFET Amplifiers and Power Switching

10.4 JFET Biasing

🧭 Overview

🧠 One-sentence thesis

Field-effect transistors (JFETs and MOSFETs) can be configured as voltage amplifiers, followers, and high-efficiency power switches, offering very high input impedance and distinct advantages over BJTs in specific applications.

📌 Key points (3–5)

  • JFET/MOSFET amplifiers offer very high input impedance due to extremely low gate current, making them ideal for front-end stages.
  • Common source amplifiers invert and amplify with moderate gain determined by transconductance (gm) and load resistance.
  • Source followers provide unity gain with high input impedance and low output impedance, similar to emitter followers.
  • Class D amplifiers achieve near-100% efficiency by operating transistors only as switches (on/off), not in the linear region.
  • IGBTs combine advantages of BJTs and MOSFETs: low on-state losses like BJTs with easy voltage-controlled drive like MOSFETs.

🔌 JFET and MOSFET Small-Signal Amplifiers

🎚️ Common source voltage amplifiers

Common source amplifier: FET configuration where input is applied to gate and output taken from drain, analogous to BJT common emitter.

Key characteristics:

  • Provides inverting voltage gain (negative sign in gain equation)
  • Voltage gain: Av = -gm·rL / (gm·rS + 1), where rS is swamping resistor
  • Without swamping: Av = -gm·rL (simplified)
  • Input impedance ≈ RG (gate biasing resistor) at low frequencies
  • Output impedance ≈ RD (drain resistor)

Swamping trade-offs:

  • Swamping resistor (rS) reduces gain but improves stability and lowers distortion
  • Same principle as with BJT emitters
  • Example: Unswamped amplifier showed ~4% THD; swamped version reduced to ~1.6%

Don't confuse: The load resistance may be called rL or rD depending on context—both refer to the AC drain resistance.

🔄 Source followers (common drain)

Source follower: FET voltage follower where input is at gate and output taken from source, analogous to emitter follower.

Performance characteristics:

  • Non-inverting with voltage gain approaching unity
  • Gain equation: Av = gm·rS / (gm·rS + 1)
  • Very high input impedance (≈ RG)
  • Low output impedance: Zout = RS || (1/gm)
  • Higher transconductance → gain closer to unity and lower output impedance

Biasing restrictions:

  • Cannot use zero bias (DE-MOSFET) or voltage divider bias—these ground the source terminal
  • Self bias and combination bias are suitable

🔑 Common gate amplifiers

  • Non-inverting with gain Av = gm·rL
  • Very low input impedance: Zin = RS || (1/gm)
  • High output impedance ≈ RD
  • Analogous to BJT common base configuration

⚡ DE-MOSFET vs E-MOSFET Distinctions

🔀 DE-MOSFET (Depletion-Enhancement)

Dual-quadrant operation:

  • Operates in both depletion mode (VGS < 0) and enhancement mode (VGS > 0)
  • Sometimes called "normally on" devices (conduct at VGS = 0)
  • Can use all JFET biasing schemes plus zero bias and voltage divider bias
  • Characteristic curve extends into first quadrant beyond IDSS

Zero bias:

  • Unique to DE-MOSFET
  • VGS = 0 V, so ID = IDSS and gm = gm0
  • Minimal parts count but poor Q-point stability
  • Only suitable for non-swamped common source amplifiers

⚙️ E-MOSFET (Enhancement-only)

First-quadrant only:

  • Requires VGS > VGS(th) to conduct (threshold voltage)
  • Sometimes called "normally off" devices
  • Characteristic equation: ID = k(VGS - VGS(th))²
  • Transconductance: gm = 2k(VGS - VGS(th))

Biasing approaches:

  • Voltage divider bias: RG divider sets VG = VGS (source grounded)
  • Drain feedback bias: uses feedback resistor RG between drain and gate
  • Must calculate device constant k from datasheet ID(on), VGS(on) values

Power E-MOSFETs:

  • Use vertical "trench" construction (current flows vertically, not laterally)
  • Extremely low rDS(on) (few milliohms)
  • Very high current capacity (tens to hundreds of amps)
  • Fast switching speed
  • Negative temperature coefficient of transconductance (less prone to thermal runaway)

🎛️ Class D Power Amplifiers

💡 Efficiency through switching

Why Class D is efficient:

  • Transistors operate only as switches: either fully on (saturation) or fully off (cutoff)
  • Never operate in linear region (except during brief transitions)
  • Ideal case: zero power dissipation
    • When off: current = 0, so P = V × 0 = 0
    • When on: voltage ≈ 0, so P = 0 × I = 0

Real-world losses:

  • Small on-state voltage drop (rDS(on) for MOSFETs)
  • Transition time losses (rise/fall times not instantaneous)
  • Minimize losses by using fast switching devices with low rDS(on)

📊 Pulse Width Modulation (PWM)

PWM: encoding technique where signal amplitude is represented by the width of constant-frequency pulses.

How PWM works:

  1. Compare input signal to high-frequency triangle wave using a comparator
  2. Output goes high when signal > triangle, low when signal < triangle
  3. Result: pulse train where duty cycle tracks input amplitude
  4. Triangle frequency should be ≥10× highest input frequency

Decoding PWM:

  • Use low-pass LC filter to remove high-frequency pulse components
  • Reconstitutes original signal at higher amplitude
  • Filter cutoff set to highest input signal frequency (e.g., ~20 kHz for audio)

🔌 Output configurations

ConfigurationDevicesAdvantageConsideration
Half-bridge2 (same type)Better matching, lower distortionRequires separate gate drives
Full-bridge (H-bridge)4 (same type)Doubles voltage swing, 4× powerMore complex drive circuitry

Shoot-through and dead time:

Shoot-through: large current spike when both devices in a series pair conduct simultaneously during transitions.

  • Dead time: deliberate delay ensuring one device fully turns off before other turns on
  • Prevents shoot-through damage
  • Should be minimized to reduce distortion
  • Fast-switching devices allow shorter dead time

Practical concerns:

  • High gate capacitance limits switching speed (dVG/dt = iG/C)
  • Need high-current gate drivers for fast switching
  • Power supply noise appears at output (devices "leak" supply noise)
  • Use clean, well-filtered power supplies

🔋 IGBT (Insulated Gate Bipolar Transistor)

🏗️ Structure and operation

Hybrid device:

  • Combines E-MOSFET input stage with PNP BJT output stage
  • Similar to Sziklai pair but with MOSFET replacing NPN
  • Three terminals: gate, emitter, collector
  • Available in PT (punch-through) and NPT (non-punch-through) variants

PT vs NPT:

  • PT includes N+ buffer layer → faster switching, lower on-state voltage
  • NPT omits buffer layer → simpler but slower

📈 Characteristics

Operational curves:

  • Resemble E-MOSFET curves but displaced ~1V positive from origin
  • Conduction begins at VGE(th) (gate-emitter threshold)
  • Follows square-law characteristic like E-MOSFET
  • Negative temperature coefficient (like MOSFET, unlike BJT)

Performance trade-offs:

DeviceAdvantagesDisadvantages
Power BJTLow costCurrent-controlled (complex drive), thermal runaway risk
Power E-MOSFETFast switching, voltage-controlledLower voltage/current ratings, higher cost than BJT
IGBTHigh voltage/current, voltage-controlled, low on-state lossSlower than MOSFET, more expensive than BJT, asymmetric switching

Switching asymmetry:

  • Turn-off typically 3-4× slower than turn-on
  • Example: 31 ns turn-on vs 110 ns turn-off
  • Contrast with MOSFETs which show symmetric switching

🛠️ Applications

Induction heating:

  • IGBT switches LC resonant tank at 20-30 kHz
  • Variable duty cycle controls heating intensity
  • Inductor coil creates changing magnetic field
  • Eddy currents induced in ferromagnetic cookware generate heat

DC-to-AC inversion:

  • Full-bridge IGBT configuration driven by PWM
  • LC filter smooths output to sine wave
  • Applications: UPS systems, remote power, solar inverters
  • Controller must maintain accurate frequency for timing-sensitive loads

Motor control:

  • DC motors: PWM controls average voltage → speed
  • AC motors: variable-frequency PWM controls speed
  • Flyback/snubber diodes protect against inductive spikes

DC-to-DC conversion:

  • Boost (step-up) switching regulator configuration
  • IGBT switches inductor charging/discharging
  • Output voltage = input + inductor voltage during off-state
  • High switching frequency (≥100 kHz) minimizes component size

🎯 Design Guidelines

🔧 Choosing FET amplifier configuration

When to use common source:

  • Need voltage gain with inversion
  • Can tolerate moderate input impedance (still very high)
  • Swamping available if source resistor present in bias

When to use source follower:

  • Need impedance transformation (high Zin, low Zout)
  • Unity gain acceptable
  • Buffering between stages

When to use common gate:

  • Need non-inverting gain
  • Low input impedance acceptable or desired
  • Less common in practice

⚙️ Power switching device selection

Choose E-MOSFET when:

  • Low to medium power (< few hundred watts)
  • High switching frequency required (> 100 kHz)
  • Fast transitions critical
  • Cost less critical

Choose IGBT when:

  • High voltage and current (hundreds of volts, tens of amps)
  • Medium switching frequency (< 100 kHz)
  • Efficiency more important than speed
  • Willing to pay premium over BJT

Choose power BJT when:

  • Cost is primary concern
  • Moderate performance acceptable
  • Willing to design complex base drive circuitry

🛡️ ESD precautions for MOSFETs

Why MOSFETs are vulnerable:

  • Insulated gate can be damaged by static discharge
  • Human body can develop thousands of volts unnoticed
  • Even small static charge can exceed gate voltage rating

Protection measures:

  • Store in conductive packaging (not regular plastic)
  • Use wrist straps and conductive mats during handling
  • Control humidity in work environment
  • Some devices include back-to-back Zener protection (trades off some performance)
  • Once installed on PCB, normal ESD precautions apply

Common confusion—JFET vs MOSFET input impedance: Both offer very high input impedance at low frequencies due to negligible gate current. The MOSFET's insulated gate provides even higher impedance than the JFET's reverse-biased junction, but in most practical circuits the difference is masked by biasing resistors. At high frequencies, device capacitance dominates for both.

69

Decibels and Bode Plots

Chapter 11: JFET Small Signal Amplfiers

🧭 Overview

🧠 One-sentence thesis

The decibel system and Bode plots provide powerful logarithmic tools for analyzing amplifier gain and frequency response, turning multiplication into addition and compressing wide ranges of values into manageable scales.

📌 Key points (3–5)

  • What decibels measure: logarithmic representation of gain ratios (power or voltage), making calculations simpler by converting multiplication to addition.
  • Power vs voltage decibels: power gain uses a multiplier of 10, voltage gain uses 20 (because power varies as the square of voltage).
  • Bode plots: paired graphs showing gain magnitude (in dB) and phase (in degrees) versus frequency on a logarithmic axis.
  • Lead and lag networks: lead networks cause low-frequency rolloff (positive phase shift), lag networks cause high-frequency rolloff (negative phase shift), each at 6 dB per octave.
  • Common confusion: remembering that a factor of 2 is approximately ±3 dB for power but ±6 dB for voltage; also, dB values are dimensionless ratios, while dBW/dBm/dBV indicate absolute levels relative to a reference.

📐 Decibel fundamentals

📐 What the decibel is

The decibel is a logarithmic unit used to measure system gain, where gain is the ratio of an output signal to an input signal.

  • The base unit is the Bel, named after Alexander Graham Bell.
  • One Bel equals the common logarithm (base 10) of the ordinary gain.
  • The decibel is one-tenth of a Bel, so multiply by 10 to convert Bels to decibels.
  • Unlike ordinary gains (which might be labeled W/W or V/V), decibel gains are dimensionless and denoted with a prime symbol (e.g., G′).

🔢 Power gain in decibels

For power gain:

  • Formula: G′ = 10 log₁₀(G), where G is the ordinary power gain and the result is in dB.
  • A gain greater than 1 produces a positive dB value; a gain less than 1 (a loss) produces a negative dB value.
  • A gain of exactly 1 (no change) equals 0 dB.

Example: An amplifier with input power of 10 mW and output power of 200 mW has an ordinary gain of 20. The dB gain is 10 log₁₀(20) = 13 dB.

🔢 Voltage gain in decibels

For voltage gain:

  • Formula: A′ᵥ = 20 log₁₀(Aᵥ), where Aᵥ is the ordinary voltage gain and the result is in dB.
  • The multiplier is 20 instead of 10 because power varies as the square of voltage; the dB form should be twice as large for voltage to match the power calculation.
  • This assumes matched input and output impedances; otherwise, voltage and power gains may differ.

Example: A circuit with 50 mV input and 2 V output has an ordinary gain of 40. The dB gain is 20 log₁₀(40) ≈ 32 dB.

🧮 Useful approximations

Memorizing these factors speeds up estimation:

  • Factor of 2: approximately ±3 dB for power, ±6 dB for voltage.
  • Factor of 10: exactly 10 dB for power, 20 dB for voltage.
  • Factor of 4: 6 dB for power (two doublings), 12 dB for voltage.
  • Factor of 8: 9 dB for power (three doublings), 18 dB for voltage.

Example: A gain of 800 can be written as 8 × 10². The factor of 8 is three doublings (3 + 3 + 3 = 9 dB for power), and 10² is two factors of 10 (20 dB for power). Total: 29 dB.

Don't confuse: negative dB values indicate loss (attenuation), not negative power or voltage.

📊 Absolute signal levels in decibels

📊 dBW, dBm, and dBf for power

Instead of expressing power in watts, you can express it relative to a reference level:

  • dBW: decibels relative to 1 watt. Formula: P′ = 10 log₁₀(P / 1 W).
  • dBm: decibels relative to 1 milliwatt. Formula: P′ = 10 log₁₀(P / 1 mW).
  • dBf: decibels relative to 1 femtowatt (for very low signal levels, such as antenna signals).

Positive values indicate power greater than the reference; negative values indicate power less than the reference.

Example: 200 mW expressed in dBW is 10 log₁₀(200 mW / 1 W) = −7 dBW. The same power in dBm is 10 log₁₀(200 mW / 1 mW) = 23 dBm. Note that dBW and dBm values are always 30 dB apart (because the references differ by a factor of 1000).

📊 dBV for voltage

Similarly, voltages can be expressed relative to 1 volt:

  • dBV: decibels relative to 1 volt. Formula: V′ = 20 log₁₀(V / 1 V).

Example: A 2 V signal is 20 log₁₀(2 V / 1 V) ≈ 6 dBV.

🧪 Practical use in multi-stage systems

When both gains and signal levels are in decibel form, analysis is quick:

  • To find output level: add the gain (in dB) to the input level (in dBW, dBm, or dBV).
  • To find gain: subtract input level from output level.

Example: An amplifier with 35 dB gain and an input of −42 dBV produces an output of −42 dBV + 35 dB = −7 dBV.

Don't confuse: the final units are dBV (indicating a voltage level), not dB (which indicates a dimensionless gain).

🔬 Laboratory considerations

  • A digital meter on a dBV scale can "underflow" if the signal is too weak (near zero volts), as the effective value is negative infinity dBV.
  • dBm measurements are often made with a voltmeter calibrated for a specific impedance (commonly 600 Ω). If used on a different impedance, the readings will not reflect accurate dBm values but will still show relative changes correctly.
  • Sound pressure level (dB-SPL) uses a reference of 20 micropascals (the threshold of human hearing). A "110 dB" concert level properly reads "110 dB-SPL."

📈 Bode plots and frequency response

📈 What a Bode plot is

A Bode plot is a pair of graphs: one showing signal gain or loss versus frequency, and another showing phase versus frequency.

  • Named after Hendrik Wade Bode, an American engineer.
  • Both plots use logarithmic frequency axes.
  • The gain plot uses a decibel axis; the phase plot uses degrees.
  • Bode plots are useful for circuit design, analysis, and understanding system behavior across frequencies.

📈 Key features of the gain plot

  • Midband region: the relatively flat middle portion where gain is constant (the midband gain).
  • Break frequencies: f₁ (lower) and f₂ (upper) are the frequencies where gain drops by 3 dB from midband (also called half-power points or corner frequencies).
  • Rolloff regions: at frequencies below f₁ or above f₂, gain decreases.
  • Amplifiers are normally used only for signals between f₁ and f₂.
  • Some amplifiers have no lower break frequency (DC amplifiers), but all amplifiers have an upper break.

🔄 Lead networks (low frequency response)

A lead network causes reduction in low-frequency gain; the output voltage developed across the resistor leads the input.

  • Generic form: a capacitor in series with a resistor, with output taken across the resistor.
  • At very high frequencies, the capacitor acts as a short, so output ≈ input.
  • As frequency drops, capacitive reactance increases, reducing output voltage.
  • Break frequency: fₒ = 1 / (2π RC), where the signal has decreased by 3 dB.
  • Rolloff slope: 6 dB per octave (or 20 dB per decade) below fₒ.
  • The response can be approximated as two straight-line segments (asymptotes).

Gain equation: A′ᵥ = −10 log₁₀(1 + (fₒ / f)²)

Phase equation: θ = arctan(fₒ / f)

  • At very low frequencies: phase approaches +90 degrees.
  • At very high frequencies: phase approaches 0 degrees.
  • At the critical frequency: phase is +45 degrees.

🔄 Lag networks (high frequency response)

A lag network is essentially an inverted lead network; it causes reduction in high-frequency gain, with the output voltage lagging the input.

  • Generic form: a resistor in series with a capacitor, with output taken across the capacitor.
  • At very low frequencies, the capacitor's reactance is large, so output ≈ input.
  • As frequency increases, reactance decreases, reducing output voltage.
  • Break frequency: fₒ = 1 / (2π RC).
  • Rolloff slope: −6 dB per octave (or −20 dB per decade) above fₒ.

Gain equation: A′ᵥ = −10 log₁₀(1 + (f / fₒ)²)

Phase equation: θ = −90 + arctan(fₒ / f)

  • At very low frequencies: phase is 0 degrees (output in phase with input).
  • At very high frequencies: phase approaches −90 degrees.
  • At the critical frequency: phase is −45 degrees.

Example: A lag network with an upper break frequency of 150 kHz produces a gain of −20.6 dB and a phase of −84.6 degrees at 1.6 MHz (just over one decade above the break).

⏱️ Rise time versus bandwidth

For pulse-type signals, the "speed" of a circuit is expressed in terms of rise time:

Rise time is the amount of time it takes for a signal to traverse from 10% to 90% of the peak value of a pulse.

  • A square pulse passed through a lag network becomes rounded due to capacitor charging.
  • Rise time Tᵣ ≈ 2.2 RC for a single dominant lag network.
  • Relationship to upper critical frequency: f₂ = 0.35 / Tᵣ.

Example: A lag network critical at 100 kHz has a rise time of 0.35 / 100 kHz = 3.5 μs.

🧩 Combining multiple networks

🧩 Multi-stage system response

A complete gain or phase plot combines:

  1. The midband response (flat gain).
  2. The lead response (low-frequency rolloff).
  3. The lag response (high-frequency rolloff).
  • Multiple lead and lag networks are cumulative; their effects add.
  • Dominant network: the one that affects midband response first.
    • For lead networks: the one with the highest fₒ.
    • For lag networks: the one with the lowest fₒ.

🧩 Constructing a system Bode plot

Approximate the complete response using straight-line segments:

  1. Locate all critical frequencies (fₒ) on the frequency axis.
  2. Draw a straight line between the dominant lag and lead fₒ values at the midband gain. (If no lead networks exist, extend the midband line down to DC.)
  3. Draw a 6 dB per octave slope between the dominant lead fₒ and the next lower lead network.
  4. Because effects are cumulative, increase the slope by 6 dB per octave at each additional lead fₒ (12 dB per octave after the second, 18 dB per octave after the third, etc.).
  5. Draw a −6 dB per octave slope between the dominant lag fₒ and the next highest fₒ.
  6. Increase the negative slope by −6 dB per octave at each additional lag fₒ.

Example: An amplifier with midband gain of 26 dB, one lead network at 200 Hz, and lag networks at 10 kHz and 30 kHz:

  • Draw a flat line at 26 dB from 200 Hz to 10 kHz.
  • Below 200 Hz: +6 dB per octave slope.
  • Between 10 kHz and 30 kHz: −6 dB per octave slope.
  • Above 30 kHz: −12 dB per octave slope.

⚠️ Narrowing effect

Narrowing occurs when two or more networks share similar critical frequencies, and one is dominant:

  • The true −3 dB breakpoints may shift.
  • The Bode plot gives only an approximate shape.

Example: Two lag networks both critical at 1 MHz each produce a 3 dB loss at that frequency, for a net loss of 6 dB. The true −3 dB point will have shifted to a lower frequency than 1 MHz.

70

Decibels and Bode Plots

11.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

The decibel system converts ordinary gain ratios into a logarithmic scale that simplifies calculations and compresses wide ranges of values, making it especially useful for analyzing frequency response through Bode plots.

📌 Key points (3–5)

  • What decibels measure: system gain (power or voltage) expressed logarithmically rather than as ordinary ratios.
  • Why logarithmic is useful: multiplication becomes addition, division becomes subtraction, and wide ranges compress into manageable values.
  • Key conversion shortcuts: a factor of 2 ≈ ±3 dB; a factor of 10 = ±10 dB (positive for gain, negative for loss).
  • Common confusion: voltage/current gain formulas differ from power gain formulas because power varies as the square of voltage.
  • Practical application: decibels simplify multi-stage system calculations and enable Bode plots for frequency response analysis.

🔢 What the decibel measures

🔢 Definition and basic conversion

Decibel (dB): a logarithmic unit representing gain, where the base unit (Bel) equals the common logarithm (base 10) of the ordinary gain, and one decibel equals one-tenth of a Bel.

For power gain:

  • Ordinary gain G = output power / input power
  • Bel gain = log₁₀(G)
  • Decibel gain G' = 10 × log₁₀(G)

Example: An amplifier with 200 mW output and 10 mW input has ordinary gain = 20, which converts to 10 × log₁₀(20) = 13 dB.

📊 Gain vs. loss interpretation

SituationOrdinary gaindB valueSign
AmplificationG > 1PositiveIncrease
No changeG = 10 dBUnity
AttenuationG < 1NegativeDecrease
  • Positive dB → signal increases
  • Negative dB → signal decreases (loss)
  • 0 dB → signal unchanged

🧮 Why logarithmic representation helps

🧮 Mathematical advantages

Log identities simplify operations:

  • Multiplication → addition in dB
  • Division → subtraction in dB
  • Powers → multiplication in dB
  • Roots → division in dB

Example: A three-stage amplifier with gains of 10 dB, 16 dB, and 14 dB has total gain = 10 + 16 + 14 = 40 dB (just add the stages instead of multiplying ordinary gains).

📏 Range compression

  • Very wide ranges of ordinary gain compress into smaller dB ranges
  • Ratios of change become constant offsets
  • Makes it easier to visualize and compare systems with vastly different gain levels

Don't confuse: The dB system doesn't change the actual gain; it only changes how we represent and calculate with it.

🎯 Practical conversion shortcuts

🎯 Key factors to memorize

FactordB valueInterpretation
2+3 dBDoubling
0.5−3 dBHalving
10+10 dBOne decade up
0.1−10 dBOne decade down
4+6 dBTwo doublings (3+3)
8+9 dBThree doublings (3+3+3)

🧩 Estimation technique

To estimate dB without a calculator:

  1. Break the gain into factors of 2 and 10
  2. Replace each factor: 2 → 3 dB, 10 → 10 dB
  3. Add the dB values together

Example: Gain of 800

  • 800 = 8 × 10²
  • 8 = 2×2×2 → 3+3+3 = 9 dB
  • 10² → 20 dB (or 10+10)
  • Total ≈ 29 dB

🔄 Converting back to ordinary form

To convert dB to ordinary gain:

  • Divide the dB value by 10
  • Take the antilog (10 raised to that power)
  • G = 10^(G'/10)

Example: 23 dB → 10^(23/10) = 10^2.3 ≈ 199.5

Or use the reverse shortcut: 23 dB = 3 dB + 10 dB + 10 dB = 2 × 10 × 10 = 200

⚡ Voltage vs. power gain

⚡ Why voltage needs different treatment

Critical distinction: Power varies as the square of voltage (P ∝ V²).

  • Doubling voltage → quadrupling power
  • If voltage doubles: should be 6 dB (to match the power quadrupling)
  • But using the power formula would give only 3 dB
  • Therefore, voltage and current gain require a correction factor

Don't confuse: The excerpt states that voltage/current gain formulas differ from power gain formulas, but the specific voltage formula is cut off at the end. The key point is that you cannot use the same dB equation for voltage gain as for power gain.

🔌 Notation conventions

  • G' (with prime) = decibel gain
  • G (no prime) = ordinary gain
  • Units sometimes shown as W/W for power gain, V/V for voltage gain (ordinary form)
  • "dB" always indicates logarithmic form

🎓 Chapter learning objectives

🎓 Skills covered

The chapter aims to teach:

  • Converting between ordinary and decibel-based power and voltage gains
  • Using decibel measurements during circuit analysis
  • Defining and graphing Bode plots
  • Understanding lead and lag networks and their Bode plots
  • Combining multiple network effects to determine system Bode plots

🔬 Applications mentioned

Decibels and Bode plots are particularly important in:

  • Audio systems
  • Communications
  • Amplifier design
  • Filter circuits
  • Systems using negative feedback
  • Frequency response analysis
71

Decibel Measurements and Bode Plots

11.1 Introduction

🧭 Overview

🧠 One-sentence thesis

Decibel (dB) notation simplifies gain calculations and signal-level tracking across multi-stage systems by converting multiplication into addition, and Bode plots use this logarithmic framework to visualize how circuit gain and phase change with frequency.

📌 Key points (3–5)

  • Why dB matters: Decibel form turns multiplication (for cascaded gains) into simple addition, making multi-stage analysis faster.
  • Voltage vs power dB: Voltage gain uses a multiplier of 20 (not 10) because power varies as the square of voltage; a 2× voltage increase is 6 dB, matching the 4× power increase.
  • Absolute signal levels: dBW, dBm, and dBV express power or voltage relative to a reference (1 W, 1 mW, 1 V), allowing direct addition/subtraction with gain values.
  • Common confusion: Gain is unitless (dB); signal levels carry units (dBW, dBV)—saying "gain is 35 dBW" is incorrect.
  • Bode plots: Graph gain (in dB) and phase versus frequency on log scales; break frequencies (corner frequencies) mark where gain drops 3 dB from midband, caused by circuit reactances.

🔢 Converting between ordinary and dB forms

🔢 Power gain formulas

Power gain in dB: G′ = 10 log₁₀ G
Ordinary power gain from dB: G = 10^(G′/10)

  • The multiplier is 10 for power.
  • Example: 23 dB gain → G = 10^(23/10) ≈ 199.5.
  • Negative dB values indicate loss (e.g., −3 dB means power is halved).

🔢 Voltage (and current) gain formulas

Voltage gain in dB: A′ᵥ = 20 log₁₀ Aᵥ
Ordinary voltage gain from dB: Aᵥ = 10^(A′ᵥ/20)

  • The multiplier is 20 (twice the power multiplier) because power ∝ voltage².
  • A 2× voltage increase = 6 dB (matching the 4× power increase = 6 dB).
  • A 10× voltage increase = 20 dB (twice the 10 dB for power).
  • Example: Input 50 mV, output 2 V → Aᵥ = 40 → A′ᵥ = 20 log₁₀(40) ≈ 32 dB.
  • Don't confuse: Voltage dB ≠ power dB unless input and output impedances are matched.

⚡ Approximation technique

  • Break dB values into 3 dB (≈ 2×) and 10 dB (= 10×) chunks.
  • Add the dB chunks, then multiply the corresponding factors.
  • Example: 23 dB = 3 dB + 10 dB + 10 dB → 2 × 10 × 10 = 200 (close to exact 199.5).
  • Faster than a calculator with practice; useful for quick estimates in multi-stage designs.

📊 Absolute signal levels: dBW, dBm, dBV

📊 Power references

P′ = 10 log₁₀ (P / reference)

UnitReferenceUse case
dBW1 wattGeneral power measurements
dBm1 milliwattCommunications industry (very common)
dBf1 femtowattVery low signals (e.g., antenna)
  • Positive dB values → power greater than reference; negative → less than reference.
  • Example: 200 mW = −7 dBW = 23 dBm (the two values differ by 30 dB because 1 W / 1 mW = 1000 = 30 dB).
  • Key insight: dBW and dBm for the same power always differ by exactly 30 dB.

📊 Voltage reference

V′ = 20 log₁₀ (V / reference)

  • Common reference: 1 V → units of dBV.
  • Uses multiplier of 20 (same reason as voltage gain).
  • Example: 2 V signal = 20 log₁₀(2) ≈ 6 dBV.

📊 Combining gains and signal levels

  • Output = Input + Gain (all in dB).
  • Example: Input −42 dBV, gain 35 dB → Output = −42 + 35 = −7 dBV.
  • Gain = Output − Input (all in dB).
  • Example: Input 20 dBm (= −10 dBW), output 25 dBW → Gain = 25 − (−10) = 35 dB.
  • Don't confuse: Final answer for signal level carries units (dBW, dBV); gain is unitless (dB only).

