Electromagnetics

1

What is Electromagnetics?

1.1 What is Electromagnetics?

🧭 Overview

🧠 One-sentence thesis

Applied engineering electromagnetics bridges the gap between basic circuit theory and the more general theory needed to address problems where material properties and geometry matter, becoming essential at higher frequencies and in applications where electromagnetic fields and waves are primary engineering considerations.

📌 Key points (3–5)

  • Why circuit theory isn't enough: lumped-element models (e.g., a resistor as simply R) hide material properties and geometry, making it impossible to answer design questions about power handling, parasitic reactance, or how to achieve specific values.
  • Electromagnetics as generalization: it extends circuit theory by including what disappears in lumped models—material properties and geometry—while circuit theory becomes a special case when these don't matter.
  • Frequency and wavelength: electromagnetic considerations become unavoidable as frequency increases, because wavelength shrinks; when circuit dimensions approach or exceed a wavelength, different parts see different signal phases.
  • Common confusion—when to use which: if wavelength is large compared to the region of interest, DC-like circuit analysis works; when dimensions are comparable to wavelength, electromagnetic analysis is required.
  • Primary vs secondary role: in some applications (antennas, fiber optics, radar), fields and waves are the main engineering focus, not just corrections to circuit models.

🔌 The limits of lumped-element circuit theory

🔌 What lumped models hide

Lumped element abstraction: a device model (e.g., resistor as V = IR with value R) that completely describes behavior without requiring knowledge of underlying electromagnetic principles like electrical potential, conduction current, or resistance.

  • This abstraction greatly simplifies analysis and design—students can work with resistors knowing only R.
  • However, it makes certain practical questions impossible to answer:
    • What determines R? How do you design a resistor for a specific resistance?
    • Why do resistors have power ratings (1/8-W, 1/4-W)? How do you adjust this?
    • Why do real resistors show reactance as well as resistance? How is this determined and mitigated?
    • Why do non-resistor components (pins, interconnects) also exhibit resistance and reactance?

🧱 What's missing: materials and geometry

  • The answers to all these questions require properties of materials and the geometry in which those materials are arranged.
  • These are precisely what disappear in lumped-element models.
  • The same limitation applies to capacitors, inductors, transformers, and any device exhibiting unintentional capacitance, inductance, or mutual impedance.

🌐 Electromagnetics as generalization of circuit theory

🌐 The relationship between the two theories

PerspectiveDescription
Electromagnetics as generalizationExtends circuit theory to address material properties and geometry
Circuit theory as special caseApplies when material properties and geometry are not important
  • Many instances of this "generalization vs. special case" dichotomy appear throughout electromagnetics.
  • Example: when you can ignore how a resistor is physically constructed, circuit theory suffices; when you need to design the resistor or predict its high-frequency behavior, electromagnetics is required.

🎯 When fields and waves are primary

  • Beyond generalizing circuit theory, some devices and applications require direct treatment of electromagnetic fields and waves as the main engineering concern.
  • Examples include:
    • Electrical generators and motors
    • Antennas
    • Printed circuit board stackup and layout
    • Persistent data storage (hard drives)
    • Fiber optics
    • Radio, radar, remote sensing, and medical imaging systems
    • Signal integrity and electromagnetic compatibility (EMC)

📏 The role of frequency and wavelength

📏 Why frequency matters

  • Electromagnetic considerations pertain to all frequencies but become increasingly difficult to avoid with increasing frequency.
  • Reason: wavelength decreases as frequency increases.

📏 Wavelength vs. circuit dimensions

  • When wavelength is large compared to the region of interest (e.g., a circuit):

    • Analysis and design resemble zero-frequency ("DC") methods.
    • Example: at 3 MHz, free-space wavelength ≈ 100 m; a 10 cm × 10 cm circuit is only 0.1% of a wavelength across.
    • An electromagnetic wave present has about the same value over the entire circuit.
  • When wavelength is comparable to circuit dimensions:

    • Different parts of the circuit observe the same signal with very different magnitude and phase.
    • Example: at 3 GHz, free-space wavelength ≈ 10 cm; the same 10 cm × 10 cm circuit is one full wavelength across.
    • Behaviors associated with non-negligible dimensions can be undesirable if not accounted for in design.
    • These behaviors can also be exploited (e.g., to launch waves via antennas, or create filters and impedance matching devices from metallic shapes alone, without discrete capacitors or inductors).

⚡ High-frequency and non-electrical bands

  • Above a few hundred MHz, and especially in millimeter-wave, infrared (IR), optical, and ultraviolet (UV) bands, electromagnetic considerations become central to analysis and design.
  • Ironically, applications in these ranges may not operate on principles considered "electrical," yet electromagnetic theory still applies.
  • Don't confuse: "electrical engineering" encompasses these frequency ranges even when the physics isn't strictly "electrical."