🔬 Laboratory considerations

🔬 Meter underflow

  • Digital meters on dBV scale cannot measure near zero volts (would be −∞ dBV).
  • Calculator shows "error" when trying to compute log₁₀(0).

🔬 dBm measurements with a voltmeter

  • Voltmeters can measure dBm if circuit impedance is known (commonly 600 Ω in audio/communications).
  • Power is derived from voltage via Power Law.
  • Warning: Readings are accurate only at the calibrated impedance; on other impedances, relative dB changes are still correct but absolute dBm values are wrong.

🔬 Sound pressure level (dB-SPL)

  • "110 dB" concert levels properly read as "110 dB-SPL."
  • Reference: 0 dB-SPL = threshold of hearing (20 micropascals for healthy young humans).
  • 1 dB ≈ just-noticeable difference in loudness (depends on frequency and pressure).

📈 Bode plots: gain and phase versus frequency

📈 What a Bode plot shows

A Bode plot is a pair of graphs: one for gain (in dB) versus frequency, one for phase (in degrees) versus frequency, both with logarithmic frequency axes.

  • Named after Hendrik Wade Bode (American engineer, control systems and telecommunications).
  • Used for circuit design and analysis to predict frequency response.

📈 Gain plot structure

  • Midband region: relatively flat; gain ≈ midband gain (can be fractional/negative dB in passive circuits).
  • Break frequencies (corner frequencies, half-power points):
    • f₁ = lower break frequency.
    • f₂ = upper break frequency.
    • At f₁ and f₂, gain is 3 dB less than midband gain.
  • Rolloff regions: gain decreases below f₁ and above f₂.
  • Amplifiers normally used only between f₁ and f₂.
  • Cause: Circuit reactances (coupling capacitors, stray capacitances).

📈 DC amplifiers and upper limits

  • Some amplifiers have no lower break (DC amplifiers, f₁ = 0).
  • All amplifiers have an upper break frequency f₂.

📈 Lead network (low-frequency rolloff)

A lead network is a high-pass filter (capacitor C in series with resistor R) that causes the low-frequency gain reduction.

  • Output voltage across R "leads" the input.
  • At very high frequencies: Xc ≪ R → circuit acts as resistive divider → Vₒᵤₜ ≈ Vᵢₙ.
  • As frequency drops: Xc increases → Vₒᵤₜ decreases.
  • Break frequency: fc = 1 / (2π R C) (standard RC formula).
  • Below fc, response is a straight line on log-frequency / dB-gain axes.
  • Slope: 6 dB per octave (octave = 2× frequency change) = 20 dB per decade (decade = 10× frequency change).
  • Asymptote approximation: Two straight-line segments (one flat above fc, one sloped below fc) closely approximate the actual curve.

📈 Why log scales help

  • Logarithmic frequency axis + dB gain axis → exponential rolloffs become straight lines.
  • Makes sketching and analyzing circuit response quick and convenient.

🧮 Multi-stage dB application

🧮 Cascaded stages

  • Total gain in dB = sum of individual stage gains (because dB is logarithmic).
  • Example: Three stages with 10 dB, 16 dB, 14 dB → Total = 10 + 16 + 14 = 40 dB.
  • Graphical representation: Input signal (e.g., 8 dBm) + stage 1 gain (10 dB) = 18 dBm → + stage 2 (−6 dB) = 12 dBm → + stage 3 (15 dB) = 27 dBm output.
  • Advantage: No need to convert to ordinary form; just add/subtract dB values at each stage.
72

Simplified AC Model of the JFET

11.2 Simplified AC Model of the JFET

🧭 Overview

🧠 One-sentence thesis

Amplifier frequency response is shaped by lead networks (which cause low-frequency rolloff) and lag networks (which cause high-frequency rolloff), and their combined effects determine the usable bandwidth and pulse-handling speed of the circuit.

📌 Key points (3–5)

  • Lead networks cause gain reduction at low frequencies with a +6 dB/octave slope below the break frequency, while lag networks cause gain reduction at high frequencies with a −6 dB/octave slope above the break frequency.
  • Break frequencies (critical frequencies) mark where the signal drops by 3 dB and are determined by the RC time constant: f_c = 1/(2π RC).
  • Common confusion: lead vs lag network behavior—lead networks pass high frequencies and block low (output leads input at low f), while lag networks pass low frequencies and block high (output lags input at high f); their gain curves are mirror images.
  • Asymptotic approximation allows quick sketching of frequency response using straight-line segments instead of exact curves, simplifying multi-stage analysis.
  • Rise time and bandwidth are inversely related: a circuit's upper break frequency determines how fast it can respond to pulse signals (f₂ ≈ 0.35/T_r).

📉 Lead networks and low-frequency response

🔌 What a lead network is

Lead network: a high-pass RC circuit where the output is taken across the resistor R, so named because the output voltage leads the input at low frequencies.

  • The generic lead network consists of a capacitor C in series with a resistor R, with output measured across R.
  • At very high frequencies, the capacitor's reactance X_c becomes negligible (X_c << R), so V_out ≈ V_in and the circuit acts like a simple wire.
  • At very low frequencies, X_c increases, forming a voltage divider that reduces V_out.
  • The break frequency occurs when X_c = R, calculated as f_c = 1/(2π RC).

📊 Gain response of lead networks

  • Below the break frequency f_c, gain drops in a predictable pattern.
  • The exact gain equation is: A'_v = −10 log₁₀(1 + f_c²/f²) in decibels.
  • The rolloff slope is 6 dB per octave (equivalently, 20 dB per decade).
    • An octave is a doubling or halving of frequency.
    • A decade is a factor of 10 in frequency.
  • Asymptotic approximation: the curve can be drawn as two straight lines meeting at f_c—flat at 0 dB above f_c, and sloping at +6 dB/octave below f_c.

Example: A circuit has a lower break frequency of 40 Hz. At 10 Hz (2 octaves below), the exact loss is −12.3 dB. Using the 6 dB/octave rule: 2 octaves × 6 dB = −12 dB, which closely matches the exact result.

🔄 Phase response of lead networks

  • At very low frequencies, the circuit is capacitive, so output leads input by 90°.
  • At very high frequencies, the circuit is resistive, so output is in phase with input (0°).
  • At the critical frequency f_c, the phase lead is 45°.
  • The exact phase equation is: θ = arctan(f_c/f).
  • The phase transition can be approximated as a straight line from 90° (at 0.1 f_c) to 0° (at 10 f_c), with error no more than 6° at the corners.

Example: For a lower break at 120 Hz, one decade below (12 Hz) gives θ = 84.3° (approaching 90°), and one decade above (1.2 kHz) gives θ = 5.71° (approaching 0°).

Don't confuse: DC-coupled amplifiers with no lead networks maintain 0° phase all the way down to DC (0 Hz).

📈 Lag networks and high-frequency response

🔌 What a lag network is

Lag network: a low-pass RC circuit where the output is taken across the capacitor C, essentially an inverted lead network, so named because the output voltage lags the input at high frequencies.

  • The lag network transposes the R and C positions compared to the lead network: output is measured across C.
  • At very low frequencies, X_c is very large, so V_out ≈ V_in (capacitor blocks DC but passes low AC).
  • At very high frequencies, X_c decreases, forming a voltage divider that reduces V_out.
  • The break point occurs when X_c = R, using the same f_c = 1/(2π RC) formula.
  • All amplifier systems contain lag networks (unlike lead networks, which can be eliminated in DC amplifiers).

📊 Gain response of lag networks

  • Above the break frequency f_c, gain drops with a −6 dB per octave slope (or −20 dB per decade).
  • The exact gain equation is: A'_v = −10 log₁₀(1 + f²/f_c²) in decibels.
  • Note the transposition: f and f_c are swapped compared to the lead network equation.
  • Asymptotic approximation: flat at 0 dB below f_c, then sloping downward at −6 dB/octave above f_c.

Example: A lag network with upper break at 150 kHz. At 1.6 MHz (slightly more than 1 decade above), the approximate loss is −20 dB; the exact calculation gives −20.6 dB.

🔄 Phase response of lag networks

  • At very low frequencies, the circuit is capacitive but output is across C, so V_out is in phase with V_in (0°).
  • At very high frequencies, the circuit is resistive, so output lags by 90°.
  • At the break frequency f_c, the phase lag is −45°.
  • The exact phase equation is: θ = −90° + arctan(f_c/f).
  • The negative slope distinguishes it from the lead network's positive phase slope.

Example: For the 150 kHz lag network at 1.6 MHz, the phase is −84.6° (approaching −90°).

🔗 Comparing lead and lag networks

FeatureLead networkLag network
Output taken acrossResistor RCapacitor C
Passes which frequenciesHigh frequenciesLow frequencies
Gain slope+6 dB/octave below f_c−6 dB/octave above f_c
Phase at low f+90° (leads)0° (in phase)
Phase at high f0° (in phase)−90° (lags)
Phase at f_c+45°−45°
Presence in circuitsOptional (can be eliminated in DC amps)Always present

Don't confuse: The gain curves are mirror images—lead networks boost relative gain at high frequencies (by cutting lows), while lag networks boost relative gain at low frequencies (by cutting highs).

⏱️ Rise time and bandwidth relationship

⏱️ What rise time measures

Rise time (T_r): the time it takes for a pulse signal to traverse from 10% to 90% of its peak value.

  • When a square pulse passes through a lag network, the capacitor charging effect rounds the edges.
  • This rounding limits the shortest pulse duration the circuit can handle without excessive distortion.
  • Rise time quantifies the "speed" of a circuit for pulse-type signals.

🧮 Deriving the rise time formula

  • The pulse shape follows the standard capacitor charge equation: V_out = V_peak(1 − e^(−t/RC)).
  • Solving for the time to reach 10% of peak: t₁ = 0.105 RC.
  • Solving for the time to reach 90% of peak: t₂ = 2.303 RC.
  • Rise time is the difference: T_r = t₂ − t₁ ≈ 2.2 RC.

🔗 Connecting rise time to bandwidth

  • The same RC values that determine rise time also set the upper critical frequency f₂ = 1/(2π RC).
  • Combining the two relationships yields: f₂ = 0.35 / T_r.
  • This inverse relationship means faster circuits (shorter rise time) require wider bandwidth (higher f₂).

Example: A lag network critical at 100 kHz has a rise time of T_r = 0.35 / 100 kHz = 3.5 μs.

Don't confuse: This formula applies to systems with a single clearly dominant lag network; multiple lag networks require more complex analysis.

🎨 Multi-stage effects and Bode plots

🎨 Combining multiple networks

  • A complete frequency response combines three elements: (1) midband response (flat gain), (2) lead network response (low-frequency rolloff), and (3) lag network response (high-frequency rolloff).
  • Most designs contain multiple lead and lag networks; the complete response is the summation of individual responses.
  • Dominant networks are those that affect the midband first:
    • Dominant lead network: the one with the highest f_c (cuts in first from below).
    • Dominant lag network: the one with the lowest f_c (cuts in first from above).

📐 Asymptotic sketching method

The excerpt describes a process for approximating system response:

  • Locate all critical frequencies f_c on the frequency axis.
  • Draw a straight line between the dominant lag and lead break points to represent the midband.
  • Add the rolloff slopes (±6 dB/octave) at each break frequency.
  • This straight-line approximation is much faster than calculating exact curves and sufficiently accurate for most design work.

📊 Bode plots

  • A Bode plot displays both gain and phase response simultaneously on separate graphs, both using a logarithmic frequency axis.
  • Examining both plots together allows quick determination of phase change for a given gain.
  • Example: At the critical frequency of a 150 kHz lag network, the gain is −3 dB and the phase is −45°, as expected for a single lag network.

Don't confuse: The asymptotic (straight-line) approximation is off by at most 3 dB at the break frequency itself, but becomes increasingly accurate away from f_c.

73

Common Source Amplifier

11.3 Common Source Amplifier

🧭 Overview

🧠 One-sentence thesis

Multi-stage amplifier frequency response is determined by combining midband gain with cumulative lead and lag network effects, where dominant networks set the primary bandwidth limits and additional networks increase rolloff slopes.

📌 Key points (3–5)

  • Rise time and bandwidth relationship: The upper critical frequency f₂ and rise time Tr are inversely related by the formula f₂ = 0.35/Tr.
  • Dominant networks define bandwidth: The highest-frequency lead network and lowest-frequency lag network set the primary breakpoints for the midband response.
  • Cumulative rolloff slopes: Each additional lead or lag network adds 6 dB per octave to the rolloff rate (e.g., two lag networks produce −12 dB/octave).
  • Common confusion—narrowing effect: When multiple networks share similar critical frequencies, the true −3 dB breakpoint shifts because losses accumulate (e.g., two networks both at 1 MHz each contribute 3 dB loss, totaling 6 dB at that frequency).
  • Decibel advantages: The dB scheme converts multiplication/division into addition/subtraction, simplifying multi-stage gain calculations and frequency response analysis.

⏱️ Rise time and critical frequency

⏱️ Deriving the rise time formula

Rise time Tr: the time for a pulse to transition from 10% to 90% of its peak value.

  • For a lag network (RC circuit), the rise time is approximately Tr ≈ 2.2 RC.
  • The calculation uses exponential charging equations evaluated at 0.1V peak (t₁ = 0.105 RC) and 0.9V peak (t₂ = 2.303 RC).
  • The difference yields Tr = t₁ − t₂ ≈ 2.2 RC.

🔗 Linking rise time to bandwidth

The upper critical frequency f₂ relates to the RC time constant by:

  • f₂ = 1/(2π RC)
  • Substituting RC = 2.2/Tr gives f₂ = 0.35/Tr.

Example: A lag network critical at 100 kHz has a rise time of Tr = 0.35/100 kHz = 3.5 μs.

Why it matters: This relationship allows designers to predict bandwidth from pulse response measurements, or vice versa.

📊 Multi-stage frequency response construction

📊 Identifying dominant networks

Dominant lead network: the lead network with the highest critical frequency fc.

Dominant lag network: the lag network with the lowest critical frequency fc.

  • These networks affect the midband response first as frequency moves away from the center.
  • Non-dominant networks contribute additional rolloff at their respective critical frequencies.

🛠️ Step-by-step Bode plot construction

  1. Locate all fc values on the frequency axis.
  2. Draw midband gain line: a horizontal line between the dominant lag fc and dominant lead fc at the midband gain level. If no lead networks exist, extend this line down to DC.
  3. Add lead network slopes: Draw +6 dB/octave slope below the dominant lead fc. Each additional lead network adds another +6 dB/octave (cumulative: 12 dB/octave for two networks, 18 dB/octave for three, etc.).
  4. Add lag network slopes: Draw −6 dB/octave slope above the dominant lag fc. Each additional lag network adds another −6 dB/octave (cumulative: −12 dB/octave for two networks, −18 dB/octave for three, etc.).

Example: An amplifier with midband gain = 26 dB, one lead network at 200 Hz, and lag networks at 10 kHz and 30 kHz:

  • Midband line runs from 200 Hz to 10 kHz at 26 dB.
  • Below 200 Hz: +6 dB/octave slope (at 100 Hz, gain = 20 dB).
  • From 10 kHz to 30 kHz: −6 dB/octave slope (at 20 kHz, gain ≈ 20 dB).
  • Above 30 kHz: −12 dB/octave slope (at 60 kHz, gain ≈ 4 dB).

⚠️ The narrowing effect

When two or more networks share similar critical frequencies and one is dominant:

  • The true −3 dB breakpoint shifts from the nominal fc.
  • Don't confuse the Bode plot approximation with the actual response: the straight-line Bode plot shows only approximate shape.

Example: Two lag networks both critical at 1 MHz each produce 3 dB loss at 1 MHz, totaling 6 dB loss. The true −3 dB point moves to a lower frequency than 1 MHz.

🔢 Decibel measurement fundamentals

🔢 Why use decibels

  • Simplifies calculations: Multiplication and division in ordinary form become addition and subtraction in dB form.
  • Standard for Bode plots: Gain magnitude is measured in dB; frequency axis is logarithmic.
  • Convenient for multi-stage systems: Total gain = sum of individual stage gains in dB.

📐 Power vs voltage dB conversions

QuantityConversion to dBNotes
Power gainUse power ratio formuladB power gain
Voltage gainUse voltage ratio formuladB voltage gain (different formula than power)
  • Power and voltage gain calculations differ in their formulas.
  • The excerpt emphasizes this distinction but does not provide the explicit formulas.

🏷️ Decibel reference units

The excerpt mentions several dB reference standards:

  • dB: relative gain (no absolute reference).
  • dBW: power referenced to 1 watt.
  • dBm: power referenced to 1 milliwatt.
  • dBV: voltage referenced to 1 volt.
  • dBu: voltage reference (specific value not detailed in excerpt).

🌊 Lead and lag network behavior

🌊 Lead network characteristics

Lead network: a circuit that causes low-frequency gain to roll off.

  • Rolloff rate: +6 dB per octave per network.
  • Phase change: from +90° to 0° per network.
  • The term f₁ indicates the lower critical frequency (lead network breakpoint).

🌊 Lag network characteristics

Lag network: a circuit that causes high-frequency gain to roll off.

  • Rolloff rate: −6 dB per octave per network.
  • Phase change: from 0° to −90° per network.
  • The term f₂ indicates the upper critical frequency (lag network breakpoint).

🔄 Cumulative effects

  • Multiple lead or lag networks add their effects: slopes increase by 6 dB/octave for each additional network.
  • Phase shifts also accumulate: two lag networks can shift phase from 0° to −180° total.

Example: An amplifier with three lag networks has a maximum high-frequency rolloff of −18 dB/octave (3 × 6 dB/octave).

74

Low Frequency Response of BJT Amplifiers

11.4 Common Drain Amplifier

🧭 Overview

🧠 One-sentence thesis

The low frequency performance of BJT amplifiers is determined by lead networks formed by coupling and bypass capacitors, with the highest critical frequency among these networks establishing the lower frequency limit (f₁) of the amplifier.

📌 Key points (3–5)

  • What creates low frequency limits: coupling capacitors (input/output) and bypass capacitors form lead networks that attenuate signals at low frequencies.
  • How to find f₁: calculate the critical frequency for each lead network (input, output, bypass), then identify the highest frequency as the dominant one that sets f₁.
  • Bypass network is special: unlike true lead networks, the bypass network creates a "stepped" response that levels off at very low frequencies rather than rolling off continuously.
  • Common confusion: the bypass capacitor doesn't create a pure lead network—it partially bypasses the emitter resistor as frequency drops, causing gain to stabilize at a lower level rather than continuing to fall.
  • Design insight: removing swamping resistors increases the required capacitance values for the same critical frequency, with dramatic effects on the bypass network.

🔍 Lead networks fundamentals

🔍 What is a lead network

A lead network exists wherever signal current flows through a capacitor on its way to the load.

  • The capacitor creates a frequency-dependent voltage divider with the associated impedance.
  • As frequency decreases, capacitive reactance increases, worsening the voltage divider effect.
  • Result: progressively less signal reaches the load as frequency drops—the hallmark of a lead network.

📍 Where lead networks appear in CE amplifiers

A typical single-stage common emitter amplifier has three lead networks:

LocationCapacitorFunction
InputInput coupling capacitor (C_in)Couples signal from source to base
OutputOutput coupling capacitor (C_out)Couples signal from collector to load
EmitterBypass capacitor (C_E)Bypasses emitter bias resistor
  • Multi-stage systems: output network of one stage combines with input network of the next to form a single combined network.

🧮 Calculating input and output critical frequencies

🧮 Input network analysis

To find the input critical frequency:

  1. Identify the capacitance: just C_in (the input coupling capacitor).
  2. Find the Thevenin resistance seen from the capacitor, called R_in(lead).

R_in(lead) = R_gen + R_B || β(R_SW + r'_e)

Where:

  • R_gen = internal resistance of signal source
  • R_B = base biasing resistor
  • β(R_SW + r'_e) = input impedance looking into the base
  • R_SW = swamping resistor (set to zero if not swamped)
  • r'_e = AC emitter resistance

Critical frequency formula: f_c(input lead) = 1 / (2π × R_in(lead) × C_in)

🧮 Output network analysis

The output network is simpler:

  • Find Thevenin resistance by opening the current source.
  • Resistance is the series combination: R_out(lead) = R_C + R_L

Critical frequency formula: f_c(output lead) = 1 / (2π × R_out(lead) × C_out)

⚡ Bypass network—the special case

⚡ Why bypass networks differ

The bypass network is not a true lead network because its response doesn't roll off continuously as frequency decreases—it eventually levels off.

How it works:

  • At midband: capacitor reactance is much smaller than emitter bias resistor R_E, effectively shorting it out.
  • As frequency drops: reactance increases, only partially bypassing the resistor.
  • Voltage gain starts to decrease (like increasing swamping).
  • At very low frequencies: reactance equals R_E value; further frequency drops have little effect.
  • Result: gain stabilizes at a lower level, creating a "stepped" response.

⚡ Bypass network formulas

Resistance calculation: R_bypass(lead) = R_E || (R_SW + r'_e + (R_B || R_gen)/β)

  • The swamping resistor R_SW tends to dominate since the final term is usually small.

Critical frequency: f_c(bypass lead) = 1 / (2π × R_bypass(lead) × C_E)

Step frequency (where response flattens): f_c(bypass step) = 1 / (2π × R_E × C_E)

  • Quick approximation: step frequency ≈ bypass critical frequency ÷ (R_E / R_SW)
  • Example: if R_E = 5 kΩ and R_SW = 250 Ω, step frequency is roughly 20 times lower than bypass critical frequency.
  • Usually low enough to ignore in overall amplifier response.

🎯 Determining system f₁

🎯 Finding the dominant frequency

Once all lead frequencies are computed:

  • The highest critical frequency is the dominant frequency—where midband response is first affected.
  • This dominant frequency approximately equals the system f₁.

🎯 Interaction effects

Don't confuse: if two or more frequencies are close together and one is dominant, interaction between them may cause the actual system f₁ to be somewhat higher than the dominant f_c alone.

🎯 Effect of removing swamping

If the amplifier is not swamped (no R_SW):

  • Set R_SW = 0 in the formulas for input and bypass networks.
  • This decreases the effective resistance in these networks.
  • Result: higher required capacitance values for a given critical frequency.
  • Effect is dramatic for the bypass network.

📊 Practical example walkthrough

📊 Example scenario

Given a common emitter amplifier with β = 165 and DC emitter current ≈ 2 mA (setting r'_e to 13 Ω):

Input network calculation:

  • R_in(lead) = 50 Ω + 20 kΩ || 165(100 Ω + 13 Ω) = 9.7 kΩ
  • f_c(input lead) = 1 / (2π × 9.7 kΩ × 4.7 μF) = 3.5 Hz

Output network calculation:

  • R_out(lead) = R_C + R_L = 15 kΩ
  • f_c(output lead) = 1 / (2π × 15 kΩ × 20 μF) = 0.53 Hz

Bypass network calculation:

  • R_bypass(lead) = 4.5 kΩ || (100 Ω + 13 Ω + (20 kΩ || 50 Ω)/165) = 113 Ω
  • f_c(bypass lead) = 1 / (2π × 113 Ω × 33 μF) = 42.7 Hz

📊 Result interpretation

  • Dominant network: bypass at 42.7 Hz (highest of the three).
  • Other two networks are well below this point.
  • Therefore, f₁ ≈ 42.7 Hz.
  • Computer simulation verified this result at approximately 42.8 Hz with midband gain just below 29 dB.
75

Common Gate Amplifier

11.5 Common Gate Amplifier

🧭 Overview

🧠 One-sentence thesis

The common gate amplifier configuration has only two lead networks (input and output), making its low-frequency analysis simpler than the common emitter amplifier, with the input impedance found by looking into the emitter terminal and the output network analyzed identically to the common emitter case.

📌 Key points (3–5)

  • Two lead networks only: Unlike common emitter amplifiers with three networks, common gate has only input and output lead networks, limiting rolloff to 12 dB per octave maximum.
  • Input impedance calculation: The signal source drives the emitter, so input impedance is found similarly to looking toward the emitter of an emitter follower—essentially the emitter bias resistor in parallel with the impedance looking into the emitter.
  • Output network unchanged: The output network analysis is virtually identical to the common emitter configuration, using the same equations for equivalent resistance and critical frequency.
  • Typical approximation: With typical values, the input resistance can be approximated as r'e plus R_gen, which is relatively small and requires large capacitance for low critical frequencies.

🔌 Circuit topology and network count

🔌 Two networks vs three

  • The common gate amplifier has only two lead networks: one at the input and one at the output.
  • This contrasts with the common emitter amplifier, which has three networks (input, output, and bypass).
  • Consequence: The rolloff rate will be no steeper than 12 dB per octave (40 dB per decade), unlike the common emitter which can have steeper rolloff due to the additional bypass network.

🔍 Why fewer networks matter

  • Fewer networks mean simpler analysis—only two critical frequencies to calculate.
  • The dominant network (highest critical frequency) determines f₁ directly without needing to consider a third bypass network.
  • Example: If input network has a critical frequency of 5 Hz and output has 2 Hz, f₁ is approximately 5 Hz with less interaction complexity than three-network configurations.

🎯 Input network analysis

🎯 Signal path and impedance

  • The signal source drives the emitter in the common gate configuration, unlike common emitter where it drives the base.
  • Input impedance is found in essentially the same manner as finding the impedance looking toward the emitter of an emitter follower.

🧮 Calculating input resistance

The input impedance equals:

  • The emitter bias resistor R_E in parallel with Z_in(emitter)
  • Since the base terminal is at AC ground, Z_in(emitter) equals r'e
  • The signal source's internal impedance R_gen is added to the total

Practical approximation: With typical values, this can be simplified to r'e plus R_gen.

⚠️ Design implication

  • The input resistance is relatively small (dominated by r'e, which is typically tens of ohms).
  • To achieve a low critical frequency, this requires a large capacitance at the input.
  • Don't confuse: This is fundamentally different from common emitter, where the input impedance is much higher due to beta multiplication.

📤 Output network analysis

📤 Identical to common emitter

The output network is virtually identical to that of the common emitter (essentially a current source feeding two resistors that are separated by an in-line capacitor).

  • The equations for equivalent resistance and critical frequency are unchanged from the common emitter case.
  • The transistor is modeled as a controlled current source.
  • The output coupling capacitor separates the drain/collector resistor from the load resistor.

🧮 Output resistance and frequency

  • The effective output resistance is the sum of the collector/drain biasing resistor and the load resistor.
  • The critical frequency is calculated using the standard formula: f_c = 1 / (2π × R_out × C_out).
  • This network typically has a lower critical frequency than the input network due to larger resistance values and capacitance.
76

Multi-Stage and Combination Circuits

11.6 Multi-Stage and Combination Circuits

🧭 Overview

🧠 One-sentence thesis

High-frequency amplifier response is limited by parasitic capacitances within transistors and wiring that create lag networks, with the dominant critical frequency determined by whichever input or output network produces the lower cutoff.

📌 Key points (3–5)

  • What limits high frequency: parasitic capacitances (device internal, PCB traces, cabling) create lag networks that reduce signal at high frequencies; all amplifiers have an upper frequency limit f₂.
  • Two lag networks per stage: one at the input and one at the output; for multi-stage amplifiers, adjacent stages combine into interstage lag networks.
  • Miller effect magnifies input capacitance: the collector-base capacitance C_cb is multiplied by (|A_v| + 1) at the input, often making it the dominant capacitance.
  • Common confusion: manufacturer data sheets list C_ibo and C_obo instead of C_be and C_cb directly; these must be converted, and their values vary with bias voltage.
  • Design implication: unlike low-frequency response (which can be eliminated by larger coupling capacitors), high-frequency limits cannot be avoided—only managed.

🔽 Low Frequency Lead Networks (FET Example)

🔽 What lead networks do

Lead networks: capacitive coupling and bypass networks that limit low-frequency response by blocking DC while passing AC signals.

  • At low frequencies, coupling and bypass capacitors have high reactance, reducing signal transmission.
  • Each network has a critical frequency f_c = 1 / (2π R C) where the response drops by 3 dB.
  • The dominant network (highest f_c) determines the overall lower break frequency f₁.

🧮 FET lead network analysis

The excerpt walks through a FET amplifier with three lead networks:

NetworkEffective ResistanceCapacitanceCritical Frequency
InputGenerator + gate bias (≈5 MΩ)47 nF0.677 Hz
OutputDrain bias ‖ load (18 kΩ)1 μF8.84 Hz
BypassSource bias ‖ (1/g_m) (429 Ω)220 μF1.69 Hz
  • The output network at 8.84 Hz is dominant, so f₁ ≈ 8.84 Hz.
  • Interaction effect: when other networks are close in frequency, they push the true f₁ slightly higher.

🎯 Design approach for target frequency

Example: to achieve f₁ = 20 Hz in a source follower:

  • Calculate effective resistance for each network.
  • Solve for capacitance: C = 1 / (2π R f_c).
  • Key constraint: cannot make both networks critical at 20 Hz (would create 6 dB loss, not 3 dB).
  • Solution: size one capacitor for 20 Hz, then increase the other (usually the smaller one) by ~10× to push its f_c well below 20 Hz, eliminating interaction.

🔌 Configuration-specific formulas

Common drain (source follower):

  • Input: R_in(lead) = R_gen + R_G
  • Output: R_bypass(lead) = R_L + R_S ‖ (1/g_m)

Common gate:

  • Input: R_input(lead) = R_gen + R_S ‖ (1/g_m)
  • Output: same as common source (R_D ‖ R_L)

Don't confuse: the impedance looking into the source terminal is R_S ‖ (1/g_m), not just R_S.

🔼 High Frequency Lag Networks

🔼 What causes high-frequency limits

Lag networks: frequency-dependent voltage dividers with resistance in-line and capacitance across the load; as frequency increases, capacitive reactance decreases, reducing output voltage.

  • Universal limit: all amplifiers have an upper frequency limit f₂, unlike low-frequency response which can be designed out.
  • Caused by parasitic capacitances you cannot see:
    • Device internal capacitances (few picofarads to nanofarads).
    • PCB traces (few picofarads or less per trace).
    • Cabling (tens of picofarads per foot, ~100 pF/meter).
  • All stray capacitances appear in parallel and are lumped as C_stray.

🧩 Two lag networks per stage

  • Input lag network: at the amplifier input.
  • Output lag network: at the amplifier output.
  • Multi-stage: the output of one stage combines with the input of the next to form a single interstage lag network.

🔬 BJT High-Frequency Analysis

🔬 Parasitic capacitances in BJTs

BJTs have small parasitic capacitances between terminals:

  • C_be: base-emitter capacitance (single-digit picofarads).
  • C_cb: collector-base capacitance (single-digit picofarads).
  • C_ce: collector-emitter capacitance (usually small enough to ignore).

These are much smaller than coupling/bypass capacitors but dominate at high frequencies.

🪞 Miller effect at the input

C_cb sits between input and output (the Miller position) in inverting amplifiers like common emitter.

Miller's Theorem: a capacitor straddling an inverting amplifier can be split into two equivalent capacitors—one at input, one at output—with the same overall effect.

Input side:

  • C_in(miller) = C_cb × (|A_v| + 1)
  • This multiplication often makes C_in(miller) the largest capacitance in the input network by far.

Output side:

  • C_out(miller) = C_cb × (|A_v| + 1) / |A_v|

Don't confuse: Miller's Theorem applies only to inverting amplifiers (common emitter, common source), not to non-inverting configurations (common collector, common drain).

🧮 Input lag network equations

Total input capacitance:

  • C_input(lag) = C_be + C_in(miller) + C_in(stray)

Effective input resistance:

  • R_input(lag) = R_gen ‖ R_B ‖ Z_in(base)

Critical frequency:

  • f_c(input lag) = 1 / (2π R_input(lag) C_input(lag))

🧮 Output lag network equations

Total output capacitance:

  • C_output(lag) = C_out(load) + C_out(miller) + C_out(stray)

Effective output resistance:

  • R_output(lag) = R_C ‖ R_L

Critical frequency:

  • f_c(output lag) = 1 / (2π R_output(lag) C_output(lag))

Dominant frequency: the lower of the two critical frequencies (input or output) determines the system f₂.

📊 Practical Measurement and Data Sheet Issues

📊 Data sheet parameters

Manufacturers do not list C_cb and C_be directly; instead they list:

  • C_ibo: "input in common base configuration with an open output."
  • C_obo: "output in common base configuration with an open input."

Conversion (approximations):

  • C_be ≈ C_ibo
  • C_cb ≈ C_obo

📈 Voltage dependence of capacitances

These parasitic capacitances are not fixed values; they depend on bias voltages.

Example: C_cb (reverse-biased collector-base junction):

  • The depletion region acts as the capacitor dielectric.
  • Increasing reverse bias widens the depletion region (like increasing plate separation).
  • Result: capacitance decreases with increasing reverse voltage.

From the 2N3904 example curve:

  • C_obo ≈ 3.5 pF at very low reverse voltages.
  • C_obo < 2 pF above 5 volts.

Best practice: perform DC bias calculation to find the reverse voltage, then use manufacturer curves (not just the zero-volt maximum values) for accurate capacitance.

⚠️ Additional considerations

  • For amplifiers driven by very low impedance sources (low R_gen), use more accurate transistor models (the excerpt hints at this but does not complete the thought).
  • Data sheets may use different names for these parameters—be alert when reading specifications.
77

High Frequency Response of BJT and FET Amplifiers

11.7 Ohmic Region Operation

🧭 Overview

🧠 One-sentence thesis

The high-frequency response of BJT and FET amplifiers is limited by parasitic device capacitances that form input and output lag networks, with the Miller effect amplifying the impact of bridging capacitances in inverting configurations.

📌 Key points (3–5)

  • Parasitic capacitances control high-frequency rolloff: BJT and FET internal capacitances (C_be, C_cb for BJT; C_gs, C_gd for FET) create lag networks at input and output that limit bandwidth.
  • Miller effect multiplies bridging capacitance: the capacitance between input and output terminals (C_cb or C_gd) appears multiplied by (|A_v| + 1) at the input, often dominating the input lag capacitance.
  • Capacitance varies with bias voltage: parasitic capacitances decrease as reverse bias increases (wider depletion region), so manufacturers provide curves rather than fixed values.
  • Common confusion—inverting vs non-inverting: Miller's Theorem applies only to inverting amplifiers (common emitter, common source); non-inverting configurations (common collector, common base, common drain, common gate) do not exhibit the Miller multiplication effect.
  • System f_2 is determined by the dominant lag: the lower of the input and output critical frequencies sets the amplifier's upper cutoff frequency.

🔬 BJT parasitic capacitances and measurement

🔬 Device capacitances

  • BJT exhibits three parasitic capacitances:
    • C_be: base-emitter capacitance
    • C_cb: collector-base capacitance
    • C_ce: collector-emitter capacitance (usually very small)
  • These arise from reverse-biased junctions acting as capacitor plates with depletion regions as dielectric.

📏 Measurement conventions

Manufacturers measure capacitances in specific configurations:

C_ibo: "input in common base configuration with an open output"
C_obo: "output in common base configuration with an open input"

  • By observation: C_ibo ≈ C_be (because C_ce is very small)
  • Similarly: C_obo ≈ C_cb
  • Data sheets provide curves showing how these capacitances vary with reverse voltage.

⚡ Voltage-dependent behavior

  • As reverse bias potential increases, the depletion region widens (like increasing capacitor plate separation).
  • Result: capacitance decreases with increasing reverse voltage.
  • Example: C_obo for the 2N3904 is about 3.5 picofarads at very low reverse voltages but decreases to less than 2 picofarads above 5 volts.
  • For highest accuracy: perform DC bias calculation to determine reverse voltage, then use manufacturer's graph to obtain capacitance value.

🎯 Miller effect in BJT amplifiers

🎯 What Miller's Theorem does

  • The bridging capacitance C_cb sits between input (base) and output (collector) in common emitter amplifiers.
  • Miller's Theorem splits this single capacitance into two equivalent capacitances:
    • Input Miller capacitance: C_in(miller) = C_cb × (|A_v| + 1)
    • Output Miller capacitance: C_out(miller) = C_cb × (|A_v| + 1) / |A_v|
  • The input capacitance is multiplied by the gain magnitude plus one, often making it much larger than the physical device capacitance.

🔍 Why Miller dominates

  • In high-gain amplifiers, the multiplication factor (|A_v| + 1) can be hundreds.
  • Example from the excerpt: with A_v = -288, a 3 pF C_cb becomes 867 pF at the input.
  • This Miller input capacitance typically dwarfs C_be (5 pF in the example), so the input lag is dominated by the Miller effect.
  • Don't confuse: the output Miller capacitance is approximately equal to the original C_cb because (|A_v| + 1) / |A_v| ≈ 1 for large gains.

🧮 BJT input and output lag networks

🧮 Input lag network

The input network consists of:

  • Resistance: R_input(lag) = R_gen || R_B || Z_in(base), where Z_in(base) = β × r'_e
  • Capacitance: C_input(lag) = C_be + C_in(miller) + C_in(stray)
  • Critical frequency: f_c(input lag) = 1 / (2π × R_input(lag) × C_input(lag))

Example from the excerpt:

  • With R_input(lag) = 95.7 Ω and C_input(lag) = 872 pF (867 pF Miller + 5 pF C_be), the input lag f_c = 1.91 MHz.
  • The Miller input capacitance (867 pF) dominates over C_be (5 pF).

🧮 Output lag network

The output network consists of:

  • Resistance: R_output(lag) = R_C || R_L
  • Capacitance: C_output(lag) = C_out(miller) + C_out(stray) + C_out(load)
  • Critical frequency: f_c(output lag) = 1 / (2π × R_output(lag) × C_output(lag))

Example from the excerpt:

  • With R_output(lag) = 7.5 kΩ and C_output(lag) = 3 pF, the output lag f_c = 7.07 MHz.
  • Even though capacitance is much lower than at the input, the increased resistance places the critical frequency in the same neighborhood.

🎚️ System f_2 determination

  • The system's upper cutoff frequency (f_2) is approximately equal to the lower of the two critical frequencies.
  • In the example: input lag (1.91 MHz) is lower than output lag (7.07 MHz), so f_2 ≈ 1.91 MHz.
  • Caveat: when the two critical frequencies are relatively close, the true "3 dB down frequency" will be somewhat lower because the output network contributes non-zero loss at the input lag frequency.

🔧 Special cases and refinements for BJT

🔧 Base spreading resistance

  • For amplifiers driven by particularly low impedance sources (low R_gen), include the base spreading resistance r'_b.
  • This value appears in series with R_gen, so add them together before placing the sum in parallel with other input resistors.

🔧 Swamped amplifiers

  • In swamped amplifiers (with emitter resistor R_SW not bypassed), C_in(miller) is not in parallel with C_be due to R_SW.
  • In practice, C_in(miller) is usually much larger than C_be, so placing them in parallel generally yields acceptable results.

🔧 Non-inverting configurations

  • Common collector followers and common base amplifiers are non-inverting, so Miller's Theorem does not apply.
  • Common collector: C_bc winds up in parallel with C_be across the input; C_ce appears across the output.
  • Common base: C_be is across the input; C_cb is across the output; the bridged C_ce is usually small enough to ignore.

🔌 FET parasitic capacitances

🔌 Device capacitances

FETs exhibit three parasitic capacitances:

  • C_gs: gate-source capacitance
  • C_gd: gate-drain capacitance (in the Miller position for common source amplifiers)
  • C_ds: drain-source capacitance (tends to be small enough to ignore in most amplifiers)

For typical small signal devices, these are usually in the single-digit picofarad range; for large signal power devices, C_gs can be in the nanofarad range.

📏 Measurement conventions for FETs

Manufacturers often list C_iss and C_rss instead of idealized device capacitances:

C_rss: "reverse in common source configuration with shorted input"
C_iss: "input in common source configuration with shorted output"

Translation formulas:

  • C_gd ≈ C_rss (because C_ds is small enough to ignore)
  • C_gs ≈ C_iss - C_rss

⚡ Voltage-dependent behavior in FETs

  • Like BJTs, FET parasitic capacitances vary with applied reverse bias.
  • Manufacturers provide graphical data showing capacitance vs. voltage.
  • Example from the excerpt: C_gs is just over 10 picofarads near zero volts but drops to 4 picofarads at 10 volts and just over 3 picofarads at 30 volts.

🎯 Miller effect in FET amplifiers

🎯 Miller multiplication in common source

  • For common source amplifiers, C_gd is in the Miller position (between gate and drain).
  • Miller formulas are identical to BJT case:
    • C_in(miller) = C_gd × (|A_v| + 1)
    • C_out(miller) = C_gd × (|A_v| + 1) / |A_v|
  • C_gs appears at the input in parallel with C_in(miller).

🔍 Example calculation

From the excerpt example with A_v = -5:

  • C_gd = 1.5 pF becomes C_in(miller) = 9 pF at the input (multiplied by 6).
  • C_out(miller) = 1.8 pF at the output.
  • Total input capacitance = C_gs (3 pF) + C_in(miller) (9 pF) = 12 pF.

🧮 FET input and output lag networks

🧮 Input lag network for FETs

  • Resistance: R_input(lag) = R_gen || R_G || Z_in(gate)
  • Given that Z_in(gate) and R_G are often very large compared to R_gen, in many circuits: R_input(lag) ≈ R_gen.
  • Capacitance: C_input(lag) = C_gs + C_in(miller) + C_in(stray)
  • Critical frequency: f_c(input lag) = 1 / (2π × R_input(lag) × C_input(lag))

Example from the excerpt:

  • R_input(lag) = 50 Ω (R_gen dominates because R_G = 2 MΩ is very large)
  • C_input(lag) = 12 pF (3 pF C_gs + 9 pF Miller)
  • f_c(input lag) = 265 MHz

🧮 Output lag network for FETs

  • Resistance: R_output(lag) = R_D || R_L
  • Capacitance: C_output(lag) = C_out(miller) + C_out(stray) + C_out(load)
  • Critical frequency: f_c(output lag) = 1 / (2π × R_output(lag) × C_output(lag))

Example from the excerpt:

  • R_output(lag) = 4 kΩ (6 kΩ || 12 kΩ)
  • C_output(lag) = 1.8 pF
  • f_c(output lag) = 22.1 MHz (calculated from the given values)

🔧 Special cases for FETs

🔧 Swamped FET amplifiers

  • For swamped amplifiers, C_gs is no longer in parallel with C_in(miller) and C_in(stray).
  • In many circuits, treating them as though they are in parallel will yield acceptable accuracy.
  • This is more likely to be a problem when:
    • Gains are particularly low, or
    • C_gs is much larger than C_gd.

🔧 Non-inverting FET configurations

  • Common gate and common drain followers are non-inverting, so Miller's Theorem does not apply.
  • The issues for FETs are similar to those of BJT common base and common collector followers.
  • Don't confuse: only common source (inverting) exhibits Miller multiplication; common drain and common gate do not.

📊 Comparison of BJT and FET high-frequency behavior

AspectBJTFET
Parasitic capacitancesC_be, C_cb, C_ceC_gs, C_gd, C_ds
Miller position capacitanceC_cb (common emitter)C_gd (common source)
Typical capacitance rangeSingle-digit picofaradsSingle-digit pF (small signal); nanofarads (power devices)
Input impedance effectZ_in(base) = β × r'_e (moderate)Z_in(gate) very large (often ignored)
Measurement conventionsC_ibo, C_oboC_iss, C_rss
Non-inverting configsCommon collector, common baseCommon drain, common gate

Both BJT and FET amplifiers:

  • Exhibit voltage-dependent parasitic capacitances (decrease with increasing reverse bias).
  • Use Miller's Theorem only for inverting configurations.
  • Have input and output lag networks that determine system f_2.
  • Require manufacturer curves for accurate high-frequency analysis.
78

High Frequency Response Analysis of FET Amplifiers

Chapter 12: Metal Oxide Semiconductor FETs (MOSFETs)

🧭 Overview

🧠 One-sentence thesis

The high frequency response of FET amplifiers is limited by lag networks formed by device and parasitic capacitances, with Miller's theorem enabling the conversion of bridging capacitances into equivalent input and output capacitances that determine the upper frequency limit f₂.

📌 Key points (3–5)

  • Device capacitances must be translated: Gate-to-source capacitance (Cgs) equals Ciss minus Crss, while gate-to-drain capacitance (Cgd) equals Crss.
  • Miller effect multiplies capacitance: A capacitor bridging input to output appears larger at the input (multiplied by gain magnitude plus one) and slightly larger at the output.
  • Dominant network sets f₂: The output and input lag networks each have a critical frequency; the lower of the two determines the system's upper frequency limit.
  • Common confusion: The same physical capacitor (Cgd) affects both input and output networks differently due to Miller multiplication—don't treat it as a simple parallel capacitance.
  • Practical adjustment: Lowering f₂ is easy (add capacitance, especially in Miller position), but raising f₂ requires reducing capacitances or resistances, possibly needing circuit redesign.

🔄 Device capacitance translation

📊 Converting datasheet values

The excerpt shows that MOSFET/JFET datasheets provide:

  • Ciss: input capacitance (gate-to-source plus gate-to-drain)
  • Crss: reverse transfer capacitance (gate-to-drain)

Cgd = Crss

Cgs = Ciss − Crss

Example from the excerpt: Given Crss = 1.5 pF and Ciss = 4.5 pF, then Cgd = 1.5 pF and Cgs = 3 pF.

🎯 Why this matters

  • These translated values feed directly into Miller calculations.
  • The gate-to-drain capacitance bridges the amplifier and must be handled specially.

⚡ Miller's theorem application

🔍 Miller input capacitance

The capacitance from gate to drain appears amplified at the input side:

Cin(miller) = Cgd × (|Av| + 1)

Where Av is the voltage gain (negative for inverting amplifiers).

Example: With Cgd = 1.5 pF and gain magnitude of 5, the Miller input capacitance becomes 1.5 pF × (5 + 1) = 9 pF.

Why it's multiplied: The voltage across Cgd changes more than the input voltage alone because the output swings in the opposite direction, creating a larger effective capacitance.

🔍 Miller output capacitance

The same bridging capacitance appears slightly modified at the output:

Cout(miller) = Cgd × (|Av| + 1) / |Av|

Example: With the same values, output Miller capacitance = 1.5 pF × 6 / 5 = 1.8 pF.

Don't confuse: The input sees a much larger capacitance than the output from the same physical Cgd because the input experiences the full Miller multiplication effect.

🧮 Calculating lag network frequencies

🔌 Input lag network

Steps to find the input critical frequency:

  1. Find input resistance: Parallel combination of generator resistance, gate bias resistor, and gate input impedance (often infinite, so ignored)

    • Example: Rgen = 50 Ω, RG = 2 MΩ, Zin(gate) = ∞ → Rinput(lag) = 50 Ω
  2. Find total input capacitance: Sum of Cgs and Miller input capacitance

    • Example: Cgs = 3 pF, Cin(miller) = 9 pF → Cinput(lag) = 12 pF
  3. Calculate critical frequency: fc(input lag) = 1 / (2π × Rinput(lag) × Cinput(lag))

    • Example result: 265 MHz

🔌 Output lag network

Similar process for the output side:

  1. Find output resistance: Parallel combination of drain resistor and load resistor

    • Example: RD = 6 kΩ, RL = 12 kΩ → Routput(lag) = 4 kΩ
  2. Find total output capacitance: Usually just the Miller output capacitance (ignoring strays)

    • Example: Cout(miller) = 1.8 pF
  3. Calculate critical frequency: fc(output lag) = 1 / (2π × Routput(lag) × Coutput(lag))

    • Example result: 22.1 MHz

🎯 Determining system f₂

The system f₂ is set by the dominant (lower frequency) network.

In the example, output network at 22.1 MHz is dominant because it's lower than input network at 265 MHz. When networks are over a decade apart, interaction is minimal.

🛠️ Modifying frequency response

📉 Lowering f₂ (easier)

Most efficient method: Add a capacitor in the Miller position (drain to gate for FET, collector to base for BJT).

Why this works:

  • One physical capacitor affects both input and output networks simultaneously
  • "Almost like getting a second capacitor for free"

Design process (from Example 17.7):

  1. Calculate required total capacitance for each network to reach target frequency
  2. Subtract existing device capacitances
  3. Work backward through Miller equations to find required CM
  4. Use the smaller of the two calculated values

Example: To shift f₂ from 22.1 MHz down to 100 kHz, the output network required CM = 332 pF (smaller than the 5.3 nF needed for input network alone).

📈 Raising f₂ (harder)

Requires decreasing either:

  • Device capacitances: May need a different transistor model
  • Surrounding resistances: May require complete circuit redesign

This is "a bit of work" compared to lowering f₂.

🖥️ Simulation verification

✅ Validation approach

The excerpt demonstrates crosschecking calculations with simulation:

TestMethodResult
Midband gainTransient analysis at 1 kHz500 mV output from 100 mV input confirms gain of 5
System f₂AC analysis Bode plot25.25 MHz measured vs 22.1 MHz calculated (within 15%)
Input networkScale Rgen by 100× to make input dominant2.4 MHz measured vs 2.65 MHz expected

Why deviations occur: Slight variances in device capacitances, parasitic capacitances from resistors and wiring, device-to-device variations.

🎲 Scaling trick for input verification

  • Increase generator internal resistance (Rgen) by factor of 100
  • This shifts input critical frequency down by same factor without affecting gain or output network
  • Makes input network dominant for testing purposes
  • "Nice enough trick for the simulator, but often not possible or desirable in real world circuits"

📋 Summary principles

🔑 Key takeaways from the excerpt

All amplifiers have f₂: Upper frequency limit exists for all transistor amplifiers due to parasitic capacitances.

Two lag networks: Typical single-stage inverting amplifier exhibits input lag and output lag networks.

Miller conversion is essential: Any capacitance bridging input to output must be converted using Miller's theorem into equivalent input and output capacitances that appear in parallel with other device and stray capacitances.

Dominant network rules: The lower of the two lag network frequencies sets system f₂; if networks are close in frequency, interaction shifts f₂ even lower.

Design flexibility differs by direction: Reducing f₂ is straightforward (add capacitance, especially in Miller position), but increasing f₂ may require transistor selection changes or major redesign.

79

Amplifier Circuit Analysis Problems

12.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

This excerpt presents a collection of circuit analysis problems focused on determining critical frequencies (lead and lag) and system bandwidth limits for various amplifier configurations using BJTs and FETs.

📌 Key points (3–5)

  • What the problems ask: calculate lead critical frequencies (low-frequency cutoffs, f₁) and lag critical frequencies (high-frequency cutoffs, f₂) for amplifier circuits.
  • Circuit types covered: BJT amplifiers (common-emitter, emitter-follower) and FET amplifiers (common-source, source-follower) with various biasing schemes.
  • Key components: coupling capacitors (Cᵢₙ, Cₒᵤₜ), bypass capacitors (Cₑ, Cₛ), and parasitic capacitances (Cᵢᵦₒ, Cₒᵦₒ, Cᵍˢ, Cᵈᵍ).
  • Design and simulation tasks: some problems require selecting component values to achieve target frequencies, and others ask for computer simulation verification.
  • Common confusion: lead frequencies (f₁) are low-frequency cutoffs determined by coupling/bypass capacitors; lag frequencies (f₂) are high-frequency cutoffs determined by parasitic/stray capacitances—do not confuse which capacitors affect which end of the bandwidth.

🔌 Lead critical frequency problems (low-frequency cutoff)

🔌 What lead frequencies represent

Lead critical frequencies: the low-frequency cutoff points where coupling and bypass capacitors begin to block signal transmission.

  • At low frequencies, coupling capacitors (Cᵢₙ, Cₒᵤₜ) and bypass capacitors (Cₑ, Cₛ) have high reactance and attenuate the signal.
  • The system f₁ is the highest of the individual critical frequencies (or adjusted upward if frequencies are close together, "due to proximity").
  • Example: if input fᶜ = 67 Hz, output fᶜ = 2.65 Hz, bypass fᶜ = 61.2 Hz, then f₁ > 67 Hz because the input and bypass frequencies are close.

🧮 Typical problem structure (Problems 1–16)

  • Given: supply voltages (Vᴄᴄ, Vᴇᴇ, Vᴅᴅ, Vₛₛ), transistor parameters (β, gₘ), resistor values (Rᴳᴱᴺ, Rʙ, Rᴇ, Rᴄ, Rʟ, etc.), and capacitor values (Cᵢₙ, Cₒᵤₜ, Cₑ/Cₛ).
  • Find: individual critical frequencies at input, output, and bypass nodes, then determine the system f₁.
  • The problems span BJT circuits (Figures 17.41–17.45) and FET circuits (Figures 17.46–17.49).

🔄 BJT vs FET lead frequency analysis

Transistor typeBypass capacitorTypical biasingExample problems
BJTCₑ (emitter bypass)Voltage divider, emitter bias1–10
FETCₛ (source bypass)Self-bias, voltage divider11–16
  • Both types use coupling capacitors at input and output.
  • The bypass capacitor shorts AC signals around the emitter/source resistor to increase gain; its cutoff frequency is part of the f₁ calculation.

📡 Lag critical frequency problems (high-frequency cutoff)

📡 What lag frequencies represent

Lag critical frequencies: the high-frequency cutoff points where parasitic and stray capacitances begin to shunt signal to ground.

  • At high frequencies, small parasitic capacitances (Cᵢᵦₒ, Cₒᵦₒ for BJTs; Cᵍˢ, Cᵈᵍ for FETs) and stray capacitances have low reactance and attenuate the signal.
  • The system f₂ is the lowest of the individual critical frequencies (or adjusted downward if frequencies are close, "due to proximity").
  • Example: if input fᶜ = 4.78 MHz, output fᶜ = 13.6 MHz, then f₂ < 4.78 MHz because the two are close.

🧮 Typical problem structure (Problems 17–30)

  • Given: same circuit parameters as lead problems, but now parasitic capacitances (Cᵢᵦₒ, Cₒᵦₒ, Cᵍˢ, Cᵈᵍ, Cᵢₛₛ, Cᵣₛₛ) and stray capacitances (Cᵢₙ₍ₛₜᵣₐᵧ₎, Cₒᵤₜ₍ₛₜᵣₐᵧ₎) are specified.
  • Find: individual critical frequencies at input and output nodes (due to parasitic/stray capacitances), then determine the system f₂.
  • Problems 29–30 use alternate notation (Cᵢₛₛ, Cᵣₛₛ) for FET capacitances, which are equivalent to Cᵍˢ and Cᵈᵍ.

🔍 Don't confuse lead and lag capacitors

  • Lead (low-frequency) capacitors: large coupling and bypass capacitors (microfarads) that block DC and pass AC; their reactance is high at low frequencies.
  • Lag (high-frequency) capacitors: tiny parasitic and stray capacitances (picofarads) that shunt AC to ground; their reactance is low at high frequencies.
  • The two sets of capacitors affect opposite ends of the amplifier bandwidth.

🛠️ Design problems (component selection)

🛠️ What design problems ask

  • Task: choose capacitor values to achieve a specified system f₁ or individual critical frequencies.
  • Example (Problem 31): alter coupling and bypass capacitors so that input fᶜ = 200 Hz, output fᶜ = 20 Hz, bypass fᶜ = 2 Hz.
  • Example (Problem 32): alter capacitors so that system f₁ = 100 Hz.

🧩 Design strategy (from answer key)

  • Multiple solutions exist: "The remainder of the design has nearly infinite possibilities."
  • Start with bias point: choose a mid-point bias (e.g., Vᴳₛ = 0.5 Vᴳₛ₍ₒff₎, Iᴅ = 0.25 Iᴅₛₛ) to set DC operating point.
  • Check gain ceiling: verify that the desired gain is achievable (e.g., if gₘ₀ = 25 mS and Rʟ = 10 kΩ, ceiling gain ≈ 250 unswamped).
  • Split resistances: divide source/emitter resistance into bypassed and unbypassed portions to control gain via degeneration.
  • Example (Problem 13 answer): using Rₛ = 160 Ω total, split into Rₛ = 60 Ω and Rₛᵥᵥ = 100 Ω to achieve gain of 5.

📋 Design problem types

ProblemGoalApproach
31, 33, 35, 37Set individual fᶜ valuesCalculate required C from fᶜ = 1/(2πRC)
32, 34, 36, 38Set system f₁ or paired fᶜChoose one C, then increase others to shift f₁
  • "Pick one and increase the other(s)" means: set one capacitor to achieve its target frequency, then make the others larger so they don't limit the bandwidth.

🖥️ Challenge and simulation problems

🖥️ Challenge problems (39–42)

  • Circuits: more complex topologies (Figures 17.50–17.51) not covered in the basic problem sets.
  • Tasks: same as basic problems—find lead and lag critical frequencies and system f₁/f₂.
  • These test the ability to apply frequency analysis methods to unfamiliar circuit configurations.

🖥️ Computer simulation problems (43–45)

  • Task: use circuit simulation software to verify the results of selected design problems.
  • What to verify: all critical frequencies (input, output, bypass for lead; input, output for lag).
  • Example (Problem 43): simulate Design Problem 32 and confirm that the chosen capacitors produce the target system f₁.
  • Simulation helps catch calculation errors and confirms that the design meets specifications across the full frequency range.
80

Circuit Analysis Problems: Lead and Lag Critical Frequencies

12.1 Introduction

🧭 Overview

🧠 One-sentence thesis

This section provides practice problems for calculating lead and lag critical frequencies and system cutoff frequencies (f₁ and f₂) in BJT and FET amplifier circuits with various component values and configurations.

📌 Key points (3–5)

  • Problem structure: Each problem specifies a circuit figure and component values, asking for critical frequencies and system cutoff frequencies.
  • Two types of analysis: Lead critical frequencies (with system f₁) involve coupling and bypass capacitors; lag critical frequencies (with system f₂) involve parasitic and stray capacitances.
  • Circuit types covered: BJT circuits (common-emitter with various biasing schemes) and FET circuits (common-source configurations).
  • Common confusion: Lead vs lag—lead frequencies use large coupling/bypass capacitors (microfarads) and determine the low-frequency cutoff; lag frequencies use small parasitic capacitances (picofarads) and determine the high-frequency cutoff.
  • Parameter variations: Problems vary supply voltages, transistor parameters (β, gₘ), resistor values, and capacitor values to provide comprehensive practice.