📋 Topics requiring electromagnetic principles

📋 Applications list

The excerpt provides an alphabetical list of topics where explicit electromagnetic consideration is important or essential:

  • Antennas
  • Coaxial cable
  • Design and characterization of discrete passive components (resistors, capacitors, inductors, diodes)
  • Distributed (e.g., microstrip) filters
  • Electromagnetic compatibility (EMC)
  • Fiber optics
  • Generators
  • Magnetic resonance imaging (MRI)
  • Magnetic storage (data)
  • Microstrip transmission lines
  • Modeling non-ideal behaviors of discrete components
  • Motors
  • Non-contact sensors
  • Photonics
  • Printed circuit board stackup and layout
  • Radar
  • Radio wave propagation
  • Radio frequency electronics
  • Signal integrity
  • Transformers
  • Waveguides

🎓 Definition and learning objectives

🎓 Formal definition

Applied engineering electromagnetics: the study of those aspects of electrical engineering in situations in which the electromagnetic properties of materials and the geometry in which those materials are arranged is important. This requires understanding electromagnetic fields and waves, which are of primary interest in some applications.

🎓 Two broad learning goals

  1. Learn techniques of engineering analysis and design that apply when electromagnetic principles are important.
  2. Better understand the physics underlying electrical devices and systems, so that when issues associated with these physical principles emerge, one is prepared to recognize and address them.
2

Electromagnetic Spectrum

1.2 Electromagnetic Spectrum

🧭 Overview

🧠 One-sentence thesis

The electromagnetic spectrum spans at least 20 orders of magnitude from DC to beyond 10²⁰ Hz, and classical electromagnetic theory applies well to most practical problems from DC through optical frequencies, though quantum effects become important at higher frequencies.

📌 Key points (3–5)

  • Spectrum range: Electromagnetic fields exist from 0 Hz (DC) to at least 10²⁰ Hz, divided into named bands (radio, infrared, optical, ultraviolet, X-rays, gamma rays) based on how they manifest physically.
  • DC vs higher frequencies: At DC, electromagnetics splits into electrostatics (electric fields) and magnetostatics (magnetic fields); at higher frequencies, electric and magnetic fields interact to form propagating waves.
  • Wavelength relationship: In free space, wavelength λ equals c divided by frequency f, where c (speed of light) is approximately 3.00 × 10⁸ m/s for any electromagnetic wave, not just visible light.
  • Classical vs quantum limits: Classical electromagnetic theory (presented in the book) works well for DC, radio, IR, and optical waves, but quantum mechanical effects become important at very high frequencies (usually X-ray and above, sometimes lower).
  • Common confusion: The "speed of light" c is actually the speed of all electromagnetic waves in free space, not just optical waves.

📡 Structure of the electromagnetic spectrum

📡 Overall frequency range

The electromagnetic spectrum: the various forms of electromagnetic phenomena that exist over the continuum of frequencies.

  • Spans from DC (0 Hz) to at least 10²⁰ Hz—at least 20 orders of magnitude.
  • Each frequency range is named based on how electromagnetic waves manifest as physical phenomena.
  • The continuum is divided into: radio, infrared (IR), optical (light), ultraviolet (UV), X-rays, and gamma rays (γ-rays), in order of increasing frequency.

🌈 Named bands and their ranges

BandFrequency RangeWavelength Range
γ-Ray> 3 × 10¹⁹ Hz< 0.01 nm
X-Ray3 × 10¹⁶ – 3 × 10¹⁹ Hz10–0.01 nm
Ultraviolet (UV)2.5 × 10¹⁵ – 3 × 10¹⁶ Hz120–10 nm
Optical4.3 × 10¹⁴ – 2.5 × 10¹⁵ Hz700–120 nm
Infrared (IR)300 GHz – 4.3 × 10¹⁴ Hz1 mm – 700 nm
Radio3 kHz – 300 GHz100 km – 1 mm
  • These ranges are arbitrary but consistent with common usage.
  • The radio portion alone spans 12 orders of magnitude, so it exhibits a broad range of phenomena and is further subdivided.

📻 Radio spectrum subdivisions

BandFrequenciesWavelengthsTypical Applications
EHF30–300 GHz10–1 mm60 GHz WLAN, Point-to-point data links
SHF3–30 GHz10–1 cmTerrestrial & Satellite data links, Radar
UHF300–3000 MHz1–0.1 mTV broadcasting, Cellular, WLAN
VHF30–300 MHz10–1 mFM & TV broadcasting, LMR
HF3–30 MHz100–10 mGlobal terrestrial comm., CB Radio
MF300–3000 kHz1000–100 mAM broadcasting
LF30–300 kHz10–1 kmNavigation, RFID
VLF3–30 kHz100–10 kmNavigation
  • Band identification uses common acronyms (EHF, SHF, UHF, VHF, HF, MF, LF, VLF).
  • Applications vary widely across bands due to different propagation and interaction properties.

🎨 Optical spectrum subdivisions

BandFrequenciesWavelengths
Violet668–789 THz450–380 nm
Blue606–668 THz495–450 nm
Green526–606 THz570–495 nm
Yellow508–526 THz590–570 nm
Orange484–508 THz620–590 nm
Red400–484 THz750–620 nm
  • The optical band is partitioned into the familiar "rainbow" colors.
  • Other portions of the spectrum are sometimes similarly subdivided in certain applications.