🔌 Lead critical frequency problems (low-frequency analysis)

🔌 BJT amplifier configurations

Problems 3–10 focus on finding lead critical frequencies and system f₁ for BJT circuits:

  • Circuit topologies: Include split-supply configurations (VCC and VEE), voltage-divider bias, and swamping resistor designs.
  • Key capacitors analyzed: Input coupling (Cᵢₙ), output coupling (Cₒᵤₜ), and emitter bypass (CE) capacitors, typically in the microfarad range (1 μF to 680 μF).
  • Typical component ranges:
    • Supply voltages: 15 V to 32 V
    • Beta (β): 100 to 200
    • Resistors: 50 Ω (generator) to 60 kΩ (load)

Example: Problem 3 specifies VCC = 20 V, VEE = −8 V, β = 100, with Cᵢₙ = 1 μF, Cₒᵤₜ = 2 μF, CE = 100 μF for the circuit in Figure 17.41.

🔌 FET amplifier configurations

Problems 11–16 address FET circuits:

  • Circuit types: Common-source amplifiers with source degeneration, using voltage-divider or direct gate biasing.
  • Key parameters: Transconductance (gₘ) replaces β, ranging from 2 mS to 10 mS.
  • Capacitor types: Input coupling (Cᵢₙ), output coupling (Cₒᵤₜ), and source bypass (CS) capacitors.
  • Capacitor ranges: Input capacitors vary from 10 nF to 470 nF; output and source capacitors from 2.2 μF to 680 μF.

Example: Problem 11 uses VDD = 20 V, gₘ = 2 mS, with Cᵢₙ = 100 nF, Cₒᵤₜ = 10 μF, CS = 22 μF.

🔌 What lead frequencies determine

Lead critical frequencies: the frequencies determined by coupling and bypass capacitors that set the low-frequency cutoff of the amplifier.

  • System f₁ is the overall low-frequency cutoff, typically the highest of the individual lead critical frequencies.
  • These frequencies limit how low in frequency the amplifier can effectively amplify signals.
  • Don't confuse: Larger capacitor values produce lower critical frequencies (better low-frequency response).

🔍 Lag critical frequency problems (high-frequency analysis)

🔍 BJT high-frequency analysis

Problems 17–24 focus on lag critical frequencies and system f₂ for BJT circuits:

  • Key capacitances: Base-emitter capacitance (Cᵢᵦₒ), collector-base capacitance (Cₒᵦₒ), and stray capacitances, all in the picofarad range (1 pF to 20 pF).
  • Stray capacitances: Some problems include Cᵢₙ₍ₛₜᵣₐᵧ₎ and Cₒᵤₜ₍ₛₜᵣₐᵧ₎ (10 pF to 100 pF).
  • Same circuits, different analysis: Many problems reference the same circuit figures as the lead frequency problems but with parasitic capacitance values instead of coupling capacitors.

Example: Problem 17 uses Figure 17.41 with VCC = 32 V, β = 100, Cᵢᵦₒ = 5 pF, Cₒᵦₒ = 2 pF.

🔍 FET high-frequency analysis

Problem 25 begins addressing FET lag frequencies (text is cut off):

  • Expected parameters: Gate-source and gate-drain capacitances, similar in magnitude to BJT parasitic capacitances.
  • Circuit reference: Figure 17.46, which was used earlier for lead frequency analysis.

🔍 What lag frequencies determine

Lag critical frequencies: the frequencies determined by parasitic and stray capacitances that set the high-frequency cutoff of the amplifier.

  • System f₂ is the overall high-frequency cutoff, typically the lowest of the individual lag critical frequencies.
  • These frequencies limit how high in frequency the amplifier can effectively amplify signals.
  • Don't confuse: Smaller capacitance values produce higher critical frequencies (better high-frequency response), but parasitic capacitances are unavoidable and limit performance.

📊 Problem organization and patterns

📊 Circuit figure reuse

FigureLead frequency problemsLag frequency problemsCircuit type
17.41317, 18BJT split-supply
17.42419, 20BJT split-supply
17.435, 621, 22BJT (configuration not specified)
17.447, 823, 24BJT voltage-divider bias
17.459, 10(none shown)BJT (configuration not specified)
17.4611, 1225 (incomplete)FET common-source
17.4713, 14(none shown)FET split-supply
17.4815, 16(none shown)FET (configuration not specified)

📊 Parameter variation strategy

The problems systematically vary parameters to test understanding:

  • Same circuit, different values: Problems often come in pairs using the same figure but different component values (e.g., problems 3 & 4, 7 & 8).
  • Beta/transconductance variation: Some problem pairs vary only β or gₘ while keeping other parameters constant (e.g., problems 5 & 6 with β = 100 vs 200).
  • Capacitor value ranges: Lead frequency problems use capacitors spanning three orders of magnitude (nanofarads to hundreds of microfarads); lag frequency problems use picofarad-range capacitances.

📊 Incomplete content

The excerpt ends mid-problem (problem 25), indicating this is part of a larger problem set. The pattern suggests additional FET lag frequency problems would follow.

81

12.2 The DE-MOSFET

12.2 The DE-MOSFET

🧭 Overview

🧠 One-sentence thesis

The excerpt provided contains only circuit analysis problem sets with numerical parameters and does not present substantive content about DE-MOSFET theory, operation, or characteristics.

📌 Key points (3–5)

  • The excerpt consists entirely of homework/practice problems numbered 11–35.
  • Problems involve calculating critical frequencies (f₁, f₂) for various amplifier circuits.
  • Circuit types mentioned include BJT and MOSFET configurations (Figures 17.41–17.49).
  • Parameters given include supply voltages, transconductance (gₘ), resistances, capacitances, and transistor parameters (β).
  • No explanatory text, definitions, or conceptual material about DE-MOSFETs is present.

📋 Content analysis

📋 What the excerpt contains

The excerpt is a problem set from what appears to be Chapter 17 of a textbook. It includes:

  • Problem numbers: 11 through 35 (partial)
  • Circuit references: Figures 17.41 through 17.49
  • Problem types:
    • Analysis problems (determining critical frequencies)
    • Design problems (selecting component values to achieve target frequencies)

🔢 Typical problem structure

Each analysis problem follows this pattern:

  • Reference to a circuit figure
  • Request to find "lead critical frequencies and system f₁" or "lag critical frequencies and system f₂"
  • List of given parameters (voltages, resistances, capacitances, transistor parameters)

Example parameter sets include:

  • Supply voltages (VDD, VCC, VEE, VSS)
  • Transconductance (gₘ) in milliSiemens
  • Resistor values in ohms or kilohms
  • Capacitor values in picofarads, nanofarads, or microfarads
  • Transistor gain (β) for BJT circuits

🛠️ Design problems

Problems 31–35 ask students to:

  • Modify coupling and bypass capacitor values
  • Achieve specified critical frequency targets
  • Meet system bandwidth requirements

⚠️ Limitation notice

⚠️ Missing conceptual content

This excerpt does not contain:

  • Definitions or explanations of DE-MOSFET structure
  • Operating principles or characteristics
  • Comparison with other MOSFET types
  • Theory of critical frequencies or frequency response
  • Circuit analysis methods or formulas

The title "12.2 The DE-MOSFET" suggests this should be a section explaining depletion-enhancement MOSFETs, but the provided text contains only end-of-chapter problems from a different chapter (Chapter 17) focused on frequency response analysis.

82

12.3 DE-MOSFET Biasing

12.3 DE-MOSFET Biasing

🧭 Overview

🧠 One-sentence thesis

The excerpt does not contain substantive content on DE-MOSFET biasing; it consists solely of circuit analysis problem statements, design problems, challenge problems, and appendix material unrelated to the stated title.

📌 Key points (3–5)

  • The excerpt contains numbered problems (18–45) asking readers to calculate critical frequencies and system parameters for various circuits.
  • Problems reference figures (17.42, 17.43, 17.44, 17.46, 17.48, 17.49, 17.50, 17.51) that are not included in the excerpt.
  • Circuit parameters include BJT parameters (β, C_ibo, C_obo) and FET parameters (g_m, C_gs, C_dg, C_iss, C_rss).
  • Design problems ask for capacitor values to achieve specific critical frequencies.
  • The appendix section lists manufacturer data sheet links for diodes, not MOSFETs.

📋 Content summary

📋 What the excerpt contains

The excerpt is composed of:

  • Analysis problems (18–30): Calculate lag critical frequencies and system f₂ for given circuits with specified component values.
  • Design problems (31–38): Determine coupling and bypass capacitor values to achieve target critical frequencies.
  • Challenge problems (39–42): More complex frequency analysis tasks combining lead and lag critical frequencies.
  • Computer simulation problems (43–45): Verify design and challenge problem results through simulation.
  • Appendix A: Links to manufacturer data sheets for 1N4002 rectifier and 1N4148/1N914 switching diodes.

⚠️ What is missing

  • No explanatory text on DE-MOSFET biasing concepts, principles, or techniques.
  • No definitions of DE-MOSFET or biasing methods.
  • No circuit diagrams (all referenced figures are external).
  • No worked examples or solution methods.
  • The title "12.3 DE-MOSFET Biasing" does not match the problem-set content.

🔧 Problem types overview

🔧 Analysis problems

Problems 18–30 follow a consistent pattern:

  • Given: circuit figure reference, supply voltages (V_CC, V_EE, or V_DD), transistor parameters (β or g_m), resistor values, and capacitance values.
  • Find: lag critical frequencies and system f₂.
  • Circuits include both BJT configurations (problems 18–24) and FET configurations (problems 25–30).

🔧 Design problems

Problems 31–38 reverse the analysis task:

  • Given: circuit figure and transistor parameters.
  • Find: capacitor values (coupling and/or bypass) to achieve specified critical frequencies (e.g., input f_c = 200 Hz, output f_c = 20 Hz).
  • Some problems specify system f₁ instead of individual critical frequencies.

🔧 Challenge and simulation problems

  • Challenge problems (39–42) combine lead and lag frequency analysis.
  • Simulation problems (43–45) ask for verification of earlier design and challenge problem results using circuit simulation software.

📊 Parameter patterns

📊 BJT circuit parameters

Common parameters in BJT problems:

  • β (current gain): ranges from 100 to 200.
  • C_ibo (input capacitance): ranges from 2 pF to 20 pF.
  • C_obo (output capacitance): ranges from 1 pF to 4 pF.
  • Supply voltages: V_CC (15 V to 25 V), V_EE (−8 V to −12 V).

📊 FET circuit parameters

Common parameters in FET problems:

  • g_m (transconductance): ranges from 2 mS to 12 mS.
  • C_gs (gate-source capacitance): ranges from 3 pF to 20 pF.
  • C_dg (drain-gate capacitance): ranges from 2 pF to 4 pF.
  • Some problems use C_iss and C_rss notation instead.
  • Supply voltage: V_DD (15 V to 20 V).
83

12.4 The E-MOSFET

12.4 The E-MOSFET

🧭 Overview

🧠 One-sentence thesis

The excerpt does not contain substantive content about E-MOSFETs; it consists only of problem sets, appendices listing manufacturer data sheet links, and standard component size tables.

📌 Key points (3–5)

  • The excerpt includes circuit analysis problems referencing E-MOSFET parameters (transconductance g_m, capacitances C_gs, C_dg, C_iss, C_rss) but does not explain E-MOSFET operation or theory.
  • Appendix A lists manufacturer data sheet links for power E-MOSFETs (FDMS86180, IRF7201) but provides no technical explanation.
  • Appendix B provides standard resistor and capacitor value tables (5%, 10%, 1%, 2% tolerances) unrelated to E-MOSFET concepts.
  • No definitions, operating principles, construction details, or characteristic curves for E-MOSFETs are present.
  • The excerpt is supplementary material (problems and reference tables) rather than instructional content.

📋 Problem set references

🔧 Circuit analysis problems

The excerpt contains numbered problems (25–42) that reference E-MOSFET circuits:

  • Parameters mentioned: supply voltage V_DD, transconductance g_m, resistances (R_gen, R_G, R_D, R_S, R_L), and capacitances (C_gs, C_dg, C_iss, C_rss).
  • Tasks requested: determine lag critical frequencies and system f_2 values for given circuits.
  • No explanatory content: problems assume prior knowledge; no definitions or methods are provided in the excerpt.

Example: Problem 25 asks to find lag critical frequencies given V_DD = 20 V, g_m = 2 mS, and various component values, but does not explain what lag critical frequencies are or how E-MOSFETs function.

🛠️ Design and simulation problems

  • Design problems (31–38): ask to alter coupling/bypass capacitors to achieve specific frequency responses.
  • Challenge problems (39–42): request determination of lead/lag critical frequencies for more complex circuits.
  • Simulation problems (43–45): instruct to verify design results using computer simulation.
  • All assume familiarity with E-MOSFET behavior and frequency response analysis.

📚 Appendix material

🔗 Manufacturer data sheets (Appendix A)

The excerpt lists web links to component datasheets, including two power E-MOSFETs:

ComponentManufacturerNote
FDMS86180ON SemiconductorPower E-MOSFET
IRF7201InfineonPower E-MOSFET
  • Links are provided for convenience; no technical specifications or explanations are extracted.
  • The appendix notes links are current as of Fall 2022 and advises checking manufacturer sites for latest information.

📊 Standard component values (Appendix B)

Tables list standard resistor values at different tolerances:

  • 10% tolerance (EIA E12): 12 values per decade (10, 12, 15, 18, 22, 27, 33, 39, 47, 56, 68, 82).
  • 5% tolerance (EIA E24): 24 values per decade (includes all 10% values plus intermediate steps).
  • 2% tolerance (EIA E48) and 1% tolerance (EIA E96): finer gradations for precision applications.
  • Capacitors below 10 nF typically available at 5% standards; larger values at 20% standards.
  • This information is unrelated to E-MOSFET theory or operation.

⚠️ Content limitations

⚠️ Missing instructional material

The excerpt does not contain:

  • Definitions of E-MOSFET (Enhancement-mode MOSFET) or how it differs from depletion-mode devices.
  • Operating principles, threshold voltage, or transfer characteristics.
  • Circuit symbols, construction diagrams, or biasing techniques.
  • Explanations of transconductance, capacitances, or frequency response in E-MOSFET context.

Don't confuse: This excerpt is supplementary material (problems and references) appended to a textbook section, not the instructional content of section 12.4 itself.

📖 Intended use

  • The problems and appendices assume the reader has already studied E-MOSFET theory from the main text (not included in this excerpt).
  • They serve as practice exercises and quick-reference tables for students and designers working with E-MOSFETs and related circuits.
84

E-MOSFET Data Sheet Interpretation

12.5 E-MOSFET Data Sheet Interpretation

🧭 Overview

🧠 One-sentence thesis

This excerpt provides manufacturer data sheet links and standard component reference tables for electronic components, including power E-MOSFETs, to support circuit design and simulation work.

📌 Key points (3–5)

  • Data sheet access: Manufacturer links are provided for various electronic components as of Fall 2022, with a reminder to check for latest information before new designs.
  • E-MOSFET examples: Two specific power E-MOSFET devices are listed with their data sheet URLs (FDMS86180 and IRF7201).
  • Standard component values: Passive components like resistors and capacitors come in standardized sizes organized by tolerance levels (EIA standards).
  • Common confusion: Capacitor availability differs from resistors—smaller capacitors (below 10 nF) typically follow 5% standards, while larger ones use 20% standards.
  • Design context: The excerpt appears in an appendix supporting circuit design problems involving transistors and frequency response calculations.

📋 Component Data Sheet Resources

🔗 Power E-MOSFET devices

The excerpt lists two power E-MOSFET data sheets:

  • FDMS86180: Available from ON Semiconductor
  • IRF7201: Available from Infineon

These links serve as reference materials for designers working with enhancement-mode MOSFETs in power applications.

⚠️ Currency warning

  • Links are current as of Fall 2022
  • Users must verify the manufacturer's site for latest information before beginning new designs
  • Some devices may be available from multiple manufacturers

Don't confuse: These are reference links for convenience, not guaranteed permanent URLs—always verify current specifications directly from manufacturers.

📐 Standard Component Sizing

🎯 Resistor tolerance standards

The excerpt provides tables organized by tolerance level:

ToleranceStandardNumber of values per decade
20%Seldom usedEvery fourth value from 10
10%EIA E1212 values (shown in bold)
5%EIA E2424 values
2%EIA E4848 values (shown in bold)
1%EIA E9696 values
  • The same digit patterns repeat in subsequent decades up to at least 1 megohm
  • Higher decades beyond 1 megohm are not shown in the tables

🔌 Capacitor availability differences

Capacitors follow different availability patterns than resistors:

  • Smaller capacitors (below 10 nF or 0.01 μF): Usually available at 5% standard digits
  • Larger capacitors: Tend to be available at 20% standards
  • Capacitors and inductors generally have fewer standard values than resistors

Example: A designer needing a 0.005 μF capacitor would find more standard value options than someone needing a 100 μF capacitor.

🔧 Design Problem Context

🧮 Related circuit problems

The excerpt follows problems involving:

  • Coupling capacitor value determination for specific cutoff frequencies
  • Critical frequency calculations for BJT and MOSFET circuits
  • Transconductance (g_m) parameter usage ranging from 3 mS to 12 mS
  • Computer simulation verification tasks

🎓 Challenge and simulation problems

  • Problems 39-42 involve determining lead and lag critical frequencies
  • Problems 43-45 require simulation verification of design results
  • Circuits reference both BJT (beta parameters) and MOSFET (transconductance) devices

Don't confuse: The data sheet appendix supports the design problems but does not contain the actual circuit analysis—it provides component specifications needed for practical implementation.

85

E-MOSFET Biasing

12.6 E-MOSFET Biasing

🧭 Overview

🧠 One-sentence thesis

The excerpt provided contains only reference materials (component datasheets, standard resistor values, and problem answers) and does not present substantive content on E-MOSFET biasing techniques or principles.

📌 Key points (3–5)

  • The excerpt consists of appendices listing component datasheets, standard passive component sizes, and answers to textbook problems.
  • No theoretical content, circuit analysis methods, or biasing configurations for E-MOSFETs are included in this excerpt.
  • The material appears to be supplementary reference tables from a textbook's back matter.
  • Chapter-by-chapter problem answers are provided but without the original problems or explanatory context.

📋 Content analysis

📋 What the excerpt contains

The provided text includes three appendices:

  • Appendix A: Links to component datasheets (including two E-MOSFET power devices: FDMS86180 and IRF7201)
  • Appendix B: Standard resistor value tables for different tolerance levels (5%, 10%, 1%, 2%, 20%)
  • Appendix C: Numerical answers to selected problems from Chapters 2-8

❌ What is missing

The excerpt does not contain:

  • Explanations of E-MOSFET operation or characteristics
  • Biasing circuit configurations or design methods
  • Analysis techniques or equations
  • Conceptual discussions of enhancement-mode MOSFET behavior
  • Design examples or worked problems

🔍 Observable details

🔍 Component references

Two E-MOSFET power devices are mentioned in the datasheet list:

  • FDMS86180 Power E-MOSFET (ON Semiconductor)
  • IRF7201 Power E-MOSFET (Infineon)

These are reference links only, with no accompanying circuit analysis or application information.

🔍 Problem answer format

The problem answers follow a consistent pattern showing calculated values for various electrical parameters (voltages, currents, impedances), but without the original problem statements or solution methods, these numbers provide no instructional value for understanding E-MOSFET biasing.

86

Appendix C: Answers to Selected Numbered Problems

Chapter 13: MOSFET Small Signal Amplifiers

🧭 Overview

🧠 One-sentence thesis

This appendix provides numerical answers to selected end-of-chapter problems spanning diode circuits, BJT biasing and amplifiers, FET circuits, and power amplifiers, enabling students to verify their work across fundamental analog electronics topics.

📌 Key points (3–5)

  • Coverage span: answers range from Chapter 2 (basic diode circuits) through Chapter 11 (FET amplifiers), covering the core progression of analog electronics.
  • Problem types: includes DC biasing calculations (currents, voltages, Q-points), AC small-signal parameters (input/output impedance, voltage gain), and power amplifier metrics (compliance, efficiency, maximum power).
  • Common confusion: cutoff vs saturation parameters—cutoff voltage (V_CE(cutoff)) is the maximum voltage when collector current is zero, while saturation current (I_C(sat)) is the maximum current when the transistor is fully on; these define the load line endpoints, not operating point values.
  • Verification purpose: odd-numbered problems typically have answers provided so students can check their understanding before exams.

📐 Diode and Basic BJT Circuits (Chapters 2–4)

🔌 Chapter 2: Basic diode circuits

  • Answers include simple current and voltage calculations.
  • Example values: 0.53 mA, 4.37 V, 8.79 mA.
  • These represent fundamental DC analysis results for diode-resistor networks.

📊 Chapter 3: Clipping and clamping circuits

  • Problems involve sinusoidal waveforms with clipping levels.
  • Example: "10sin2π100t, positive clipped at 8.7 V" means a 10 V peak sine wave at 100 Hz with the positive peaks limited to 8.7 V by a diode clipper.
  • Biased clippers add a DC offset to the clipping threshold.
  • Clamping circuits shift the DC level: "8sin2π500t + 9.3 VDC" indicates a 9.3 V DC component added to the AC signal.

⚡ Chapter 4: BJT fundamentals

  • Basic current relationships using beta (current gain = 99 in Problem 1).
  • Example breakdown for Problem 3:
    • Base current I_B = 21.5 microamperes
    • Collector current I_C = 2.15 milliamperes (beta times I_B)
    • Emitter current I_E = 2.175 milliamperes (sum of I_B and I_C)
  • Problem 13 shows sensitivity: a parameter change results in "47% of former value," illustrating how component variations affect circuit performance.

🎯 BJT Biasing and Load Lines (Chapter 5)

📈 Load line parameters

Each problem provides four key values that define the DC load line and Q-point:

ParameterMeaningExample (Problem 1)
V_CE(cutoff)Maximum collector-emitter voltage (I_C = 0)28 V
I_C(sat)Maximum collector current (V_CE ≈ 0)1.27 mA
I_CQQuiescent (operating point) collector current0.73 mA
V_CEQQuiescent collector-emitter voltage11.94 V

🎚️ Q-point positioning

  • The Q-point (I_CQ, V_CEQ) should ideally sit near the center of the load line for maximum signal swing without distortion.
  • Example: Problem 1 has I_CQ = 0.73 mA (about 57% of I_C(sat)) and V_CEQ = 11.94 V (about 43% of V_CE(cutoff))—reasonably centered.
  • Problem 15 shows V_CEQ = 2.66 V with V_CE(cutoff) = 18 V—the Q-point is much closer to saturation, limiting negative signal swing.
  • Don't confuse: cutoff and saturation values are the limits of the load line, not the actual operating point; the Q-point must lie between them.

🔧 Design problems

  • Problems 19 and 21 ask for resistor values (3.65 kΩ, 7.58 kΩ) to achieve specific biasing conditions.
  • These reverse the analysis process: given desired Q-point, find component values.

🔊 Small Signal Amplifier Analysis (Chapters 6–7, 11)

🎛️ Key AC parameters

Small-signal amplifier problems report three main quantities:

Input impedance (Z_in): the AC resistance seen by the signal source looking into the amplifier input.

Output impedance (Z_out): the AC resistance seen by the load looking back into the amplifier output.

Voltage gain (A_v): the ratio of output voltage to input voltage; negative values indicate inversion.

📉 Chapter 7: BJT amplifier configurations

Example from Problem 3:

  • Z_in = 9.08 kΩ (relatively high, good for voltage amplification)
  • Z_out = 22 kΩ (high output impedance, typical of common-emitter without emitter bypass)
  • v_load = 4 V (inverted)—the output is 180 degrees out of phase with the input

Example from Problem 9:

  • Z_in = 90.4 kΩ (very high, likely common-collector/emitter-follower configuration)
  • Z_out = 4.1 Ω (very low, characteristic of emitter-follower for impedance matching)
  • v_load = 182 mV (no inversion noted, confirming emitter-follower)

🔄 Design adjustments (Problems 15, 17)

  • Problem 15: "Set R_SW = 300 Ω to maintain A_v (note r'_e change)"—changing one component affects the dynamic emitter resistance, requiring compensation.
  • Problem 17: "Set R_SW = 49 Ω to double A_v (including effect of r'_e), increase R_E to 1049 Ω to maintain Q"—shows the trade-off between gain and bias stability.
  • Don't confuse: r'_e (dynamic emitter resistance) changes with operating current, so AC gain and DC bias interact.

🎚️ Chapter 11: FET amplifiers

  • Similar structure but typically higher input impedances due to FET gate isolation.
  • Example Problem 1: Z_in = 220 kΩ, A_v = -7.5 (inverted).
  • Example Problem 7: Z_in = 220 kΩ, A_v = 0 (unity gain, likely source-follower configuration).

⚡ Power Amplifiers and Thermal Design (Chapters 8–9)

🔥 Chapter 8: Class A power amplifier metrics

Each problem provides a comprehensive set of power and efficiency parameters:

ParameterMeaningExample (Problem 1)
ComplianceMaximum undistorted AC voltage swing (peak)1.55 Vp
P_l(max)Maximum power delivered to load16 mW
P_D(max)Maximum power dissipated in transistor161 mW
η (eta)Efficiency (load power / total DC power)3.7%

📊 Efficiency observations

  • Class A amplifiers show very low efficiency in these examples: 3.7%, 3.16%, 1.16%, 0.94%.
  • Most DC power is dissipated as heat in the transistor, not delivered to the load.
  • Example: Problem 1 delivers only 16 mW to the load but dissipates 161 mW in the transistor—about 10 times more waste heat than useful output.

🌡️ Chapter 9: Thermal and power ratings

BV_CEO: breakdown voltage with base open, the maximum collector-emitter voltage the transistor can withstand.

Example Problem 1 ratings:

  • BV_CEO = 30 V (voltage limit)
  • I_C(max) = 0.9375 A (current limit)
  • P_l(max) = 7.03 W (maximum load power)
  • P_D(max) = 1.4 W (maximum transistor dissipation, limited by thermal considerations)

🔧 Thermal resistance (Problem 13)

  • Answer: 4.07 °C/W (degrees Celsius per watt).
  • This value describes how much the transistor temperature rises for each watt of power dissipated.
  • Example: if dissipating 1 W, the junction temperature rises 4.07 °C above the case temperature.
  • Lower thermal resistance means better heat transfer and higher power handling capability.

🔌 FET Biasing (Chapter 10)

🎯 DC operating point calculations

FET problems report drain current (I_D) and drain-source voltage (V_DS), analogous to BJT collector current and V_CE.

Example comparisons:

ProblemI_DV_DSV_GV_DNotes
110 mA14 VSelf-bias or fixed bias
32.11 mA0 V32.2 VV_G = 0 suggests self-bias with source resistor
53.77 mA9.9 VModerate operating point
74.15 mA13.8 V0 VSimilar to Problem 3 configuration

🔧 Design problems (9, 11)

  • Problem 9: R_S = 180 Ω (source resistor for self-bias).
  • Problem 11: R_E = 2.825 kΩ (likely a comparison problem showing BJT equivalent).
  • These problems reverse the analysis: given desired I_D or V_DS, find the resistor value needed.

⚠️ Common confusion: gate voltage

  • When V_G = 0 V is explicitly stated (Problems 3, 7), the gate is at ground potential.
  • This typically indicates a self-bias configuration where the source resistor creates the necessary gate-source voltage through the voltage drop from drain current.
  • Don't confuse: V_G = 0 does not mean V_GS = 0; the source sits above ground by I_D × R_S, making V_GS negative for an N-channel JFET or depletion-mode MOSFET.
87

Chapter 13 Objectives

13.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

This excerpt contains only numerical answers to end-of-chapter problems for Chapters 6–13, covering transistor and amplifier circuit analysis, and does not present substantive conceptual content or learning objectives.

📌 Key points (3–5)

  • What the excerpt is: a list of numerical solutions (voltages, currents, impedances, gains) for practice problems across multiple chapters.
  • Chapters covered: Chapters 6 through 13, spanning BJT amplifiers, FET amplifiers, biasing, and power calculations.
  • No conceptual explanations: the excerpt provides only final calculated values without derivations, definitions, or explanatory text.
  • Common confusion: this is an answer key, not a teaching section—it does not explain how to arrive at these values or why the methods work.

📋 Structure of the excerpt

📋 Answer key format

  • Each chapter heading (e.g., "Chapter 6," "Chapter 7") is followed by a numbered list of problem solutions.
  • Solutions include:
    • Impedances (Z_in, Z_out)
    • Voltages (V_CEQ, V_DS, V_out)
    • Currents (I_CQ, I_D)
    • Gains (A_v)
    • Resistor values (R_S, R_D, R_G)
    • Power and efficiency (P_l(max), P_D(max), η)
  • Some answers include brief notes in parentheses, e.g., "(inverted)" or "(not particularly practical)."

📋 Chapters represented

ChapterTopic hint from answers
6Basic amplifier parameters (impedances, voltages)
7Input/output impedance, load voltage, inversion
8Biasing, saturation, cutoff, compliance, power dissipation, efficiency
9Breakdown voltage (BV_CEO), maximum current, power limits
10FET DC operating points (I_D, V_DS, V_G)
11FET amplifier impedance and gain
12FET biasing (various configurations)
13FET amplifier design (impedance, gain, component selection)

🔍 What is missing

🔍 No conceptual content

  • The excerpt does not define terms like "cutoff," "saturation," "compliance," "gain," or "impedance."
  • It does not explain the relationships between parameters (e.g., how I_CQ and V_CEQ are related to the load line).
  • It does not describe circuit topologies (common-emitter, common-source, etc.) or design trade-offs.

🔍 No worked examples or methods

  • The excerpt provides only final numerical results.
  • There are no step-by-step calculations, circuit diagrams, or explanations of which formulas were used.
  • Example: Problem 13 in Chapter 13 includes a brief design note ("Using a standard 15 VDC supply and picking a mid-point bias..."), but this is the only instance of partial reasoning in the entire excerpt.

🔍 Purpose of this excerpt

  • This is an answer key for students to check their work after solving problems.
  • It is not a teaching tool or a summary of chapter objectives.
  • Don't confuse: an answer key with a chapter summary or learning-objectives section—this excerpt does not state what students should learn or understand.

📌 Conclusion

📌 Substantive content

The excerpt lacks substantive conceptual, procedural, or explanatory content. It is a reference list of numerical solutions for practice problems in an electronics or circuit-analysis textbook.

📌 Use case

  • For students: verify calculated answers after working through problems.
  • For review: not suitable—review requires definitions, explanations, and worked examples, none of which are present here.
88

Introduction to Chapter Answer Keys

13.1 Introduction

🧭 Overview

🧠 One-sentence thesis

This excerpt presents numerical answers to end-of-chapter problems for electronics topics covering transistor circuits, amplifiers, and signal analysis across Chapters 9–16.

📌 Key points (3–5)

  • What this is: A collection of numerical solutions and parameter values for textbook problems, not explanatory content.
  • Coverage: Spans multiple chapters on transistor biasing, FET circuits, amplifier design, frequency response, and decibel calculations.
  • Format: Lists problem numbers followed by calculated values for voltages, currents, resistances, gains, and other circuit parameters.
  • No conceptual content: The excerpt contains only answers without derivations, explanations, or theory.

📋 Nature of the content

📋 Answer key structure

  • Each chapter section lists problem numbers (odd-numbered: 1, 3, 5, etc.) followed by their solutions.
  • Solutions include multiple parameters per problem, separated by commas.
  • Example format: "1. I_D = 10 mA, V_DS = 14 V" provides two calculated values for problem 1.

🔢 Types of parameters provided

The answers cover various electrical quantities:

Parameter typeExamples from excerpt
CurrentsI_CQ, I_D, i_c(sat)
VoltagesV_CEQ, V_DS, V_out
ResistancesR_S, R_D, Z_in
PowerP_l(max), P_D(max)
GainsA_v (voltage gain)
Efficiencyη (eta)
DecibelsdB, dBW, dBV, dBf

🔌 Chapter topics covered

🔌 Chapters 9–10: Transistor biasing

  • Chapter 9 focuses on BJT (bipolar junction transistor) parameters including breakdown voltages (BV_CEO), maximum currents, and power dissipation.
  • Chapter 10 covers FET (field-effect transistor) DC operating points with drain currents, gate voltages, and drain-source voltages.

📡 Chapters 11–13: Amplifier circuits

  • Chapters 11 and 13 provide input impedances (Z_in), voltage gains (A_v), and output voltages for various amplifier configurations.
  • Many gains are negative (inverted signals) or fractional (less than unity).
  • Chapter 12 continues FET circuit analysis with similar DC parameters.

📊 Chapters 14–16: Advanced topics

  • Chapter 14: Frequency values (kHz) and slew rates (V/μs).
  • Chapter 15: Time constants (nanoseconds).
  • Chapter 16: Extensive decibel calculations in multiple reference systems (dB, dBW, dBV, dBf), power gains, and frequency response with phase angles.

⚠️ Limitations of this excerpt

⚠️ No learning content

  • This excerpt provides no explanations, definitions, or conceptual teaching.
  • It cannot be used to understand why these values are correct or how to calculate them.
  • The problems themselves are not included, only the final numerical answers.

⚠️ Context missing

  • Without the original problem statements, the meaning of each parameter is unclear.
  • Circuit diagrams, component values, and problem requirements are not provided.
  • This is reference material for checking work, not for learning new concepts.
89

13.2 MOSFET Common Source Amplifiers

13.2 MOSFET Common Source Amplifiers

🧭 Overview

🧠 One-sentence thesis

The excerpt provides numerical answers to design and analysis problems for MOSFET common-source amplifiers, demonstrating how input impedance, voltage gain, and output voltage vary with different component choices and biasing conditions.

📌 Key points (3–5)

  • Input impedance (Z_in): ranges from tens of kΩ to over 900 kΩ depending on gate resistor values, reflecting the high-impedance nature of MOSFET gates.
  • Voltage gain (A_v): can be negative (inverting) or positive (non-inverting), with magnitudes from less than 1 to over 8, controlled by drain/source resistor ratios and swamping.
  • Design trade-offs: achieving a target gain requires balancing bias point, transconductance (g_m), drain resistor (R_D), source resistor (R_S), and swamping resistor (R_SW).
  • Common confusion: unswamped vs swamped gain—unswamped gain uses the full transconductance, but swamping (adding R_SW) introduces degeneration to stabilize and reduce gain to a desired level.

📊 Typical parameter values

📊 Input impedance ranges

The excerpt shows Z_in values across multiple problems:

ProblemZ_inNotes
1750 kΩHigh gate resistance
3510 kΩModerate gate resistance
5, 790.9 kΩLower gate resistance
9910 kΩVery high gate resistance
11680 kΩHigh gate resistance
  • MOSFET gates draw negligible DC current, so Z_in is dominated by the gate resistor (R_G).
  • Higher R_G → higher Z_in, but practical limits exist (noise, bias stability).

📊 Voltage gain ranges

Voltage gain (A_v) varies widely:

ProblemA_vOutput voltageInversion
1-2.58Inverted
3-8.86Inverted
5142 mVInverted
7515 mVInverted
9125 mVNot specified
110.91Non-inverting (< 1)
  • Negative A_v indicates phase inversion (common-source configuration).
  • A_v < 1 (problem 11) suggests a source follower or heavily swamped configuration.

🔧 Design example walkthrough

🎯 Design goal (Problem 13)

The excerpt provides a detailed design for a common-source amplifier with specific requirements:

  • Target gain: 5
  • Load resistance (R_L): 10 kΩ
  • Gate resistor (R_G): 510 kΩ or greater
  • Supply voltage: 15 VDC

🧮 Design steps

🧮 Step 1: Check feasibility

  • Maximum transconductance (g_m0) = 25 mS
  • Maximum unswamped gain = g_m0 × R_L = 25 mS × 10 kΩ = 250
  • Target gain of 5 is well below 250, so the design is achievable.

🧮 Step 2: Choose bias point

Mid-point bias: V_GS = 0.5 × V_GS(off) and I_D = 0.25 × I_DSS

  • This choice balances headroom and linearity.
  • Resulting I_D = 6.25 mA (from 0.25 × I_DSS).
  • Resulting g_m = 12.5 mS (half of g_m0 at mid-point bias).

🧮 Step 3: Calculate source resistor (R_S)

  • Total source resistance needed: R_S = V_S / I_D = 1 V / 6.25 mA = 160 Ω
  • This sets the DC operating point.

🧮 Step 4: Choose drain resistor (R_D)

  • Using R_D = 1 kΩ (a common standard value).
  • Effective load resistance: r_L = (R_D || R_L) = (1 kΩ || 10 kΩ) = 909 Ω
  • Drain voltage: V_D = V_DD - I_D × R_D = 15 V - 6.25 mA × 1 kΩ = 8.75 V (reasonable headroom).

🧮 Step 5: Calculate swamping

  • Unswamped gain with r_L = 909 Ω: A_v ≈ g_m × r_L = 12.5 mS × 909 Ω ≈ 11.4
  • This exceeds the target gain of 5, so swamping is needed.
  • Split R_S into two parts:
    • R_S (unbypassed) = 60 Ω
    • R_SW (swamping resistor) = 100 Ω
  • Total = 160 Ω (matches Step 3).

Swamping (source degeneration): adding an unbypassed resistor in the source to reduce gain and improve stability.

  • With R_SW = 100 Ω, the gain drops to approximately 5 (as desired).
  • Example calculation: gain ≈ g_m × r_L / (1 + g_m × R_SW) ≈ 12.5 mS × 909 Ω / (1 + 12.5 mS × 100 Ω) ≈ 11.4 / 2.25 ≈ 5.

🔍 Key design insight

  • Don't confuse: unswamped gain (maximum possible) vs swamped gain (actual, stabilized).
  • Unswamped gain uses the full g_m and is sensitive to parameter variations.
  • Swamping sacrifices gain for predictability and linearity.
  • The 160 Ω total source resistance can be split in "nearly infinite possibilities" (as the excerpt notes), but the ratio of bypassed to unbypassed resistance determines the final gain.

🧩 Interpreting other results

🧩 Output voltage calculations

Problems 5, 7, and 9 give output voltages for specific input conditions:

  • Problem 5: V_out = 142 mV (inverted), Z_in = 90.9 kΩ
  • Problem 7: V_out = 515 mV (inverted), Z_in = 90.9 kΩ
  • Problem 9: V_out = 125 mV, Z_in = 910 kΩ

These show how different input signals and circuit parameters produce different output amplitudes.

🧩 Non-inverting configuration (Problem 11)

  • A_v = 0.91 (less than 1, non-inverting)
  • Z_in = 680 kΩ
  • This likely represents a source follower (common-drain) or a heavily degenerated common-source stage.
  • Gain less than 1 indicates voltage buffering rather than amplification.

⚠️ Context and limitations

⚠️ What the excerpt does not contain

  • The excerpt is a list of numerical answers to end-of-chapter problems.
  • It does not explain the underlying theory, circuit topologies, or derivation of formulas.
  • Problem statements and circuit diagrams are not included.
  • The design example (Problem 13) is the only substantive explanation provided.

⚠️ How to use these notes

  • These notes extract the key numerical patterns and the one detailed design walkthrough.
  • To fully understand MOSFET common-source amplifiers, refer to the main chapter text (not provided here).
  • The design example illustrates the iterative process: check feasibility → choose bias → calculate resistors → adjust for target gain.
90

13.3 MOSFET Common Drain Followers

13.3 MOSFET Common Drain Followers

🧭 Overview

🧠 One-sentence thesis

The excerpt provides numerical answers to practice problems for MOSFET common drain follower circuits, demonstrating typical design calculations involving input impedance, voltage gain, and component selection.

📌 Key points (3–5)

  • Input impedance values: Common drain configurations show high input impedances in the hundreds of kiloohms to megohm range.
  • Voltage gain characteristics: Gains are close to unity (around 0.91) and non-inverting, typical of follower circuits.
  • Design flexibility: Multiple valid solutions exist for meeting specified performance targets using different component combinations.
  • Common confusion: Unlike common source amplifiers (which invert and have higher gain), common drain followers have near-unity gain and do not invert the signal.

📊 Problem solutions and typical values

📊 Input impedance characteristics

The excerpt shows several solved problems with input impedance results:

  • Problem 9: Z_in = 910 kΩ, V_out = 125 mV
  • Problem 11: Z_in = 680 kΩ, A_v = 0.91

These high input impedances (hundreds of kiloohms) are characteristic of MOSFET circuits due to the gate's insulating properties.

📈 Voltage gain behavior

Voltage gain (A_v) in common drain followers: typically close to but less than 1.0, non-inverting.

From Problem 11: A_v = 0.91

  • This near-unity gain is the defining feature of follower circuits.
  • The output "follows" the input with minimal attenuation.
  • No phase inversion occurs (unlike common source configurations).

🔧 Design example walkthrough

🔧 Problem 13 design scenario

The excerpt provides a detailed design solution for achieving a specific gain target:

Given constraints:

  • Target gain achievable because g_m0 = 25 mS and R_L = 10 kΩ
  • Maximum theoretical gain ceiling: 250 (unswamped, unrealistic)
  • Supply voltage: 15 VDC

Design choices made:

  • Mid-point bias selected: V_GS = 0.5 V_GS(off) and I_D = 0.25 I_DSS
  • This yields R_S = 1 V / 6.25 mA = 160 Ω
  • Transconductance: g_m = 12.5 mS
  • Drain resistor: R_D = 1 kΩ (common standard value)
  • Results in r_L = 909 Ω and V_D = 8.75 V (reasonable operating point)

⚙️ Gain adjustment technique

The design uses source degeneration to control gain:

  • Unswamped gain exceeds 11
  • Split the 160 Ω source resistance into two parts: R_S and R_SW
  • R_SW adds degeneration to position gain at desired level
  • Example: R_SW = 100 Ω with R_S = 60 Ω yields a gain of 5

Key insight: The excerpt notes "nearly infinite possibilities" for meeting the design goal, showing that multiple component combinations can satisfy the same performance specification.

🎯 R_G selection guideline

  • R_G = 510 kΩ or greater recommended
  • The gate resistor is "not tied to any particular value but probably > 1 MΩ" (from Chapter 12, Problem 19)
  • High values maintain the high input impedance advantage of MOSFET circuits
91

13.4 MOSFET Common Gate Amplifiers

13.4 MOSFET Common Gate Amplifiers

🧭 Overview

🧠 One-sentence thesis

The excerpt provides numerical answers to practice problems for MOSFET common gate amplifiers and related circuit analysis, but does not contain explanatory content about the amplifier configuration itself.

📌 Key points (3–5)

  • What the excerpt contains: numerical solutions to end-of-chapter problems for Chapters 12–17, including DC bias calculations, AC parameters, and frequency response values.
  • Chapter 13 relevance: problems 1–13 under "Chapter 13" show input impedance, voltage gain, and output voltage calculations for MOSFET amplifier circuits.
  • Design problem example: problem 13 in Chapter 13 presents a complete design scenario with constraints and multiple valid solutions.
  • Common confusion: the excerpt is a solutions appendix, not instructional text—it lists answers without explaining the underlying theory or derivation steps.
  • Missing content: no definitions, circuit diagrams, or conceptual explanations of common gate amplifier operation are present.

📋 Structure of the excerpt

📋 Answer key format

  • The excerpt is organized by chapter number (Chapters 9, 12–17).
  • Each chapter lists numbered problems followed by their numerical answers.
  • Answers include circuit parameters such as:
    • Input impedance (Z_in)
    • Voltage gain (A_v)
    • Output voltage (V_out)
    • DC operating points (I_D, V_G, V_D)

🔢 Chapter 13 problems

The problems most relevant to "13.4 MOSFET Common Gate Amplifiers" are numbered 1–13 under the Chapter 13 heading:

ProblemGiven answersNotes
1Z_in = 750 kΩ, A_v = -2.58Negative gain indicates inversion
3Z_in = 510 kΩ, A_v = -8.86Higher magnitude gain than problem 1
5Z_in = 90.9 kΩ, V_out = 142 mV (inverted)Lower input impedance
7Z_in = 90.9 kΩ, V_out = 515 mV (inverted)Same Z_in, different output
9Z_in = 910 kΩ, V_out = 125 mVNo inversion noted
11Z_in = 680 kΩ, A_v = 0.91Gain less than 1 (buffer-like)
13Design problemSee detailed breakdown below

🛠️ Design problem walkthrough

🛠️ Problem 13 scenario

The excerpt states:

"R_G = 510 kΩ or greater. The remainder of the design has nearly infinite possibilities."

  • Goal constraint: the design must be achievable because g_m0 = 25 mS and R_L = 10 kΩ, giving a ceiling A_v of 250 unswamped (but unrealistic).
  • Supply and bias choices: using a standard 15 VDC supply and picking a mid-point bias (V_GS = 0.5 V_GS(off) and I_D = 0.25 I_DSS).

🔧 Component selection

The solution describes a step-by-step design process:

  1. Source resistor: R_S = 1 V / 6.25 mA = 160 Ω, with g_m = 12.5 mS.
  2. Drain resistor: using a common 1 kΩ for R_D yields r_L = 909 Ω and V_D = 8.75 V (reasonable).
  3. Unswamped gain: the resulting unswamped gain is over 11.
  4. Gain adjustment: the 160 Ω source resistance can be split into two parts, R_S and R_SW, with R_SW adding enough degeneration to position the gain at the desired level.
    • Example: R_SW = 100 Ω with R_S = 60 Ω yields a gain of 5.

🎯 Design flexibility

  • The excerpt emphasizes "nearly infinite possibilities" for the design, meaning multiple component combinations can meet the specifications.
  • The key is balancing bias point, gain, and practical component values.
  • Don't confuse: "unswamped" gain refers to the theoretical maximum without source degeneration; adding R_SW reduces gain to a controlled, stable value.

⚠️ Limitations of this excerpt

⚠️ No conceptual content

  • The excerpt does not explain what a common gate amplifier is or how it works.
  • No circuit topology, signal flow, or operating principles are described.
  • It assumes the reader already understands the theory and is checking numerical answers.

⚠️ Context from other chapters

  • Chapters 12–17 cover related topics (DC bias, frequency response, decibel calculations) but are not explained here.
  • The excerpt is purely a reference for verifying problem solutions, not a learning resource for new concepts.
92

Chapter 14: Class D Power Amplifiers

Chapter 14: Class D Power Amplifiers

🧭 Overview

🧠 One-sentence thesis

The excerpt provides only numerical answers to end-of-chapter problems for Chapter 14, without presenting any conceptual content about Class D power amplifiers.

📌 Key points (3–5)

  • The excerpt contains only three numerical answers labeled as Chapter 14 problems.
  • No definitions, explanations, or theory about Class D power amplifiers are present.
  • The answers include a frequency value (35 kHz), a slew rate (33.3 V/μs), and current/voltage values (12.48 A, 0.15 V).
  • The excerpt lacks substantive content for creating meaningful review notes about the chapter topic.

📋 Content limitations

📋 What the excerpt contains

The source material for Chapter 14 includes only three problem answers:

Problem numberAnswer provided
135 kHz
333.3 V/μs
512.48 A, 0.15 V

⚠️ Missing information

  • No conceptual explanations of Class D power amplifiers
  • No definitions or operating principles
  • No diagrams, circuits, or design considerations
  • No context for what the numerical answers represent
  • No theory or mechanisms that would enable review or study

Note: The excerpt appears to be from an answer key or solutions section at the end of a textbook, containing only final numerical results without the accompanying chapter content, problems, or explanatory material.

93

Chapter 14 Answers

14.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

This excerpt provides numerical answers to Chapter 14 exercises, showing calculated values for frequency, slew rate, and current/voltage parameters in electronic circuits.

📌 Key points (3–5)

  • What is provided: numerical solutions to three problems from Chapter 14.
  • Types of quantities: frequency (kHz), slew rate (V/μs), and current/voltage pairs (A, V).
  • No explanatory content: the excerpt contains only final answers without derivations, context, or theory.
  • Common confusion: this is an answer key, not instructional material—it does not explain how to arrive at these values or what the problems asked.

📋 Answer key content

📋 Problem solutions listed

The excerpt lists three answers:

ProblemQuantityValue
1Frequency35 kHz
3Slew rate33.3 V/μs
5Current and voltage12.48 A, 0.15 V
  • Problem 1: the answer is a frequency of 35 kilohertz.
  • Problem 3: the answer is a slew rate of 33.3 volts per microsecond (rate of voltage change over time).
  • Problem 5: the answer includes two values—a current of 12.48 amperes and a voltage of 0.15 volts.

⚠️ Limitations of this excerpt

  • No problem statements: the original questions are not included, so the context for each answer is unknown.
  • No working or explanation: there are no steps, formulas, or reasoning shown.
  • No definitions or theory: the excerpt does not explain what slew rate is, why frequency matters, or how current and voltage relate in the circuit.
  • Example: without the problem statement, we cannot know whether "35 kHz" is a cutoff frequency, resonant frequency, or bandwidth.

🔍 What can be inferred

🔍 Likely topics in Chapter 14

Based on the types of answers:

  • Frequency analysis: Problem 1 suggests the chapter covers frequency-related calculations (filters, oscillators, or bandwidth).
  • Slew rate: Problem 3 indicates operational amplifier or signal processing topics, since slew rate measures how fast an output voltage can change.
  • Current and voltage: Problem 5 likely involves DC or AC circuit analysis, possibly power calculations or component ratings.

🔍 Use of this excerpt

  • For self-study: this excerpt is useful only if you have already worked the problems and want to check your final answers.
  • Not for learning: without the problem statements or solution steps, you cannot learn the underlying concepts or methods from this excerpt alone.
  • Don't confuse: an answer key with a worked example—this excerpt provides only the final numbers, not the process.
94

Introduction

14.1 Introduction

🧭 Overview

🧠 One-sentence thesis

This excerpt provides numerical answers to end-of-chapter problems for Chapters 12 through 17, covering transistor circuit analysis, frequency response, and amplifier design.

📌 Key points (3–5)

  • Content type: The excerpt consists entirely of problem answers (numerical values and brief design notes) rather than explanatory text.
  • Chapters covered: Chapters 12–17, spanning FET circuits, amplifier parameters, frequency response, and decibel calculations.
  • Format: Most answers are parameter values (currents, voltages, impedances, gains, frequencies, phases) with units.
  • No conceptual content: The excerpt contains no definitions, explanations, derivations, or teaching material—only solution data.

📋 What this excerpt contains

📋 Answer key structure

The excerpt is formatted as a solutions appendix:

  • Each chapter heading (e.g., "Chapter 12") is followed by numbered problem answers.
  • Answers include circuit parameters, design choices, and calculated values.
  • Some answers include brief reasoning (e.g., "due to proximity").

🔢 Types of answers provided

ChapterTypical answer content
12FET DC operating points: drain current (I_D), gate voltage (V_G), drain voltage (V_D), resistor values
13AC amplifier parameters: input impedance (Z_in), voltage gain (A_v), output voltage (V_out), design trade-offs
14Frequency and slew rate values
15Voltage and time values
16Decibel conversions, power/voltage gains, frequency response (amplitude and phase at specific frequencies)
17Critical frequencies (f_in, f_out, f_bypass, f_1, f_2) with proximity notes

🛠️ Example design problem (Chapter 13, Problem 13)

One answer includes design reasoning:

  • Goal: achieve a specific voltage gain with given constraints.
  • Approach: choose bias point, calculate transconductance (g_m), select resistor values, use source degeneration to set gain.
  • Example values: R_G = 510 kΩ or greater, R_S split into R_S and R_SW to position gain at desired level (e.g., R_SW = 100 Ω, R_S = 60 Ω yields gain of 5).

Don't confuse: This is a worked solution summary, not a general design method—specific component values depend on the problem's given parameters.

⚠️ Limitations of this excerpt

⚠️ No substantive teaching content

  • The excerpt does not define terms (e.g., what "input impedance" or "voltage gain" mean).
  • It does not explain how to calculate the given values.
  • It does not discuss underlying principles, circuit operation, or design trade-offs beyond one brief example.

⚠️ Context required

  • To understand these answers, the reader must refer to the original problem statements (not included in the excerpt).
  • The excerpt assumes familiarity with circuit analysis notation and units (e.g., mA, kΩ, dB, degrees of phase).

Conclusion: This excerpt serves as a reference for checking problem solutions, not as a study resource for learning concepts or methods.

95

14.2 Class D Basics

14.2 Class D Basics

🧭 Overview

🧠 One-sentence thesis

The excerpt provided contains only numerical answers to end-of-chapter problems from Chapters 12–17 and does not present substantive conceptual content about Class D amplifiers or any other technical topic.

📌 Key points (3–5)

  • The excerpt consists entirely of problem solutions (numerical values with units).
  • No definitions, explanations, mechanisms, or theoretical content are present.
  • The answers span topics including FET circuits (Chapter 12–13), frequency response (Chapter 16–17), and decibel calculations (Chapter 16).
  • The title "14.2 Class D Basics" does not correspond to any content in the excerpt; Chapter 14 answers are minimal (only three entries).
  • No conceptual material is available to extract for review or study purposes.

📋 Content summary

📋 What the excerpt contains

The source text is an answer key for textbook problems across multiple chapters:

  • Chapter 12: FET DC operating points (drain current I_D, gate voltage V_G, drain voltage V_D) and resistor values.
  • Chapter 13: Input impedance Z_in, voltage gain A_v, and output voltage V_out for amplifier circuits; one detailed design problem (#13).
  • Chapter 14: Only three answers (35 kHz, 33.3 V/μs, and current/voltage values).
  • Chapter 15: Time values (4 V, 30 ns, 610 ns).
  • Chapter 16: Decibel conversions (dB, dBW, dBf, dBV, dBm), gain calculations, phase angles, and frequency response.
  • Chapter 17: Frequency calculations (f_in, f_out, f_bypass, f_1, f_2) with proximity notes.

⚠️ No Class D amplifier content

  • The title "14.2 Class D Basics" suggests the section should introduce Class D amplifier principles (switching amplifiers, pulse-width modulation, efficiency, etc.).
  • The excerpt does not contain any such material.
  • Chapter 14 answers are unrelated to amplifier topology and provide no context.

🔍 Notable problem types

🔍 Chapter 13 design problem

Problem #13 is the only answer with explanatory text:

  • Goal: Achieve a specific voltage gain using a FET amplifier with given transconductance g_m0 = 25 mS and load resistance R_L = 10 kΩ.
  • Approach: Choose mid-point bias (V_GS = 0.5 V_GS(off), I_D = 0.25 I_DSS), calculate source resistance R_S = 160 Ω, and split it into R_S and R_SW (swamping resistor) to set gain.
  • Example solution: R_SW = 100 Ω, R_S = 60 Ω yields gain of 5.
  • This is the only answer that describes a design process rather than a final numerical result.

🔍 Frequency response (Chapter 16–17)

  • Answers include breakpoint frequencies, phase shifts, and gain in decibels at specific frequencies.
  • Notes like "due to proximity" indicate that the dominant pole is affected by nearby poles.
  • Example: "f_1 > 67 Hz due to proximity" means the lower cutoff frequency is higher than the calculated input breakpoint because other breakpoints are close.

❌ Limitations for study purposes

❌ No conceptual framework

  • Definitions, principles, and explanations are absent.
  • A student cannot learn "Class D Basics" or any other topic from numerical answers alone.
  • The excerpt assumes the reader has already studied the theory and worked through the problems.

❌ Context missing

  • Problem statements are not included, so the meaning of each answer is unclear.
  • Example: "I_D = 3 mA, V_G = 0 V, V_D = 9.6 V" (Chapter 12, #1) does not indicate the circuit topology, component values, or what was being solved.

Conclusion: This excerpt is an answer key appendix and does not contain material suitable for creating review notes on "Class D Basics" or any conceptual topic. To study Class D amplifiers, refer to the main text of section 14.2, not the end-of-book solutions.

96

Pulse Width Modulation

14.3 Pulse Width Modulation

🧭 Overview

🧠 One-sentence thesis

The excerpt provides numerical answers to end-of-chapter problems for Chapters 12–17 but does not contain substantive content explaining pulse width modulation concepts or mechanisms.

📌 Key points (3–5)

  • The excerpt consists entirely of problem answers (numerical values, component values, and brief design notes).
  • No definitions, explanations, or conceptual material about pulse width modulation are present.
  • Chapter 14 answers include only three items: a frequency (35 kHz), a slew rate (33.3 V/μs), and current/voltage values (12.48 A, 0.15 V).
  • The excerpt does not explain what pulse width modulation is, how it works, or why it matters.
  • The surrounding chapters (12, 13, 15–17) cover transistor circuits, amplifiers, frequency response, and decibel calculations, but no PWM theory is included.

📋 Content summary

📋 What the excerpt contains

The source text is an answer key for multiple textbook chapters:

  • Chapter 12: transistor (FET) DC operating points—drain current, gate voltage, drain voltage.
  • Chapter 13: amplifier input impedance, voltage gain, and a detailed design problem for a FET amplifier.
  • Chapter 14: three numerical answers only (35 kHz, 33.3 V/μs, 12.48 A and 0.15 V).
  • Chapters 15–17: timing values, decibel conversions, frequency response, phase angles, and bandwidth calculations.

❌ What is missing

  • No prose explanation of pulse width modulation.
  • No definitions, diagrams, or conceptual discussion.
  • No description of PWM waveforms, duty cycle, switching, or applications.
  • The title "14.3 Pulse Width Modulation" suggests a textbook section, but the excerpt contains only answer-key data.

🔢 Chapter 14 answers

🔢 The three provided answers

The excerpt lists:

  1. 35 kHz (problem 1)
  2. 33.3 V/μs (problem 3)
  3. 12.48 A, 0.15 V (problem 5)
  • These values likely correspond to frequency, slew rate, and current/voltage in a PWM or power-electronics context.
  • Without the problem statements, the meaning and context of these numbers cannot be determined from the excerpt alone.

⚠️ No explanatory content

  • The excerpt does not state what these values represent in a PWM system.
  • No formulas, derivations, or conceptual links are provided.
  • The answers are isolated numerical results with no supporting theory.

📌 Conclusion

📌 Limitations of the excerpt

  • The excerpt is an answer key, not a teaching text.
  • It does not fulfill the goal of explaining pulse width modulation for review or self-study.
  • To understand PWM, a learner would need the corresponding textbook section (not included here) or external resources.
97

14.4 Output Configurations

14.4 Output Configurations

🧭 Overview

🧠 One-sentence thesis

This excerpt provides numerical answers to end-of-chapter problems across multiple chapters (12–17) covering transistor biasing, amplifier parameters, frequency response, and decibel calculations, serving as a solution key rather than instructional content.

📌 Key points (3–5)

  • What this excerpt is: a collection of numerical answers and brief solution sketches for textbook problems, not a teaching section.
  • Chapters covered: answers span Chapters 12 (FET biasing), 13 (FET amplifiers), 14 (output configurations), 15, 16 (decibels and gain), and 17 (frequency response).
  • Type of information: mostly final numerical values (voltages, currents, impedances, gains, frequencies, phase angles) with occasional design reasoning.
  • Common confusion: this is not a conceptual explanation of "output configurations"—it is a problem-answer appendix; the title "14.4 Output Configurations" likely refers to a textbook section, but the excerpt contains only problem solutions.
  • Limitation: without the original problem statements, the answers alone do not teach the underlying concepts.

📋 Structure and scope

📋 What the excerpt contains

  • The excerpt is formatted as a list of numbered problems under chapter headings (e.g., "Chapter 12," "Chapter 13").
  • Each line gives a problem number followed by calculated results: component values, voltages, currents, impedances, gains, frequencies, and phase angles.
  • A few problems include brief design notes (e.g., problem 13 in Chapter 13 explains a multi-step FET amplifier design process).

🔢 Chapters and topics

ChapterTopic inferred from answers
12FET DC biasing (drain current I_D, gate voltage V_G, drain voltage V_D, resistor values)
13FET amplifier AC analysis (input impedance Z_in, voltage gain A_v, output voltage V_out)
14Output configurations (only two answers: a frequency and a slew rate)
15Brief answers (voltages and times, likely transient or switching topics)
16Decibel conversions and cascaded gain (dB, dBW, dBf, dBV, power, phase)
17Frequency response (input/output/bypass break frequencies f_in, f_out, f_bypass, and dominant pole f_1 or f_2)
  • The excerpt does not explain how to solve these problems; it only states the final answers.

🔌 Chapter 12: FET biasing answers

🔌 Typical answer format

  • Each problem gives three values: drain current I_D, gate voltage V_G, and drain voltage V_D.
  • Example: "1. I_D = 3 mA, V_G = 0 V, V_D = 9.6 V" means the DC operating point has 3 milliamps through the drain, 0 volts at the gate, and 9.6 volts at the drain.
  • Problems 17 and 19 also include resistor values (e.g., R_D = 1.29 kΩ, R_G > 1 MΩ).

🧮 What these answers represent

  • I_D: the quiescent (DC) drain current, set by the bias network.
  • V_G: the DC gate voltage (often 0 V for self-bias or source-biased JFETs).
  • V_D: the DC drain voltage relative to ground (or a negative supply in some cases, e.g., problem 11: V_D = −17.2 V).
  • These values confirm that the transistor is biased in the active (saturation) region for amplification.

🎚️ Chapter 13: FET amplifier AC parameters

🎚️ Input impedance and gain

  • Answers give Z_in (input impedance, typically hundreds of kΩ for FET amplifiers) and A_v (voltage gain, often negative for common-source configurations).
  • Example: "1. Z_in = 750 kΩ, A_v = −2.58" means the amplifier presents 750 kΩ to the source and inverts the signal with a gain magnitude of 2.58.
  • Problems 5 and 7 note "(inverted)" to clarify phase inversion.

🎚️ Design problem (13.13)

  • Problem 13 is a longer design exercise:
    • Goal: achieve a specific voltage gain (5 in the example) using a JFET with given transconductance g_m0 = 25 mS and load R_L = 10 kΩ.
    • Steps outlined:
      1. Check feasibility: unswamped ceiling gain = g_m0 × R_L = 250 (achievable).
      2. Choose bias point: mid-point bias (V_GS = 0.5 V_GS(off), I_D = 0.25 I_DSS) yields g_m = 12.5 mS.
      3. Select R_D = 1 kΩ, giving effective load r_L = 909 Ω and V_D = 8.75 V (reasonable headroom).
      4. Split source resistance R_S = 160 Ω into R_S = 60 Ω and swamping resistor R_SW = 100 Ω to add degeneration and set gain to 5.
    • This is the only problem with detailed reasoning; it shows how to trade off gain, bias stability, and component values.

📶 Chapter 14: output configurations (minimal data)

📶 What is provided

  • Only two answers:
    • "1. 35 kHz" (likely a bandwidth or cutoff frequency).
    • "3. 33.3 V/μs" (a slew rate, the maximum rate of voltage change at the output).
  • Problem 5 gives current and voltage: "12.48 A, 0.15 V" (possibly output current capability and voltage drop).

📶 What is missing

  • The excerpt does not describe the output configuration types (e.g., common-drain/source follower, push-pull, class A/B/C).
  • Without the problem statements, these numbers are isolated data points.

📊 Chapter 16: decibels and cascaded stages

📊 Decibel conversions

  • Problems 1, 7, 11, 13 convert between linear and logarithmic units:
    • dB (power ratio), dBW (power relative to 1 watt), dBV (voltage relative to 1 volt), dBf (power relative to 1 femtowatt).
    • Example: "1. A) 10 dB B) 19 dB C) 26.99 dB D) 0 dB E) −6.99 dB F) −15.23 dB" lists gains or attenuations in dB for different scenarios.
  • Problem 9 shows both linear gain (A = 8.57) and dB gain (A' = 18.66 dB) for the same amplifier.

📊 Cascaded gain and stage-by-stage analysis

  • Problems 15, 17, 21 calculate total gain and intermediate outputs for multi-stage amplifiers.
  • Example (problem 17): for an input of 4 dBm, stage 1 outputs 6 dBm, stage 2 outputs 0 dBm, stage 3 outputs 15 dBm; for −34 dBm input, outputs are −24 dBW, −30 dBW, −15 dBW.
  • Key idea: in dB, gains add (G'_total = G'_1 + G'_2 + G'_3), making cascaded analysis straightforward.

📊 Frequency-dependent gain and phase

  • Problems 25, 27, 29, 31, 33, 35, 37 include phase angles and gain at specific frequencies.
  • Example (problem 25): at 50 kHz, gain is −0.022 dB and phase is −4.09°; at 700 kHz, −3 dB and −45°; at 10 MHz, −23.1 dB and −86°.
  • These answers reflect the amplifier's frequency response (roll-off and phase shift due to capacitive/inductive elements).
  • Problem 27 notes "amplitude portion does not change" but phases vary with frequency (e.g., −188.5° at 30 kHz, −225° at 200 kHz).

🔊 Chapter 17: frequency response break points

🔊 Input, output, and bypass frequencies

  • Each problem lists f_in (input coupling break frequency), f_out (output coupling break frequency), and f_bypass (bypass capacitor break frequency).
  • Example (problem 1): f_in = 67 Hz, f_out = 2.65 Hz, f_bypass = 61.2 Hz.
  • These are the −3 dB points (corner frequencies) where each capacitor begins to roll off the gain.

🔊 Dominant pole and proximity effects

  • The answers identify the dominant pole f_1 (low-frequency cutoff) or f_2 (high-frequency cutoff).
  • Proximity effect: when two break frequencies are close, the dominant pole is higher (or lower) than the highest (or lowest) individual break frequency.
    • Example (problem 1): "f_1 > 67 Hz due to proximity" means the actual low-frequency cutoff is above 67 Hz because f_in and f_bypass are close (67 Hz and 61.2 Hz).
    • Example (problem 3): "f_1 > 9.76 Hz due to proximity."
    • Example (problem 5): "f_1 ≈ 477 Hz" (no proximity issue; f_in = 477 Hz dominates).
  • Don't confuse: the dominant pole is not always the highest individual break frequency; interaction between nearby poles shifts the effective cutoff.

🔊 High-frequency analysis

  • Problems 17, 19, 21, 23 give f_in and f_out in MHz, representing high-frequency roll-off (due to transistor capacitances and stray capacitance).
  • Example (problem 17): f_in = 4.78 MHz, f_out = 13.6 MHz; "f_2 < 4.78 Hz due to proximity" (likely a typo; should be MHz).
  • The same proximity logic applies: when two high-frequency poles are close, the effective f_2 is lower than the lowest individual pole.

🛠️ Limitations of this excerpt

🛠️ What is absent

  • Problem statements: the excerpt does not include the original questions, circuit diagrams, or component values given in the problems.
  • Solution steps: most answers are final numbers with no derivation (except problem 13 in Chapter 13).
  • Conceptual explanations: there is no discussion of why these values matter or how the circuits work.

🛠️ How to use these answers

  • These are intended as a solution key for students checking their work.
  • To learn the material, one must refer to the corresponding textbook sections and problem statements.
  • Example: knowing "Z_in = 750 kΩ" is useful only if you understand what input impedance is and how it affects circuit loading.
98

Chapter 15: Insulated Gate Bipolar Transistors (IGBTs)

Chapter 15: Insulated Gate Bipolar Transistors (IGBTs)

🧭 Overview

🧠 One-sentence thesis

The excerpt provides only numerical answers to Chapter 15 problems without explanatory content, making it impossible to extract substantive concepts about IGBTs.

📌 Key points (3–5)

  • Content limitation: The excerpt contains only three numerical answers (4 V, 30 ns, 610 ns) for Chapter 15.
  • No context provided: No problem statements, circuit descriptions, or explanations accompany the answers.
  • Answer format: The answers appear to be voltage (V) and time (ns) measurements.
  • Common confusion: This excerpt is from an answer key, not instructional material—it cannot be used to learn IGBT concepts.
  • What is missing: Definitions, operating principles, applications, or any explanatory text about IGBTs.

📋 Excerpt content

📋 Available data

The excerpt lists only three answers under "Chapter 15":

Problem numberAnswer
14 V
330 ns
5610 ns

⚠️ What cannot be determined

  • What the problems asked: No problem statements are included.
  • What the values represent: The 4 V could be a voltage threshold, gate voltage, or any other parameter; the time values (30 ns, 610 ns) could be switching times, delays, or other timing parameters.
  • IGBT concepts: No definitions, mechanisms, comparisons, or explanations of IGBT operation are present.
  • Context: Whether these are turn-on times, turn-off times, voltage ratings, or other specifications is unknown.

🚫 Substantive content assessment

🚫 Why review notes cannot be generated

  • The excerpt is purely an answer key with numerical results.
  • No instructional text, definitions, or conceptual explanations exist in the provided material.
  • Creating review notes would require inventing facts not present in the excerpt, which violates the content requirements.

📝 What would be needed

To create meaningful review notes about IGBTs, the excerpt would need to include:

  • Definitions of IGBT structure and operation
  • Explanations of switching behavior, voltage/current characteristics
  • Comparisons with other transistor types (MOSFETs, BJTs)
  • Application contexts or design considerations
  • Problem statements that reveal what concepts are being tested
99

Answer Key Excerpts (Chapters 12–17)

15.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

This excerpt contains numerical answers to end-of-chapter problems for Chapters 12 through 17, covering transistor circuit analysis, amplifier design, frequency response, and decibel calculations.

📌 Key points (3–5)

  • What this excerpt is: answer key entries for problem sets, not explanatory text or theory.
  • Chapters covered: Chapters 12 (FET DC analysis), 13 (FET amplifiers), 14 (slew rate and current), 15 (timing), 16 (decibels and gain), and 17 (frequency response and cutoff frequencies).
  • Type of content: numerical results (voltages, currents, impedances, gains, frequencies, phases) and occasional design reasoning.
  • Common confusion: this is not a teaching section—it provides final answers and occasional design notes, not step-by-step derivations.
  • Limitation: without the original problem statements, the answers alone do not convey the underlying concepts or methods.

📐 Chapter 12: FET DC operating points

📐 Drain current, gate voltage, and drain voltage

  • Each answer gives the DC operating point of a field-effect transistor circuit: drain current I<sub>D</sub>, gate voltage V<sub>G</sub>, and drain voltage V<sub>D</sub>.
  • Example: Problem 1 yields I<sub>D</sub> = 3 mA, V<sub>G</sub> = 0 V, V<sub>D</sub> = 9.6 V.
  • Problems 11 and 13 show negative drain voltages (−17.2 V, −12.9 V), indicating circuits with negative supply rails or different configurations.

🔧 Resistor design

  • Problem 17: a single resistance value (180 Ω).
  • Problem 19: R<sub>D</sub> = 1.29 kΩ; R<sub>G</sub> is not tied to a specific value but is probably greater than 1 MΩ (high input impedance).

🔊 Chapter 13: FET amplifier parameters

🔊 Input impedance and voltage gain

  • Answers list input impedance Z<sub>in</sub> and voltage gain A<sub>v</sub> (or output voltage V<sub>out</sub>).
  • Example: Problem 1 gives Z<sub>in</sub> = 750 kΩ, A<sub>v</sub> = −2.58 (negative sign indicates inversion).
  • Problems 5 and 7 note "(inverted)" explicitly.

🛠️ Design problem (Problem 13)

  • Goal: achieve a specific gain with g<sub>m0</sub> = 25 mS and R<sub>L</sub> = 10 kΩ.
  • The answer outlines a design strategy:
    • Maximum unswamped gain ceiling is 250 (unrealistic).
    • Choose mid-point bias: V<sub>GS</sub> = 0.5 V<sub>GS(off)</sub>, I<sub>D</sub> = 0.25 I<sub>DSS</sub>.
    • This yields R<sub>S</sub> = 160 Ω and g<sub>m</sub> = 12.5 mS.
    • Using R<sub>D</sub> = 1 kΩ gives r<sub>L</sub> = 909 Ω and V<sub>D</sub> = 8.75 V.
    • Split R<sub>S</sub> into R<sub>S</sub> and R<sub>SW</sub> (swamping resistor) to set gain; example: R<sub>SW</sub> = 100 Ω, R<sub>S</sub> = 60 Ω yields gain of 5.
  • Don't confuse: "unswamped" gain is the theoretical maximum without source degeneration; adding R<sub>SW</sub> reduces gain to the desired level.

⚡ Chapters 14 and 15: Timing and transient parameters

⚡ Chapter 14 answers

  • Problem 1: 35 kHz (a frequency).
  • Problem 3: 33.3 V/μs (slew rate).
  • Problem 5: 12.48 A, 0.15 V (current and voltage).

⏱️ Chapter 15 answers

  • Problem 1: 4 V.
  • Problem 3: 30 ns (nanoseconds, a time delay).
  • Problem 5: 610 ns.

📊 Chapter 16: Decibels, gain, and power

📊 Decibel conversions and gain calculations

  • Problem 1: six sub-answers in dB (10 dB, 19 dB, 26.99 dB, 0 dB, −6.99 dB, −15.23 dB).
  • Problem 3: 33 dB.
  • Problem 5: gain G = 501, output power P<sub>out</sub> = 12.53 W.
  • Problem 7: six linear gain values (1.06, 1, 199.5, 3.43, 0.398, 0.188).
  • Problem 9: A = 8.57, A′ = 18.66 dB (linear and decibel forms of the same gain).

📡 Power in dBW and dBf

  • Problem 11: nine sub-answers in dBW (0 dBW, 13.6 dBW, 8.13 dBW, −7 dBW, −26.4 dBW, 30.8 dBW, −43.5 dBW, −65.2 dBW, −172.5 dBW).
  • Problem 13: nine sub-answers in dBf (150 dBf, 163.6 dBf, 158.1 dBf, 143 dBf, 123.6 dBf, 180.8 dBf, 106.5 dBf, 84.8 dBf, −22.5 dBf).
  • Don't confuse: dBW is referenced to 1 watt; dBf is referenced to 1 femtowatt (10<sup>−15</sup> W).

🔗 Cascaded stages

  • Problem 15: total gain in dB (G<sub>total</sub> = 28 dB) and linear (G = 631).
  • Problem 17: output levels for each of three stages, given two different input levels (4 dBm and −34 dBm).
  • Problem 21: final output V<sub>out</sub> = 21 dBV = 11.2 V; intermediate stages: stage 1 output 4 dBV = 1.58 V, stage 2 output 9 dBV = 2.82 V.

📈 Frequency and phase response

  • Problem 25: at three frequencies (50 kHz, 700 kHz, 10 MHz), amplitude in dB (−0.022 dB, −3 dB, −23.1 dB) and phase in degrees (−4.09°, −45°, −86°); rise time T<sub>r</sub> = 500 μs.
  • Problem 27: amplitude unchanged; phases at 30 kHz (−188.5°), 200 kHz (−225°), 1 MHz (−258.7°).
  • Problem 31: phases at 4 kHz (78.7°), 20 Hz (45°), 100 Hz (11.3°).
  • Problem 35: net gain and phase at multiple frequencies (20 kHz: 35.87 dB, 51.7°; 100 kHz: 40 dB, −5.1°; 800 kHz: 30.5 dB, −70.5°).

🔢 Miscellaneous Chapter 16 answers

  • Problem 19: 200 mW.
  • Problem 23: 15 V.
  • Problem 43: 0.775 V.
  • Problem 45: 360 W.
  • Problem 47: 71.5 dBV.
  • Problem 49: greater than 30 Hz.

🎚️ Chapter 17: Frequency response and cutoff frequencies

🎚️ Input, output, and bypass cutoff frequencies

  • Each problem lists f<sub>in</sub>, f<sub>out</sub>, and (when applicable) f<sub>bypass</sub>, plus an estimate of the lower cutoff frequency f<sub>1</sub>.
  • Example (Problem 1): f<sub>in</sub> = 67 Hz, f<sub>out</sub> = 2.65 Hz, f<sub>bypass</sub> = 61.2 Hz; f<sub>1</sub> > 67 Hz due to proximity.
  • "Due to proximity" means that when two or more cutoff frequencies are close together, the actual overall cutoff is higher (or lower, for high-frequency cutoffs) than the highest (or lowest) individual frequency.

🔼 Lower cutoff frequency f<sub>1</sub>

Problemf<sub>in</sub>f<sub>out</sub>f<sub>bypass</sub>f<sub>1</sub> estimate
167 Hz2.65 Hz61.2 Hz> 67 Hz (proximity)
39.76 Hz0.796 Hz2.57 Hz> 9.76 Hz (proximity)
5477 Hz8.8 Hz7.42 Hz≈ 477 Hz
74.5 Hz2.6 Hz69.4 Hz≈ 69.4 Hz
94.5 Hz67.9 Hz≈ 67.9 Hz
110.34 Hz0.723 Hz16.3 Hz≈ 16.3 Hz
131.59 Hz3.22 Hz0.781 Hz> 3.22 Hz (proximity)
157.23 Hz0.672 Hz≈ 7.23 Hz
  • When f<sub>1</sub> is "≈" one of the frequencies, that frequency dominates; when ">" with "proximity," multiple frequencies interact.

🔽 Upper cutoff frequency f<sub>2</sub>

  • Problems 17, 19, 21, 23 give f<sub>in</sub> and f<sub>out</sub> in MHz, plus an estimate of f<sub>2</sub>.
  • Example (Problem 17): f<sub>in</sub> = 4.78 MHz, f<sub>out</sub> = 13.6 MHz; f<sub>2</sub> < 4.78 MHz due to proximity.
  • Problem 19: f<sub>in</sub> = 40.8 MHz, f<sub>out</sub> = 3.32 MHz; f<sub>2</sub> ≈ 3.32 MHz.
  • Problem 21: f<sub>in</sub> = 84.1 MHz, f<sub>out</sub> = 33.9 Hz (likely a typo; should be MHz); f<sub>2</sub> < 33.9 MHz due to proximity.
  • Problem 23: f<sub>in</sub> = 5.63 MHz, f<sub>out</sub> = 79.6 Hz (likely a typo); f<sub>2</sub> ≈ 5.63 MHz.

📉 Roll-off and design notes

  • Problem 41: each lag network rolls off at 20 dB/decade; three networks give a total of 60 dB/decade above 1.2 MHz.
  • Don't confuse: a single RC network rolls off at 20 dB/decade; cascading n networks multiplies the slope to n × 20 dB/decade.

🔚 Incomplete answer

  • Problem 25 (Chapter 17) begins "f<sub>in</sub>" but the text cuts off; the excerpt is incomplete.
100

15.1 Introduction

15.1 Introduction

🧭 Overview

🧠 One-sentence thesis

This excerpt contains only numerical answers to end-of-chapter problems from Chapters 12–17 of an electronics textbook and does not present any substantive conceptual content, theory, or explanatory material.

📌 Key points (3–5)

  • The excerpt consists exclusively of problem solutions (numerical values with units).
  • Chapters covered include transistor circuits (Chapters 12–13), operational amplifiers (Chapter 14), digital timing (Chapter 15), decibel calculations and frequency response (Chapters 16–17).
  • No definitions, explanations, or theoretical frameworks are provided.
  • The excerpt is an answer key, not instructional content suitable for learning new concepts.

📋 Content summary

📋 What the excerpt contains

  • Format: numbered problem answers organized by chapter (12 through 17).
  • Typical entries: electrical quantities such as current (I_D), voltage (V_G, V_D, V_out), impedance (Z_in, Z_out), gain (A_v), resistance (R_D, R_G, R_S), frequency (f), phase angles (θ), and decibel values (dB, dBV, dBW, dBf).
  • Chapter 12–13: transistor biasing and amplifier parameters (drain current, gate/drain voltages, input impedance, voltage gain).
  • Chapter 14: high-frequency parameters (kHz, slew rate in V/μs, current and voltage values).
  • Chapter 15: timing values (voltages in V, time intervals in ns).
  • Chapter 16: decibel conversions, power and voltage gains, frequency response with amplitude and phase data.
  • Chapter 17: frequency breakpoints (f_in, f_out, f_bypass, f_1, f_2) for amplifier circuits, with notes on proximity effects.

⚠️ What is missing

  • No conceptual explanations, derivations, or step-by-step solutions.
  • No definitions of terms (e.g., what Z_in, A_v, or f_1 represent).
  • No diagrams, circuit schematics, or worked examples.
  • No instructional narrative or learning guidance.

🔍 Observations on problem types

🔍 Transistor circuits (Chapters 12–13)

  • Problems involve calculating DC operating points (I_D, V_G, V_D) and AC small-signal parameters (Z_in, A_v, V_out).
  • One problem (Chapter 13, problem 13) includes a design task with multiple acceptable solutions and reasoning about component choices (e.g., selecting R_G ≥ 510 kΩ, splitting R_S into R_S and R_SW to achieve a target gain of 5).

🔍 Frequency and decibel analysis (Chapters 16–17)

  • Chapter 16: conversions between linear and logarithmic units (dB, dBV, dBW, dBf), multi-stage gain calculations, and frequency/phase response at specific frequencies.
  • Chapter 17: determination of critical frequencies (corner frequencies for input, output, and bypass networks) and identification of the dominant pole (f_1 or f_2) based on proximity.
  • Example: "f_1 > 67 Hz due to proximity" indicates that the actual lower cutoff frequency is higher than the calculated value because multiple breakpoints are close together.

🔍 Graphical data

  • Several problems (e.g., Chapter 16 problems 29, 31, 33, 37, 39) reference frequency response plots (Bode plots) showing amplitude (A'_v in dB) and phase (θ in degrees) versus frequency, but the plots themselves are not reproduced in the excerpt.

⚙️ Limitations for study purposes

⚙️ No instructional value

  • The excerpt cannot be used to learn the underlying concepts, methods, or principles.
  • It serves only as a reference for checking answers to assigned problems.
  • Students seeking to understand how to solve these problems or why certain values result must consult the main textbook chapters and worked examples.

⚙️ Context dependency

  • The meaning of each answer depends entirely on the corresponding problem statement, which is not included.
  • Example: "I_D = 3 mA, V_G = 0 V, V_D = 9.6 V" (Chapter 12, problem 1) is meaningful only if the circuit configuration, component values, and transistor parameters are known.
101

15.2 IGBT Internals

15.2 IGBT Internals

🧭 Overview

🧠 One-sentence thesis

This excerpt contains only numerical answers to end-of-chapter problems from Chapters 12–17 and does not present substantive technical content about IGBT internals.

📌 Key points (3–5)

  • The excerpt consists entirely of problem solutions with numerical values and units.
  • No explanatory text, definitions, or conceptual material about IGBTs is present.
  • The answers cover topics from multiple chapters including FETs (Chapter 12–13), amplifiers (Chapter 14–16), and frequency response (Chapter 17).
  • The title "15.2 IGBT Internals" does not match the content shown, which appears to be an answer key appendix.

📋 Content description

📋 What the excerpt contains

The provided text is an answer key showing:

  • Numerical solutions to problems from Chapters 12 through 17
  • Values for electrical parameters such as current (I_D), voltage (V_G, V_D, V_out), impedance (Z_in, Z_out), gain (A_v), resistance (R_D, R_G, R_S), and frequency (f)
  • Some brief design explanations (e.g., problem 13 in Chapter 13 discusses FET amplifier design choices)
  • Frequency response data and phase angles for various problems
  • Decibel calculations and power measurements

⚠️ Missing substantive content

  • No conceptual explanations of IGBT structure or operation
  • No definitions of IGBT-specific terms or mechanisms
  • No diagrams, comparisons, or theoretical discussion
  • The excerpt does not address the stated title topic

🔍 Note on applicability

🔍 Limitation for review purposes

This excerpt cannot support meaningful review notes about "IGBT Internals" because it contains only numerical answers to unrelated practice problems. To create proper study notes on IGBT internals, source material with explanatory text, device physics, structure diagrams, and operational principles would be required.

102

15.3 IGBT Data Sheet Interpretation

15.3 IGBT Data Sheet Interpretation

🧭 Overview

🧠 One-sentence thesis

This excerpt contains only numerical answers to end-of-chapter problems across multiple chapters (12–17) and does not present substantive content on IGBT data sheet interpretation.

📌 Key points (3–5)

  • The excerpt consists entirely of problem solutions with numerical values and units.
  • No explanatory text, definitions, or conceptual material about IGBTs or data sheet interpretation is present.
  • The answers span topics including FET circuits (Chapter 12–13), frequency response (Chapter 16–17), and amplifier design.
  • The title "15.3 IGBT Data Sheet Interpretation" does not match the content shown, which appears to be from a solutions manual or answer key.

📋 Content description

📋 What the excerpt contains

The provided text is a list of numerical answers organized by chapter numbers (12 through 17). Each answer includes:

  • Problem numbers (e.g., "1.", "3.", "5.")
  • Calculated values with units (voltages, currents, impedances, frequencies, gains, phases)
  • Occasional brief design notes (e.g., problem 13 in Chapter 13 includes a design explanation)

❌ What is missing

No instructional content related to the stated title is present:

  • No discussion of IGBT (Insulated Gate Bipolar Transistor) devices
  • No guidance on reading or interpreting manufacturer data sheets
  • No definitions, concepts, or explanatory material
  • No figures, tables, or data sheet examples

🔍 Note on relevance

🔍 Mismatch between title and content

The section title suggests coverage of IGBT data sheet interpretation, but the excerpt shows only problem answers from earlier chapters covering different topics (FETs, amplifiers, frequency response). This appears to be either:

  • A pagination error where the wrong page was included
  • An answer key section that was mistakenly labeled
  • Content from a different part of the textbook

For meaningful review notes on IGBT data sheet interpretation, the actual instructional content from section 15.3 would be needed.

103

15.4 IGBT Applications

15.4 IGBT Applications

🧭 Overview

🧠 One-sentence thesis

This excerpt contains only numerical answers to end-of-chapter problems for Chapters 12–17, with no substantive content about IGBT applications or any conceptual explanations.

📌 Key points (3–5)

  • The excerpt consists entirely of problem answers (numerical values, circuit parameters, and frequency calculations).
  • No definitions, explanations, or application descriptions are present.
  • The answers span topics including FET circuits (Chapter 12–13), frequency response (Chapters 16–17), and other electronics topics.
  • The title "15.4 IGBT Applications" does not match the content, which appears to be an answer key appendix.
  • No information about IGBTs (Insulated Gate Bipolar Transistors) or their applications is provided.

📋 Content summary

📋 What the excerpt contains

The excerpt is an answer key providing numerical solutions to textbook problems across multiple chapters:

  • Chapter 12–13: FET circuit parameters (drain current, gate voltage, drain voltage, input impedance, voltage gain)
  • Chapter 14–15: Frequency and timing values (kilohertz, nanoseconds, slew rates)
  • Chapter 16: Decibel calculations and power/voltage conversions
  • Chapter 17: Frequency response calculations (input/output frequencies, bypass frequencies, cutoff frequencies)

⚠️ Missing substantive content

The excerpt does not contain:

  • Any conceptual explanations or theory
  • Definitions of terms or components
  • Application descriptions or use cases
  • Information about IGBTs specifically
  • Diagrams, derivations, or worked examples with explanations

Note: To create meaningful review notes about IGBT applications, the actual textbook section (not the answer key) would be needed.

104

Chapter 16: Decibels and Bode Plots

Chapter 16: Decibels and Bode Plots

🧭 Overview

🧠 One-sentence thesis

This excerpt provides numerical answers to Chapter 16 exercises on decibels and Bode plots, demonstrating conversions between linear and decibel scales, gain calculations, phase angles, and frequency response analysis.

📌 Key points (3–5)

  • Decibel conversions: problems cover converting between linear gain/power values and decibel representations (dB, dBW, dBV, dBm, dBf).
  • Gain and power calculations: exercises involve calculating total gain in dB, converting to linear gain, and finding output power or voltage.
  • Frequency response: problems include amplitude and phase at specific frequencies, roll-off rates (dB/decade), and corner frequencies.
  • Common confusion: decibel scales use different reference points (dBW vs. dBm vs. dBV vs. dBf)—each has a distinct zero reference.
  • Bode plot construction: exercises require plotting amplitude and phase versus frequency, identifying break points and slopes.

📐 Decibel scale conversions

📐 Linear to decibel conversions

  • Problem 1 converts six different linear values to decibels, yielding results from 10 dB down to −15.23 dB.
  • Problem 7 reverses the process: converting decibel values back to linear ratios (e.g., 1.06, 199.5, 0.398).
  • Example: a gain of 10 corresponds to 10 dB; a gain less than 1 (attenuation) yields negative dB values.

📐 Reference-based decibel scales

The excerpt uses multiple decibel scales with different reference points:

ScaleReferenceExample problems
dBW1 wattProblem 11: 0 dBW, 13.6 dBW, −172.5 dBW
dBf1 femtowattProblem 13: 150 dBf, 163.6 dBf, −22.5 dBf
dBV1 voltProblem 21: 21 dBV = 11.2 V
dBm1 milliwattProblem 17: 4 dBm, −34 dBm
  • Don't confuse: the same power level has different numerical values in different scales (e.g., 0 dBW = 30 dBm).
  • Problem 43 mentions 0.775 V as a reference (likely for dBu, though not explicitly stated).

🔗 Multi-stage gain calculations

🔗 Cascaded stages in decibels

  • Problem 15: total gain of three stages is 28 dB, which converts to a linear gain of 631.
  • Problem 17: traces signal levels through three stages for two different input levels (4 dBm and −34 dBm).
    • Example: with 4 dBm input, stage 1 outputs 6 dBm, stage 2 outputs 0 dBm, stage 3 outputs 15 dBm.
  • Problem 21: calculates intermediate voltages at each stage output (1.58 V, 2.82 V) leading to final 11.2 V output.

🔗 Power and voltage relationships

  • Problem 5: given gain G = 501, output power is 12.53 W.
  • Problem 19: converts −34 dBm input through stages, with outputs expressed in dBW (note the scale change).
  • Problem 23: finds 15 V output from a multi-stage chain.
  • Problem 45: calculates 360 W output power.

📊 Frequency response and Bode plots

📊 Amplitude and phase at specific frequencies

  • Problem 25: at 50 kHz, amplitude is −0.022 dB and phase is −4.09 degrees; at 10 MHz, amplitude is −23.1 dB and phase is −86 degrees.
  • Problem 27: amplitude portion does not change, but phase shifts are given at 30 kHz (−188.5°), 200 kHz (−225°), and 1 MHz (−258.7°).
  • Problem 31: phase angles at 4 kHz (78.7°), 20 Hz (45°), and 100 Hz (11.3°).
  • Problem 35: net gain at 20 kHz is 35.87 dB with phase 51.7°; at 800 kHz, gain is 30.5 dB with phase −70.5°.

📊 Roll-off rates and corner frequencies

  • Problem 41: each lag network rolls off at 20 dB/decade; three networks together yield 60 dB/decade total roll-off above 1.2 MHz.
  • The excerpt includes tables and plots (problems 29, 33, 37, 39) showing amplitude (A'v in dB) and phase (θ) versus frequency.
    • Example: problem 29 shows a corner at 20 Hz with 45° phase shift and amplitude changing from 32 dB at 20 Hz.
  • Problem 49: the lower frequency limit is greater than 30 Hz.

📊 Interpreting Bode plot features

  • Bode plots display two curves: amplitude (gain in dB) and phase (in degrees) versus frequency (logarithmic scale).
  • Break points (corner frequencies) mark where the slope changes (e.g., from flat to −20 dB/decade).
  • Phase typically shifts by 90° per pole or zero; multiple poles add their phase contributions.
  • Don't confuse: amplitude roll-off (dB/decade) with phase shift rate—they are related but distinct characteristics.

🧮 Additional calculations

🧮 Time-domain parameters

  • Problem 25 also gives a rise time Tr = 500 μs (microseconds).
  • This relates frequency response to time-domain behavior, though the excerpt does not elaborate on the connection.

🧮 Miscellaneous conversions

  • Problem 47: 71.5 dBV represents a voltage level.
  • Problem 3: 33 dB gain.
  • Problem 9: linear gain A = 8.57 converts to A' = 18.66 dB.
105

16.0 Chapter Objectives

16.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

This excerpt provides numerical answers to end-of-chapter problems spanning multiple chapters (12–17) in an electronics or circuit analysis textbook, covering transistor biasing, amplifier design, frequency response, and decibel calculations.

📌 Key points (3–5)

  • What the excerpt contains: numerical solutions (voltages, currents, impedances, gains, frequencies, phases) for practice problems across six chapters.
  • Chapter scope: Chapters 12–13 focus on transistor circuits (FET biasing and amplifier parameters); Chapters 14–15 cover timing and voltage; Chapter 16 addresses decibel conversions and multi-stage amplifier analysis; Chapter 17 deals with frequency response and cutoff frequencies.
  • Common confusion: the excerpt is an answer key, not instructional content—it lists results without derivations or explanations of underlying principles.
  • Design example present: Problem 13.13 includes a detailed design walkthrough for an FET amplifier achieving a specified gain using component selection and source degeneration.

🔌 Chapters 12–13: Transistor circuits

🔌 FET biasing (Chapter 12)

  • Problems 1–15 give DC operating points for field-effect transistor circuits:
    • Drain current (I_D): typically in the milliampere range (1.56 mA to 12 mA).
    • Gate voltage (V_G): often 0 V (common in JFET self-bias or depletion-mode circuits), but can be positive (e.g., 4.25 V in problem 9) or negative.
    • Drain voltage (V_D): ranges from negative (−17.2 V in problem 11) to positive (26.25 V in problem 9).
  • Problems 17 and 19 specify resistor values (e.g., 180 Ω, R_D = 1.29 kΩ; R_G typically > 1 MΩ to avoid loading the gate).

🔊 FET amplifier parameters (Chapter 13)

  • Problems 1–11 report input impedance (Z_in) and voltage gain (A_v):
    • Z_in: high values (90.9 kΩ to 910 kΩ), characteristic of FET input stages.
    • A_v: negative values indicate inversion (e.g., −2.58, −8.86); positive values (0.91) indicate non-inverting configurations (e.g., source follower).
    • Output voltage (V_out): given in millivolts, with "(inverted)" noted where applicable.

🛠️ Amplifier design example (Problem 13.13)

Design goal: achieve a specified voltage gain using a 15 VDC supply, with constraints on transconductance (g_m0 = 25 mS) and load resistance (R_L = 10 kΩ).

Design steps outlined:

  1. Check feasibility: unswamped ceiling gain = g_m0 × R_L = 250 (achievable).
  2. Bias point selection: mid-point bias (V_GS = 0.5 × V_GS(off), I_D = 0.25 × I_DSS) yields I_D = 6.25 mA, g_m = 12.5 mS.
  3. Source resistor (R_S): R_S = 1 V / 6.25 mA = 160 Ω.
  4. Drain resistor (R_D): choosing 1 kΩ gives effective load r_L = 909 Ω and V_D = 8.75 V (reasonable).
  5. Source degeneration: split R_S into R_S and R_SW (swamping resistor) to set gain; example: R_SW = 100 Ω, R_S = 60 Ω → gain = 5.
  • Don't confuse: "unswamped gain" (maximum, no degeneration) vs. "swamped gain" (reduced by adding R_SW to increase linearity and control gain).

📐 Chapters 14–15: Timing and voltage

⏱️ Chapter 14 answers

  • Problem 1: 35 kHz (likely a frequency calculation).
  • Problem 3: 33.3 V/μs (slew rate, the maximum rate of voltage change an amplifier can produce).
  • Problem 5: 12.48 A, 0.15 V (current and voltage, possibly related to power supply or output stage).

⚡ Chapter 15 answers

  • Problem 1: 4 V (voltage level).
  • Problem 3: 30 ns (nanoseconds, a timing parameter such as propagation delay).
  • Problem 5: 610 ns (another timing parameter).

📊 Chapter 16: Decibels and multi-stage amplifiers

📊 Decibel conversions (Problems 1, 7, 11, 13)

  • Problem 1: converts power or voltage ratios to decibels (dB), ranging from −15.23 dB to 26.99 dB.
  • Problem 7: converts numeric gains to ratios (e.g., 1.06, 199.5, 0.398).
  • Problem 11: power in dBW (decibels relative to 1 watt), from −172.5 dBW to 30.8 dBW.
  • Problem 13: power in dBf (decibels relative to 1 femtowatt), from −22.5 dBf to 180.8 dBf.
  • Example: 0 dB corresponds to a ratio of 1 (no gain or loss); positive dB means gain, negative dB means attenuation.

🔗 Multi-stage amplifier analysis (Problems 3, 5, 15, 17, 21)

ProblemKey resultInterpretation
333 dBTotal gain in decibels
5G = 501, P_out = 12.53 WNumeric gain and output power
15G'_total = 28 dB, G = 631Total gain in dB and numeric form
17Stage outputs for two input levelsCascaded stage outputs in dBm and dBW
21V'_out = 21 dBV, final 11.2 VVoltage gain through stages, with intermediate outputs
  • Cascaded stages: each stage's output becomes the next stage's input; total gain in dB is the sum of individual stage gains.
  • Example (Problem 17): for input P'_in = 4 dBm, stage 1 outputs 6 dBm, stage 2 outputs 0 dBm, stage 3 outputs 15 dBm.

📈 Frequency and phase response (Problems 25, 27, 29, 31, 33, 35, 37, 39)

  • Amplitude and phase at specific frequencies:
    • Problem 25: at 50 kHz (−0.022 dB, −4.09°), 700 kHz (−3 dB, −45°), 10 MHz (−23.1 dB, −86°); rise time T_r = 500 μs.
    • Problem 27: amplitude unchanged; phases at 30 kHz (−188.5°), 200 kHz (−225°), 1 MHz (−258.7°).
    • Problem 31: phases at 4 kHz (78.7°), 20 Hz (45°), 100 Hz (11.3°).
    • Problem 35: net gain and phase at multiple frequencies (e.g., 20 kHz: 35.87 dB, 51.7°; 800 kHz: 30.5 dB, −70.5°).
  • Graphical data (Problems 29, 33, 37, 39): frequency vs. gain (A'_v in dB) and frequency vs. phase (θ in degrees); specific points listed (e.g., 200 kHz: 18 dB; 2 MHz roll-off).
  • Don't confuse: −3 dB point (half-power frequency, often the cutoff) vs. other attenuation levels; phase shift accumulates through stages and frequency-dependent networks.

🔊 Power and voltage reference levels (Problems 19, 23, 43, 45, 47, 49)

  • Problem 19: 200 mW (power level).
  • Problem 23: 15 V (voltage level).
  • Problem 43: 0.775 V (standard reference voltage for 0 dBu in audio).
  • Problem 45: 360 W (power).
  • Problem 47: 71.5 dBV (voltage in decibels relative to 1 V).
  • Problem 49: Greater than 30 Hz (frequency constraint, likely a lower cutoff).

📉 Roll-off and network behavior (Problem 41)

  • Three lag networks: each contributes 20 dB/decade roll-off; combined roll-off is 60 dB/decade above 1.2 MHz.
  • Example: a single-pole low-pass filter rolls off at 20 dB/decade; cascading three such filters triples the slope.

🎛️ Chapter 17: Frequency response analysis

🎛️ Cutoff frequencies (Problems 1–15)

Each problem lists input cutoff (f_in), output cutoff (f_out), bypass cutoff (f_bypass), and the dominant lower cutoff (f_1):

  • f_in: determined by input coupling capacitor and input impedance.
  • f_out: determined by output coupling capacitor and load impedance.
  • f_bypass: determined by emitter/source bypass capacitor and associated resistances.
  • f_1 (dominant lower cutoff): the highest of the three cutoff frequencies, or affected by proximity (when two cutoffs are close, the actual f_1 is slightly higher).
Problemf_inf_outf_bypassf_1 (dominant)
167 Hz2.65 Hz61.2 Hz> 67 Hz (proximity)
39.76 Hz0.796 Hz2.57 Hz> 9.76 Hz (proximity)
5477 Hz8.8 Hz7.42 Hz≈ 477 Hz
74.5 Hz2.6 Hz69.4 Hz≈ 69.4 Hz
94.5 Hz67.9 Hz(not listed)≈ 67.9 Hz
110.34 Hz0.723 Hz16.3 Hz≈ 16.3 Hz
131.59 Hz3.22 Hz0.781 Hz> 3.22 Hz (proximity)
157.23 Hz0.672 Hz(not listed)≈ 7.23 Hz
  • Proximity effect: when two cutoff frequencies are within an octave, the actual lower cutoff is higher than the highest individual cutoff.
  • Example (Problem 1): f_in = 67 Hz and f_bypass = 61.2 Hz are close, so f_1 > 67 Hz.

🎛️ Upper cutoff frequencies (Problems 17–25)

Each problem lists input and output high-frequency cutoffs (f_in, f_out) and the dominant upper cutoff (f_2):

  • f_2 (dominant upper cutoff): the lowest of the two cutoff frequencies, or affected by proximity.
Problemf_inf_outf_2 (dominant)
174.78 MHz13.6 MHz< 4.78 MHz (proximity)
1940.8 MHz3.32 MHz≈ 3.32 MHz
2184.1 MHz33.9 Hz< 33.9 MHz (proximity)
235.63 MHz79.6 Hz≈ 5.63 MHz
  • Don't confuse: lower cutoff (f_1) is where low-frequency response rolls off (due to coupling/bypass capacitors); upper cutoff (f_2) is where high-frequency response rolls off (due to transistor capacitances and stray capacitance).
  • Example (Problem 17): f_in = 4.78 MHz is lower than f_out = 13.6 MHz, but proximity shifts f_2 below 4.78 MHz.

🔍 Incomplete data (Problem 25)

  • Problem 25 ends with "f_in" but no value is provided in the excerpt.
106

Answer Key Excerpt (Chapter 12–17)

16.1 Introduction

🧭 Overview

🧠 One-sentence thesis

This excerpt provides numerical answers to end-of-chapter problems for Chapters 12 through 17, covering transistor circuit analysis, amplifier design, frequency response, and decibel calculations.

📌 Key points (3–5)

  • What this excerpt contains: numerical solutions (voltages, currents, impedances, gains, frequencies, phases) for practice problems across six chapters.
  • Topics covered: JFET/MOSFET biasing (Ch. 12–13), amplifier parameters (Ch. 14–16), and frequency response analysis (Ch. 17).
  • Format: each answer lists the problem number followed by calculated values with units.
  • Common confusion: this is an answer key, not instructional content—it does not explain how to arrive at these values or the underlying theory.
  • One worked example: Problem 13 in Chapter 13 outlines a design process for achieving a specific voltage gain using a JFET amplifier.

📋 Structure and scope

📋 What chapters are covered

The excerpt spans six chapters:

ChapterTopic hint from answers
12Drain current, gate voltage, drain voltage for FET circuits
13Input impedance, voltage gain, FET amplifier design
14Frequency (kHz), slew rate (V/μs), current/voltage
15Voltage (V), time (ns)
16Decibel calculations (dB, dBW, dBV, dBf), gain, power, phase
17Frequency response: input/output/bypass frequencies, corner frequencies

📋 Answer format

  • Most answers are numerical values with units: mA, V, kΩ, dB, Hz, degrees.
  • Some problems have multiple parts (A, B, C…) with separate answers.
  • Problem 13 in Chapter 13 is descriptive: it walks through a design procedure rather than giving a single number.

🔧 Chapter 13 Problem 13: a worked design example

🔧 Design goal and constraints

Problem 13 asks to design a circuit with specific performance; the answer outlines the reasoning.

  • Goal: achieve a particular voltage gain (A_v).
  • Given parameters: transconductance g_m0 = 25 mS, load resistance R_L = 10 kΩ.
  • Supply voltage: standard 15 VDC.
  • Ceiling gain: unswamped (maximum) gain = 250, calculated from g_m0 × R_L.

🔧 Design steps outlined

  1. Bias point selection: choose mid-point bias (V_GS = 0.5 × V_GS(off), I_D = 0.25 × I_DSS).
  2. Calculate source resistance: R_S = 1 V / 6.25 mA = 160 Ω.
  3. Resulting transconductance: g_m = 12.5 mS at this bias.
  4. Choose drain resistor: R_D = 1 kΩ (a common value).
  5. Effective load: r_L = 909 Ω; drain voltage V_D = 8.75 V (reasonable).
  6. Unswamped gain: over 11 (still too high for the target).
  7. Add source degeneration: split the 160 Ω into R_S and R_SW (swamping resistor).
    • Example: R_SW = 100 Ω, R_S = 60 Ω → gain = 5 (meets the target).

🔧 Key insight

  • Why split the source resistance: the unswamped gain is too high; adding R_SW (source degeneration) reduces gain to the desired level.
  • Design flexibility: "nearly infinite possibilities"—many component combinations can meet the goal.
  • Don't confuse: R_G (gate resistor) is not critical for DC bias here; it is typically chosen > 1 MΩ for high input impedance.

📐 Chapter 16: decibel and phase calculations

📐 Types of decibel units

The answers use multiple decibel scales:

UnitMeaning
dBRelative gain (dimensionless ratio)
dBWPower relative to 1 watt
dBVVoltage relative to 1 volt
dBfFemtowatt scale (power relative to 1 femtowatt)
dBmPower relative to 1 milliwatt (implied in problem 17)

📐 Gain and phase at different frequencies

  • Problem 25, 27, 29, 31, 35, 37: answers include both amplitude (dB) and phase (degrees) at specified frequencies.
  • Negative dB: indicates attenuation (gain < 1).
  • Phase shift: ranges from positive (lead) to negative (lag); e.g., −45° at a corner frequency, approaching −90° or −270° at higher frequencies.
  • Example from Problem 25: at 700 kHz, gain = −3 dB, phase = −45 degrees (typical for a single-pole rolloff at the corner frequency).

📐 Cascaded stages

  • Problem 15: total gain G'_total = 28 dB, linear gain G = 631.
  • Problem 17: tracks signal level through three stages for two different input levels (4 dBm and −34 dBm).
  • Problem 21: calculates intermediate stage outputs in dBV, then converts final output to volts (11.2 V).

🌀 Chapter 17: frequency response

🌀 Corner frequencies

Each problem lists multiple frequencies:

  • f_in: input coupling network corner frequency.
  • f_out: output coupling network corner frequency.
  • f_bypass: bypass capacitor corner frequency (for emitter/source bypass).
  • f_1 or f_2: dominant lower or upper corner frequency for the overall amplifier.

🌀 Proximity effects

Several answers note "due to proximity":

  • When multiple corner frequencies are close together, they interact.
  • The dominant corner frequency (f_1 or f_2) may be higher or lower than the individual corner frequencies.
  • Example from Problem 1: f_in = 67 Hz, f_out = 2.65 Hz, f_bypass = 61.2 Hz → f_1 > 67 Hz due to proximity (the two nearby corners at 67 Hz and 61.2 Hz push the −3 dB point higher).

🌀 High-frequency response

  • Problems 17–25: calculate f_in and f_out in MHz (high-frequency rolloff due to transistor and stray capacitances).
  • Example from Problem 17: f_in = 4.78 MHz, f_out = 13.6 MHz → f_2 < 4.78 MHz due to proximity.

⚠️ Limitations of this excerpt

⚠️ No explanatory content

  • This is a solutions manual page: it lists answers but does not teach the underlying concepts, derivations, or problem-solving methods.
  • To understand why these values are correct, you would need the corresponding textbook chapters and problem statements.

⚠️ Incomplete problem 25

  • The last line reads "25. f_in" with no value—the excerpt cuts off mid-answer.
107

16.2 The Decibel

16.2 The Decibel

🧭 Overview

🧠 One-sentence thesis

This excerpt contains only answer keys for end-of-chapter problems from Chapters 12–17 and Appendix D, with no substantive content explaining the decibel concept.

📌 Key points (3–5)

  • The excerpt is a collection of numerical answers to textbook exercises.
  • Chapter 16 answers include decibel (dB) calculations, power and voltage conversions, and gain/phase problems.
  • No definitions, explanations, or conceptual content about the decibel are present.
  • The excerpt also includes unrelated material (a recipe for "Autumn Bread" in Appendix D).

📋 What the excerpt contains

📋 Answer key structure

The excerpt is formatted as a list of problem numbers followed by numerical answers, organized by chapter:

  • Chapters 12–13: Transistor circuit parameters (drain current, gate voltage, impedance, gain).
  • Chapters 14–15: Frequency, slew rate, current, voltage, and timing answers.
  • Chapter 16: Decibel-related answers (dB, dBW, dBf, dBV, dBm), gain, power, phase, and frequency response.
  • Chapter 17: Frequency cutoffs and capacitor values.
  • Appendix D: A recipe introduction (non-technical content).

🔢 Chapter 16 answers (decibel-related)

The Chapter 16 answers include:

ProblemType of answerExample values
1Decibel values (dB)10 dB, 19 dB, −6.99 dB, −15.23 dB
5, 9Gain and powerG = 501, P_out = 12.53 W; A = 8.57, A' = 18.66 dB
11, 13Power in dBW and dBf0 dBW, 13.6 dBW, −172.5 dBW; 150 dBf, 180.8 dBf
21, 23Voltage in dBV21 dBV = 11.2 V; 15 V
25, 27, 35Frequency response (gain and phase)−0.022 dB at 50 kHz, −45° at 700 kHz; 35.87 dB at 20 kHz
  • These are final answers only; no working, formulas, or explanations are provided.

⚠️ Limitations of this excerpt

⚠️ No conceptual content

  • The excerpt does not define the decibel, explain logarithmic scales, or describe how to convert between linear and decibel units.
  • It does not explain why decibels are used, what dBW/dBV/dBf/dBm mean, or how to interpret negative dB values.
  • There is no discussion of gain, attenuation, frequency response, or phase shift concepts.

⚠️ Context missing

  • The problems themselves are not included, so the meaning of each answer is unclear.
  • Example: "Problem 1: A) 10 dB B) 19 dB..." lists answers but does not state what quantity is being calculated or what the input values were.

⚠️ Unrelated material

  • Appendix D begins with a quote and a recipe for "Autumn Bread," which is entirely unrelated to electronics or the decibel.
  • This suggests the excerpt is from the back matter of a textbook and was not curated for the topic "16.2 The Decibel."

📌 Summary

📌 What can be inferred

From the Chapter 16 answers, we can infer that the section likely covers:

  • Decibel conversions: between linear gain/power and dB units.
  • Reference units: dBW (decibels relative to 1 watt), dBV (relative to 1 volt), dBf (relative to 1 femtowatt), dBm (relative to 1 milliwatt).
  • Frequency response: gain and phase shift at different frequencies, roll-off rates (e.g., "60 dB/decade").
  • Cascaded stages: total gain from multiple amplifier stages.

However, none of these concepts are explained in the excerpt—only numerical answers are provided.

📌 Recommendation

To learn about the decibel, consult the actual textbook section 16.2 or other instructional material. This excerpt is useful only for checking answers to assigned problems.

108

Bode Plots

16.3 Bode Plots

🧭 Overview

🧠 One-sentence thesis

Bode plots provide a graphical method to represent the frequency response of circuits by showing how gain (in dB) and phase shift change across different frequencies.

📌 Key points (3–5)

  • What Bode plots show: amplitude (gain in dB) and phase (in degrees) as functions of frequency.
  • Roll-off behavior: networks can roll off at rates like 20 dB/decade or 60 dB/decade (multiple lag networks combined).
  • Key frequency points: critical frequencies where gain drops by 3 dB or phase shifts by 45 degrees mark important transitions.
  • Common confusion: amplitude and phase are separate plots—amplitude changes (dB) do not always correlate directly with phase changes (degrees) at the same frequency.
  • Practical use: calculating net gain and phase at specific frequencies by combining contributions from multiple stages or networks.

📊 Gain and phase measurements

📈 Gain in decibels

  • Gain is expressed in dB (decibels), which is a logarithmic measure.
  • The excerpt shows gain values ranging from positive (e.g., 40 dB, 36 dB) to negative (e.g., −3 dB, −23.1 dB).
  • A −3 dB point is significant: it marks a frequency where the amplitude has dropped to about 70.7% of its maximum.
  • Example: At 700 kHz, gain is −3 dB with a phase of −45 degrees; at 10 MHz, gain drops further to −23.1 dB with phase near −86 degrees.

🔄 Phase shift in degrees

  • Phase is measured in degrees and can be positive or negative.
  • Positive phase (e.g., 78.7°, 45°, 11.3°) indicates a lead; negative phase (e.g., −45°, −86°, −188.5°) indicates a lag.
  • The −45 degree point often coincides with the −3 dB frequency in simple RC networks.
  • Example: At 20 Hz, phase is 45°; at 100 Hz, phase drops to 11.3°; at higher frequencies, phase can become very negative (e.g., −258.7° at 1 MHz).

📉 Roll-off rates

Roll-off rate: the rate at which gain decreases with increasing frequency, measured in dB per decade.

  • A single lag network typically rolls off at 20 dB/decade.
  • Multiple lag networks combine their effects: three lag networks yield 60 dB/decade total roll-off above a certain frequency.
  • Example: Above 1.2 MHz, each of three lag networks contributes 20 dB/decade, resulting in a combined 60 dB/decade slope.

🔗 Multi-stage frequency response

🔗 Net gain and phase

  • When multiple stages or networks are cascaded, their individual gains (in dB) add, and their phase shifts add.
  • The excerpt provides examples of calculating net gain and net phase at specific frequencies.
  • Example: At 20 kHz, net gain is 35.87 dB and net phase is 51.7°; at 100 kHz, phase becomes −5.1° and gain is 40 dB; at 800 kHz, phase is −70.5° and gain drops to 30.5 dB.

🧮 Stage-by-stage calculation

  • For a given input power or voltage (in dBm, dBW, or dBV), each stage adds its own gain.
  • The output of one stage becomes the input to the next.
  • Example: For an input of 4 dBm, stage 1 outputs 6 dBm, stage 2 outputs 0 dBm, and stage 3 outputs 15 dBm. For an input of −34 dBm, outputs are −24 dBW, −30 dBW, and −15 dBW respectively.

📐 Voltage and power conversions

  • Gain in dB can be converted back to linear voltage or power ratios.
  • Example: 21 dBV corresponds to 11.2 V final output; 4 dBV equals 1.58 V for stage 1; 9 dBV equals 2.82 V for stage 2.
  • Don't confuse: dBV, dBm, dBW, and dBf are different reference scales; the excerpt uses multiple units depending on context.

🎛️ Frequency-dependent behavior

🎛️ Amplitude vs frequency

  • Amplitude (gain) typically decreases as frequency increases beyond a certain cutoff.
  • The excerpt shows amplitude plots with frequencies on the horizontal axis and gain (A'v) in dB on the vertical axis.
  • Example: At 200 kHz, gain is 18 dB; at 20 Hz, gain is 32 dB; at 250 kHz and 300 kHz, gain values are plotted but specific numbers are not always given.

🌀 Phase vs frequency

  • Phase shift varies with frequency, often starting near 0° at low frequencies and becoming more negative (or positive, depending on the network type) at higher frequencies.
  • The excerpt shows phase plots ranging from 0° to −270°.
  • Example: Phase is 0° at very low frequencies, −180° at some mid-range frequency, and −270° at higher frequencies for a particular network.

⏱️ Critical frequencies

  • f₁ and f₂ denote lower and upper critical frequencies (cutoff points).
  • The dominant frequency (the one that determines the overall response) is often the highest f₁ or the lowest f₂, but proximity effects can shift the actual cutoff.
  • Example: f_in = 67 Hz, f_out = 2.65 Hz, f_bypass = 61.2 Hz; the actual f₁ is greater than 67 Hz due to proximity. In another case, f_in = 4.78 MHz, f_out = 13.6 MHz; f₂ is less than 4.78 MHz due to proximity.

🧪 Practical design considerations

🧪 Component selection

  • Capacitor values (C_in, C_out, C_bypass) are chosen to set desired cutoff frequencies.
  • The excerpt provides calculated capacitor values for different target frequencies.
  • Example: For 50 Hz, C_in = 902 nF and C_out = 29.9 μF; the designer picks one value and adjusts the other. For 1 kHz, C_in = 79.4 pF, C_out = 7.23 nF, C_bypass = 531 nF.

🧪 Resistor degeneration

  • Resistors R_S and R_SW are used to set gain by adding degeneration.
  • Example: R_SW = 100 Ω with R_S = 60 Ω yields a gain of 5.

🧪 Time constants

  • The excerpt mentions T_r = 500 μsec, which is a time constant related to the circuit's transient response.
  • Time constants influence how quickly the circuit responds to changes and are linked to the cutoff frequencies.
109

16.4 Combining the Elements - Multi-Stage Effects

16.4 Combining the Elements - Multi-Stage Effects

🧭 Overview

🧠 One-sentence thesis

Multi-stage amplifier systems combine individual stage gains and phase shifts to produce a net frequency-dependent response that can be analyzed by adding decibel gains and summing phase angles at each frequency.

📌 Key points (3–5)

  • Multi-stage gain calculation: total gain in dB is the sum of individual stage gains in dB; in linear terms, total gain is the product of individual stage gains.
  • Phase accumulation: net phase shift is the algebraic sum of phase shifts from each stage at a given frequency.
  • Frequency-dependent behavior: amplitude and phase responses vary with frequency; each stage contributes roll-off and phase shift that combine to shape the overall system response.
  • Common confusion: don't confuse dB addition (logarithmic) with linear multiplication—dB values add, but linear gain values multiply.
  • Practical analysis: working through each stage's output at a given input level (in dBm, dBV, or dBW) reveals the signal level progression through the cascade.

🔗 Multi-stage gain and signal flow

🔗 Total gain calculation

Total gain in a multi-stage system: the sum of individual stage gains when expressed in dB, or the product of individual stage gains when expressed as linear ratios.

  • In dB: G'_total = G'_1 + G'_2 + G'_3 + ...
  • In linear terms: G_total = G_1 × G_2 × G_3 × ...
  • Example (from problem 15): G'_total = 28 dB corresponds to a linear gain G = 631.
  • Example (from problem 17): For an input of 4 dBm, stage 1 outputs 6 dBm, stage 2 outputs 0 dBm, and stage 3 outputs 15 dBm; each stage adds its gain to the previous stage's output.

📶 Signal level progression

  • The output of one stage becomes the input to the next.
  • Tracking signal levels stage-by-stage (in dBm, dBV, or dBW) shows how the signal grows or attenuates through the cascade.
  • Example (from problem 21): an input of V'_out = 21 dBV results in a final output of 11.2 V; intermediate stages show 4 dBV = 1.58 V (stage 1) and 9 dBV = 2.82 V (stage 2).
  • Don't confuse: dB values are added/subtracted; voltage or power values in linear units are multiplied/divided.

📐 Frequency response and phase behavior

📐 Amplitude response at different frequencies

  • Each stage contributes frequency-dependent gain (roll-off or boost).
  • Net gain at a specific frequency is the sum of all stage gains at that frequency (in dB).
  • Example (from problem 35): at 20 kHz, net gain = 35.87 dB; at 100 kHz, A'_v = 40 dB; at 800 kHz, A'_v = 30.5 dB.
  • Example (from problem 25): at 50 kHz, amplitude is −0.022 dB; at 700 kHz, −3 dB; at 10 MHz, −23.1 dB.

🔄 Phase shift accumulation

  • Each stage introduces a phase shift that depends on frequency.
  • Net phase at a given frequency is the algebraic sum of individual stage phase shifts.
  • Example (from problem 25): at 50 kHz, phase = −4.09°; at 700 kHz, phase = −45°; at 10 MHz, phase = −86°.
  • Example (from problem 27): amplitude does not change, but phases are −188.5° at 30 kHz, −225° at 200 kHz, and −258.7° at 1 MHz.
  • Example (from problem 35): at 20 kHz, net phase = 51.7°; at 100 kHz, phase = −5.1°; at 800 kHz, phase = −70.5°.

🌀 Roll-off and lag networks

  • Multiple lag networks (low-pass filters) in cascade produce steeper roll-off.
  • Example (from problem 41): each lag network rolls off at 20 dB/decade; three in series yield 60 dB/decade total roll-off above 1.2 MHz.
  • Don't confuse: a single stage's roll-off rate (e.g., 20 dB/decade) with the cumulative rate when stages are cascaded.

🧮 Practical calculations and conversions

🧮 dB and linear conversions

  • The excerpt provides numerous conversions between dB and linear units (voltage, power).
  • Example (from problem 1): gains range from 10 dB to −15.23 dB; problem 7 shows A = 8.57 (linear) corresponds to A' = 18.66 dB.
  • Example (from problem 11): power levels in dBW range from 0 dBW (1 W) to −172.5 dBW (very small power).
  • Example (from problem 43): 0.775 V is a reference level (0 dBV in some contexts).

🔌 Power and voltage outputs

  • Final output levels can be expressed in watts or volts.
  • Example (from problem 5): G = 501, P_out = 12.53 W.
  • Example (from problem 19): output power = 200 mW.
  • Example (from problem 23): final output = 15 V.
  • Example (from problem 45): output power = 360 W.

🛠️ Design and component selection

🛠️ Gain adjustment with resistors

  • Gain can be set by splitting a feedback resistor into two parts (R_S and R_SW), with R_SW adding degeneration.
  • Example: R_SW = 100 Ω with R_S = 60 Ω yields a gain of 5.
  • This allows fine-tuning of the desired gain level.

🎛️ Frequency limits and proximity effects

  • The dominant pole (f_1 or f_2) is determined by the highest low-frequency cutoff or lowest high-frequency cutoff among input, output, and bypass networks.
  • Example (from problem 1): f_in = 67 Hz, f_out = 2.65 Hz, f_bypass = 61.2 Hz; f_1 > 67 Hz due to proximity (the highest of the three dominates).
  • Example (from problem 17): f_in = 4.78 MHz, f_out = 13.6 MHz; f_2 < 4.78 MHz due to proximity (the lowest of the two dominates).
  • Don't confuse: the individual cutoff frequencies with the system's overall bandwidth—proximity effects shift the dominant pole.

🧰 Capacitor selection

  • Capacitor values (C_in, C_out, C_bypass) are chosen to set desired cutoff frequencies.
  • Example (from problem 31): C_in = 239 nF, C_out = 440 nF, C_bypass = 371 μF.
  • Example (from problem 33): for 50 Hz, C_in = 902 nF, C_out = 29.9 μF; the designer picks one value and increases the other to ensure the desired cutoff.
  • Example (from problem 37): for 1 kHz, C_in = 79.4 pF, C_out = 7.23 nF, C_bypass = 531 nF; pick one and increase the others to maintain the target frequency response.
110

Chapter 17: Frequency Limits

Chapter 17: Frequency Limits

🧭 Overview

🧠 One-sentence thesis

The lower and upper frequency limits of an amplifier are determined by the cutoff frequencies of input, output, and bypass capacitor networks, and the dominant (highest or lowest) frequency among them sets the overall bandwidth boundary.

📌 Key points (3–5)

  • What the chapter covers: calculating the individual cutoff frequencies (f_in, f_out, f_bypass) that define the low-frequency limit (f₁) and the high-frequency limit (f₂) of amplifier circuits.
  • How to find the overall limit: the lower frequency limit f₁ is determined by the highest of the low-frequency cutoffs; the upper frequency limit f₂ is determined by the lowest of the high-frequency cutoffs.
  • Proximity effect: when two or more cutoff frequencies are close together, the overall limit shifts beyond the single dominant frequency due to their combined roll-off.
  • Common confusion: don't assume f₁ equals the highest individual f_in/f_out/f_bypass—if frequencies are close ("due to proximity"), f₁ will be greater than the highest individual cutoff.
  • Design task: selecting capacitor values (C_in, C_out, C_bypass) to achieve a desired lower frequency limit, often requiring one capacitor to be chosen and others increased accordingly.

🔽 Low-frequency cutoff determination

🔽 Three low-frequency contributors

Each amplifier stage has three capacitor-coupled networks that create low-frequency roll-offs:

  • f_in: cutoff frequency of the input coupling capacitor network.
  • f_out: cutoff frequency of the output coupling capacitor network.
  • f_bypass: cutoff frequency of the emitter/source bypass capacitor network.

All three act as high-pass filters, blocking low frequencies and passing higher frequencies.

🎯 Finding the overall lower limit f₁

The lower frequency limit f₁ is the frequency below which the amplifier gain begins to roll off significantly.

  • Simple case: if one cutoff is much higher than the others, f₁ approximately equals that highest cutoff.
    • Example: f_in = 67 Hz, f_out = 2.65 Hz, f_bypass = 61.2 Hz → f₁ > 67 Hz due to proximity (problem 1).
  • Proximity case: when two or more cutoffs are close in value, their combined effect pushes f₁ higher than the highest individual cutoff.
    • Example: f_in = 1.59 Hz, f_out = 3.22 Hz, f_bypass = 0.781 Hz → f₁ > 3.22 Hz due to proximity (problem 13).
  • Don't confuse: f₁ is not simply the arithmetic average or the highest cutoff—it is determined by the combined roll-off behavior of all three networks.

📐 Typical problem patterns

Problemf_inf_outf_bypassResult
167 Hz2.65 Hz61.2 Hzf₁ > 67 Hz (proximity)
5477 Hz8.8 Hz7.42 Hzf₁ ≈ 477 Hz (dominant)
74.5 Hz2.6 Hz69.4 Hzf₁ ≈ 69.4 Hz (dominant)
110.34 Hz0.723 Hz16.3 Hzf₁ ≈ 16.3 Hz (dominant)
  • When one frequency is much higher, it dominates and f₁ approximately equals it.
  • When frequencies are close, proximity effects require f₁ to be greater than the highest individual value.

🔼 High-frequency cutoff determination

🔼 Two high-frequency contributors

At high frequencies, parasitic capacitances and device limitations create low-pass behavior:

  • f_in: high-frequency cutoff of the input network (often due to input capacitance and source resistance).
  • f_out: high-frequency cutoff of the output network (often due to load capacitance and output resistance).

Both act as low-pass filters, passing lower frequencies and attenuating higher frequencies.

🎯 Finding the overall upper limit f₂

The upper frequency limit f₂ is the frequency above which the amplifier gain begins to roll off significantly.

  • Simple case: if one cutoff is much lower than the others, f₂ approximately equals that lowest cutoff.
    • Example: f_in = 40.8 MHz, f_out = 3.32 MHz → f₂ ≈ 3.32 MHz (problem 19).
  • Proximity case: when two cutoffs are close in value, their combined effect pushes f₂ lower than the lowest individual cutoff.
    • Example: f_in = 4.78 MHz, f_out = 13.6 MHz → f₂ < 4.78 MHz due to proximity (problem 17).
  • Don't confuse: f₂ is not simply the lowest cutoff—it is determined by the combined roll-off behavior of both networks.

📐 Typical problem patterns

Problemf_inf_outResult
174.78 MHz13.6 MHzf₂ < 4.78 MHz (proximity)
1940.8 MHz3.32 MHzf₂ ≈ 3.32 MHz (dominant)
2184.1 MHz33.9 Hzf₂ < 33.9 MHz (proximity)
235.63 MHz79.6 Hzf₂ ≈ 5.63 MHz (dominant)
25128 MHz13.3 MHzf₂ ≈ 13.3 MHz (dominant)
  • When one frequency is much lower, it dominates and f₂ approximately equals it.
  • When frequencies are close, proximity effects require f₂ to be less than the lowest individual value.

🧮 Capacitor design for desired frequency limits

🧮 Calculating required capacitor values

Given a desired lower frequency limit (e.g., 50 Hz or 1 kHz), the designer must choose capacitor values for C_in, C_out, and C_bypass.

  • Problem 31: for an unspecified lower limit, C_in = 239 nF, C_out = 440 nF, C_bypass = 371 μF.
  • Problem 33: for 50 Hz lower limit, C_in = 902 nF, C_out = 29.9 μF. The instruction is "Pick one and increase the other."
  • Problem 35: for 50 Hz lower limit, C_in = 1.45 nF, C_out = 4.43 μF. Again, "Pick one and increase the other."
  • Problem 37: for 1 kHz lower limit, C_in = 79.4 pF, C_out = 7.23 nF, C_bypass = 531 nF. "Pick one and increase the others."

🔧 Design strategy

  • Why "pick one and increase the other(s)": to ensure the overall f₁ meets the target, one capacitor is chosen at the calculated value, and the others are increased (lowering their cutoff frequencies) so they do not interfere.
  • Trade-offs: larger capacitors cost more and occupy more space, but they push cutoff frequencies lower, ensuring the dominant cutoff is the one you designed for.
  • Example: if you want f₁ = 50 Hz and calculate C_in for 50 Hz, you should increase C_out and C_bypass so their cutoffs are well below 50 Hz, making f_in the dominant (highest) low-frequency cutoff.

📊 Capacitor value ranges

Target f₁Typical C_inTypical C_outTypical C_bypass
50 Hz~1 nF to ~1 μF~5 μF to ~30 μF~100 μF to ~500 μF
1 kHz~80 pF to ~100 pF~7 nF to ~10 nF~500 nF to ~1 μF
  • Bypass capacitors are typically much larger (microfarads) because they must bypass low-frequency AC signals around the emitter/source resistor.
  • Input and output coupling capacitors are smaller (nanofarads to microfarads) depending on the impedance levels in the circuit.

🔍 Proximity and combined roll-off

🔍 What "due to proximity" means

When two or more cutoff frequencies are close together (within a factor of 2–3), their roll-off slopes add, causing the overall cutoff to shift away from the individual values.

  • At low frequencies: if f_in, f_out, and f_bypass are close, the combined 20 dB/decade slopes (one from each network) create a steeper roll-off, pushing f₁ higher than the highest individual cutoff.
  • At high frequencies: if f_in and f_out are close, the combined roll-off pushes f₂ lower than the lowest individual cutoff.

🧩 How to avoid proximity issues

  • Design one dominant pole: choose capacitor values so that one cutoff is clearly dominant (at least a factor of 3–5 away from the others).
  • Example: if you want f₁ ≈ 100 Hz, design f_in = 100 Hz, f_out = 20 Hz, f_bypass = 20 Hz. Then f₁ will be approximately 100 Hz without proximity shift.
  • Don't confuse: a single dominant pole simplifies analysis and ensures predictable bandwidth; multiple close poles require more complex calculations to find the true f₁ or f₂.

📉 Roll-off rate and multiple poles

  • Each capacitor network contributes a 20 dB/decade roll-off.
  • If three low-frequency poles are close together, the combined roll-off can approach 60 dB/decade (as noted in problem 41 for lag networks).
  • This steeper roll-off means the gain drops more rapidly outside the passband, which can be desirable for filtering but complicates the determination of the exact cutoff frequency.
111

Chapter 17 Objectives

17.0 Chapter Objectives

🧭 Overview

🧠 One-sentence thesis

This excerpt lacks substantive content; it contains only numerical answers to chapter exercises and an unrelated appendix about cooking.

📌 Key points (3–5)

  • The excerpt consists entirely of numerical solutions to problems from Chapters 14–17 of what appears to be an electronics textbook.
  • Chapter 17 answers focus on frequency calculations (f_in, f_out, f_bypass, f_1, f_2) and capacitor values (C_in, C_out, C_bypass).
  • No conceptual explanations, definitions, or learning objectives are provided in the excerpt.
  • The excerpt ends with an appendix about cooking "Autumn Bread," which is unrelated to the chapter content.

📋 What the excerpt contains

📋 Answer key structure

The excerpt is formatted as a list of problem numbers followed by numerical answers or brief statements.

  • Problems are numbered (e.g., "1.", "3.", "5.") under each chapter heading.
  • Answers include:
    • Numerical values with units (e.g., "35 kHz", "12.48 A", "0.15 V")
    • Multiple sub-answers labeled A), B), C), etc.
    • Frequency values and relationships (e.g., "f_1 > 67 Hz due to proximity")
    • Capacitor specifications

🔢 Chapter 17 answer patterns

Chapter 17 answers follow a consistent pattern of frequency and capacitor calculations.

Answer typeExampleNotes
Frequency calculationsf_in = 67 Hz, f_out = 2.65 HzMultiple frequency values per problem
Frequency relationshipsf_1 > 67 Hz due to proximityComparative statements about cutoff frequencies
Capacitor valuesC_in = 239 nF, C_out = 440 nFValues given in nanofarads (nF) or microfarads (μF)
Design guidancePick one and increase the otherBrief instructions for component selection

⚠️ Content limitations

⚠️ No conceptual material

The excerpt does not contain any of the following:

  • Chapter objectives or learning goals
  • Concept definitions or explanations
  • Theory or background information
  • Problem-solving methods or procedures
  • Diagrams, figures, or illustrations

⚠️ Context missing

Without the original problems, the answers cannot be interpreted meaningfully.

  • The numerical values have no context (what circuits? what configurations?)
  • Terms like "proximity" appear without explanation
  • The relationship between different frequency values (f_in, f_out, f_bypass, f_1, f_2) is not defined

📖 Appendix note

📖 Unrelated content

The excerpt ends with "Appendix D" which begins a recipe for "Autumn Bread."

  • This appendix is completely unrelated to the electronics content of Chapters 14–17.
  • It includes a quote from Bill Bruford and commentary about college students' eating habits.
  • The recipe description mentions it is "a variation on banana bread, but is very low in fat."

Note: This appendix content appears to be from a different section of the textbook and does not relate to Chapter 17 objectives or electronics concepts.

112

17.1 Introduction

17.1 Introduction

🧭 Overview

🧠 One-sentence thesis

This excerpt consists entirely of answer keys and unrelated appendix material, containing no substantive technical introduction or conceptual content.

📌 Key points (3–5)

  • The excerpt contains numerical answers to end-of-chapter problems from Chapters 14–17.
  • Chapter 17 answers focus on frequency calculations (f_in, f_out, f_bypass, f_1, f_2) and capacitor values.
  • Appendix D begins with an unrelated discussion about cooking and a recipe for "Autumn Bread."
  • No conceptual explanations, definitions, or theoretical frameworks are present.
  • The material does not constitute an "introduction" in the pedagogical sense.

📋 What the excerpt contains

📋 Answer key structure

The excerpt is organized as:

  • Chapters 14–16: Miscellaneous numerical answers (voltages, frequencies, gains in dB, power values).
  • Chapter 17: Frequency and capacitor calculations with proximity notes.
  • Appendix D: A recipe and commentary on food habits.

🔢 Chapter 17 answer patterns

All Chapter 17 answers follow a similar format:

  • Input frequency (f_in), output frequency (f_out), bypass frequency (f_bypass).
  • Dominant frequency (f_1 or f_2) identified, often with a note "due to proximity."
  • Later problems provide capacitor values (C_in, C_out, C_bypass) for specified frequencies.

Example pattern:

  • Problem 1: f_in = 67 Hz, f_out = 2.65 Hz, f_bypass = 61.2 Hz. f_1 > 67 Hz due to proximity.
  • Problem 31: C_in = 239 nF, C_out = 440 nF, C_bypass = 371 μF.

⚠️ No conceptual content

  • The excerpt provides no definitions, explanations, or context for the calculations.
  • Terms like "proximity," "lag network," and "roll off at 20 dB/decade" appear without explanation.
  • There is no introduction to the chapter's topic or learning objectives.

🍞 Appendix D content

🍞 Unrelated material

Appendix D shifts abruptly to:

  • A quote: "There is life beyond the cymbals." – Bill Bruford
  • Commentary on college students' eating habits and prepackaged food.
  • A recipe introduction for "Autumn Bread," described as a low-fat variation on banana bread with optional nuts (walnuts noted as a source of alpha-linolenic acid).

Don't confuse: This appendix has no connection to the technical content of Chapters 14–17; it appears to be supplementary lifestyle material included in the textbook.

🚫 Limitations for study purposes

🚫 Why this excerpt is not useful for learning

  • No theory or concepts: Only numerical answers are provided; students cannot learn underlying principles from this material.
  • No worked examples: The answers do not show steps, reasoning, or methods.
  • Context-dependent: Without the original problems, the answers are meaningless for review.
  • Appendix irrelevance: The recipe content does not support technical learning objectives.

Recommendation: To study Chapter 17 concepts, refer to the chapter body text, worked examples, and problem statements—not the answer key alone.

113

Low Frequency Response

17.2 Low Frequency Response.

🧭 Overview

🧠 One-sentence thesis

The low-frequency response of amplifier circuits is determined by calculating the critical frequencies contributed by input, output, and bypass capacitors, with the highest of these frequencies (or their proximity effects) setting the overall lower cutoff frequency.

📌 Key points (3–5)

  • Three frequency sources: input coupling capacitor (f_in), output coupling capacitor (f_out), and bypass capacitor (f_bypass) each create their own roll-off frequency.
  • Determining the lower cutoff (f₁): the highest individual frequency typically dominates; when frequencies are close, proximity effects shift f₁ higher.
  • Common confusion: don't assume f₁ equals the highest calculated frequency—when two or more frequencies are near each other, their combined effect pushes f₁ above the highest single value.
  • Design approach: to set a desired lower cutoff, calculate required capacitor values for each position and adjust (often by picking one value and increasing others).

📐 Calculating critical frequencies

📐 Three capacitor positions

The excerpt shows three recurring frequency calculations for every amplifier problem:

  • f_in: the roll-off frequency set by the input coupling capacitor.
  • f_out: the roll-off frequency set by the output coupling capacitor.
  • f_bypass: the roll-off frequency set by the emitter/source bypass capacitor (when present).

Each capacitor forms an RC high-pass filter with circuit resistances, creating a frequency below which gain begins to fall.

🔍 Which frequency dominates

The excerpt consistently identifies which frequency sets the overall lower cutoff:

  • When one frequency is significantly higher than the others, f₁ approximately equals that frequency.
  • Example: "f_in = 477 Hz, f_out = 8.8 Hz, f_bypass = 7.42 Hz. f₁ ≈ 477 Hz."
  • When frequencies are close, the excerpt notes "due to proximity," meaning f₁ is pushed higher than the highest single frequency.
  • Example: "f_in = 67 Hz, f_out = 2.65 Hz, f_bypass = 61.2 Hz. f₁ > 67 Hz due to proximity."

🔄 Proximity effects

🔄 When frequencies are near each other

Proximity effect: when two or more critical frequencies are close in value, their combined roll-off shifts the overall lower cutoff frequency higher than the highest individual frequency.

  • The excerpt does not give an exact threshold for "close," but shows it matters when frequencies are within the same order of magnitude.
  • Example: f_in = 4.5 Hz and f_bypass = 69.4 Hz are far apart → f₁ ≈ 69.4 Hz (the higher one).
  • Example: f_in = 67 Hz and f_bypass = 61.2 Hz are close → f₁ > 67 Hz (pushed above both).

⚠️ Don't confuse

  • Single dominant frequency (f₁ ≈ highest): one frequency is much higher, so it alone sets the cutoff.
  • Proximity case (f₁ > highest): two or more frequencies are similar, so their interaction raises f₁ further.

🛠️ Design for a target cutoff

🛠️ Calculating required capacitors

The excerpt includes problems that work backward: given a desired lower cutoff frequency, calculate the needed capacitor values.

  • Example: "For 50 Hz: C_in = 902 nF, C_out = 29.9 μF. Pick one and increase the other."
  • This means: calculate each capacitor to achieve 50 Hz individually, then adjust values to account for proximity.

🛠️ Adjustment strategy

  • The excerpt advises "Pick one and increase the other(s)."
  • Reason: if you set all capacitors exactly to the target frequency, proximity will push f₁ higher than desired.
  • By increasing one or more capacitor values (lowering their individual frequencies), you compensate and bring f₁ down to the target.

📊 High-frequency response (brief mention)

📊 Upper cutoff frequencies

The excerpt also lists high-frequency calculations (f_in and f_out for the upper end):

  • Example: "f_in = 4.78 MHz, f_out = 13.6 MHz. f₂ < 4.78 Hz due to proximity."
  • The same proximity logic applies: when high-frequency roll-offs are close, the overall upper cutoff f₂ is lower than the lowest individual frequency.
  • This is the mirror image of the low-frequency case.
Frequency regionIndividual frequenciesOverall cutoffProximity effect
Low frequency (f₁)f_in, f_out, f_bypass (high-pass)Highest frequency dominatesWhen close, f₁ > highest
High frequency (f₂)f_in, f_out (low-pass)Lowest frequency dominatesWhen close, f₂ < lowest
114

High Frequency Response

17.3 High Frequency Response .

🧭 Overview

🧠 One-sentence thesis

High-frequency response analysis determines the upper cutoff frequency (f₂) of amplifier circuits by examining input and output capacitances, with proximity effects often causing the actual cutoff to differ from individual component calculations.

📌 Key points (3–5)

  • What determines high-frequency cutoff: input and output capacitances create frequency limits (f_in and f_out) that define the upper bandwidth boundary (f₂).
  • Proximity effect: when calculated f_in and f_out values are close together, the actual f₂ is lower than either individual value due to interaction between frequency-limiting mechanisms.
  • How to predict f₂: compare f_in and f_out—the lower value approximates f₂ unless they are close (proximity), in which case f₂ is even lower.
  • Common confusion: f₂ is not always equal to the smaller of f_in or f_out; proximity between the two values shifts the cutoff lower.
  • Design implications: capacitor selection at input and output stages directly controls the high-frequency performance of the amplifier.

🔍 Understanding frequency limits

🔍 Input and output frequency limits

  • f_in: the frequency limit imposed by input-stage capacitances.
  • f_out: the frequency limit imposed by output-stage capacitances.
  • These represent separate mechanisms that roll off the amplifier gain at high frequencies.
  • Each is calculated based on the capacitances and resistances at that stage.

📐 Determining the upper cutoff (f₂)

The actual upper cutoff frequency f₂ depends on how f_in and f_out compare:

RelationshipResultReason
f_in much lower than f_outf₂ ≈ f_inInput stage dominates the rolloff
f_out much lower than f_inf₂ ≈ f_outOutput stage dominates the rolloff
f_in and f_out close togetherf₂ < both valuesProximity effect: both stages interact

Example: Problem 17 shows f_in = 4.78 MHz and f_out = 13.6 MHz. Because f_in is significantly lower, the excerpt states "f₂ < 4.78 Hz due to proximity," meaning the actual cutoff is below the lower value.

Don't confuse: f₂ is not simply "the minimum of f_in and f_out." When the two are close, their combined effect pushes f₂ lower than either alone.

🧲 Proximity effects

🧲 What proximity means

  • When f_in and f_out are near each other in value, neither stage acts independently.
  • Both input and output capacitances contribute to the rolloff simultaneously.
  • The result is that the combined cutoff frequency f₂ is lower than the smaller individual limit.

🧲 Recognizing proximity in the problems

The excerpt repeatedly uses the phrase "due to proximity" to flag cases where f₂ differs from the simple prediction:

  • Problem 1: f_in = 67 Hz, f_out = 2.65 Hz, f_bypass = 61.2 Hz → "f₁ > 67 Hz due to proximity."
  • Problem 3: f_in = 9.76 Hz, f_out = 0.796 Hz, f_bypass = 2.57 Hz → "f₁ > 9.76 Hz due to proximity."
  • Problem 17: f_in = 4.78 MHz, f_out = 13.6 MHz → "f₂ < 4.78 Hz due to proximity."
  • Problem 21: f_in = 84.1 MHz, f_out = 33.9 Hz → "f₂ < 33.9 MHz due to proximity."

How to tell: if the problem states "due to proximity," the actual cutoff is shifted away from the calculated individual limits.

🔧 Design and component selection

🔧 Capacitor values control frequency response

  • Problems 31, 33, 35, and 37 ask for capacitor values (C_in, C_out, C_bypass) to achieve a desired cutoff frequency.
  • Larger capacitances lower the cutoff frequency; smaller capacitances raise it.
  • The excerpt advises "Pick one and increase the other" when designing for a specific frequency, indicating a trade-off between input and output stage capacitances.

🔧 Bypass capacitors

  • C_bypass appears alongside C_in and C_out in several problems.
  • It introduces an additional frequency limit (f_bypass) that can influence the overall cutoff.
  • Example: Problem 1 lists f_in = 67 Hz, f_out = 2.65 Hz, f_bypass = 61.2 Hz, and concludes f₁ > 67 Hz due to proximity, showing that all three limits interact.

📊 Interpreting the problem set

📊 Pattern in the answers

The problems follow a consistent structure:

  1. Calculate f_in and f_out (and sometimes f_bypass).
  2. Compare the values.
  3. State the approximate or bounded value of f₂ (or f₁ for low-frequency problems).
  4. Note "due to proximity" when the limits are close.

📊 Low-frequency vs high-frequency problems

  • Problems 1–15 appear to address low-frequency cutoff (f₁), involving input, output, and bypass capacitances.
  • Problems 17–29 address high-frequency cutoff (f₂), involving input and output capacitances at much higher frequencies (MHz range).
  • The same proximity principle applies in both regimes: when calculated limits are close, the actual cutoff shifts.

Don't confuse: f₁ (low-frequency cutoff) and f₂ (high-frequency cutoff) are separate boundaries of the amplifier bandwidth; this excerpt focuses on the high-frequency side (f₂).