⚡ DC electromagnetics

⚡ Two distinct disciplines at zero frequency

  • At DC (0 Hz), electromagnetics consists of two separate disciplines:
    • Electrostatics: concerned with electric fields.
    • Magnetostatics: concerned with magnetic fields.
  • At DC, electric and magnetic fields do not interact to form waves.

🌊 Higher frequencies: wave formation

  • At higher frequencies (above DC), electric and magnetic fields interact to form propagating waves.
  • This interaction is what creates the various named bands of the electromagnetic spectrum.

🚀 Wave propagation and wavelength

🚀 Speed of electromagnetic waves

Phase velocity c: the speed at which electromagnetic fields propagate in free space, approximately 3.00 × 10⁸ m/s.

  • This value is often called the "speed of light."
  • Don't confuse: c is the speed of any electromagnetic wave in free space, not just optical (visible light) waves.
  • Example: Radio waves, infrared waves, and gamma rays all travel at c in free space.

📏 Wavelength calculation

  • Given frequency f, wavelength λ in free space is given by: λ equals c divided by f.
  • Higher frequency → shorter wavelength.
  • Lower frequency → longer wavelength.
  • Example: Radio waves at 3 kHz have wavelengths up to 100 km; gamma rays above 3 × 10¹⁹ Hz have wavelengths less than 0.01 nm.

🔬 Applicability of classical theory

🔬 Where classical theory works

  • The classical version of electromagnetic theory presented in the book applies to:
    • DC
    • Radio
    • Infrared (IR)
    • Optical waves
    • To a lesser extent: UV waves, X-rays, and γ-rays.
  • Classical theory works well for most practical problems.

⚛️ Quantum mechanical limits

  • At very high frequencies, wavelengths become small enough that quantum mechanical effects may be important.
  • This is usually the case in the X-ray band and above.
  • In some applications, quantum effects become important at frequencies as low as the optical, IR, or radio bands.
  • Example: The photoelectric effect is a quantum phenomenon that can occur at optical frequencies.

⚠️ Caution required

  • Caution is required when applying classical electromagnetic theory at higher frequencies, especially in the optical band and above.
  • Certain phenomena in these frequency ranges—in particular quantum mechanical effects—are not addressed in the book.
  • Don't confuse: Classical theory's applicability is not a sharp cutoff; it depends on the specific application and whether quantum effects matter for that problem.
3

Fundamentals of Waves

1.3 Fundamentals of Waves

🧭 Overview

🧠 One-sentence thesis

Waves—whether sound or electromagnetic—are described by the same mathematical wave equation, and sinusoidal waves propagate with a characteristic wavelength, wavenumber, and phase velocity that determine how disturbances travel through space.

📌 Key points (3–5)

  • Sound and EM waves share mathematics: sound waves are pressure variations and electromagnetic waves are field variations, but both obey the same wave equation.
  • Sinusoidal waves have three linked parameters: wavenumber β (spatial frequency), wavelength λ (distance per cycle), and phase velocity v_p (speed of propagation), related by λ = 2π/β and v_p = λf.
  • Sign of βx determines direction: if phase decreases with βx (i.e., "−βx"), the wave travels in the +x direction; if phase increases with βx (i.e., "+βx"), the wave travels in the −x direction.
  • Common confusion: wavelength vs wavenumber—wavelength is the physical distance per cycle (meters), wavenumber is the rate of phase change per meter (rad/m); they are inversely related.
  • EM waves differ from sound: electromagnetic waves are vector quantities (have direction and magnitude) and involve both electric and magnetic fields, but the same wave concepts apply.

🌊 The wave equation and what it describes

🌊 What a wave is

  • A wave is a traveling disturbance in some quantity over space and time.
  • The excerpt uses sound as an analogy: sound is a variation in air pressure p(x,y,z,t) relative to the quiescent (quiet) pressure.
  • When there is no sound, p = 0 everywhere; a clap creates a pulse of increased pressure (p > 0) followed by decreased pressure (p < 0) that travels outward.

📐 The acoustic wave equation

The acoustic wave equation: ∂²p/∂x² − (1/c_s²) ∂²p/∂t² = 0, where c_s is the speed of sound (about 340 m/s at sea level).

  • This equation governs how pressure disturbances propagate.
  • The excerpt emphasizes that the acoustic wave equation is mathematically identical to the equations governing electromagnetic waves.
  • The only difference is the phase velocity: sound travels at ~340 m/s, electromagnetic waves travel much faster.

🎺 Transient vs sinusoidal waves

  • Transient wave: a single pulse (like a clap) that travels outward and eventually dissipates.
  • Sinusoidal wave: a continuous oscillation at a fixed frequency (like blowing a trumpet note at 440 Hz).
  • The excerpt focuses on sinusoidal waves because they are common in electromagnetics and easier to analyze mathematically.

🎵 Sinusoidal waves and their parameters

🎵 The general form of a sinusoidal wave

  • For a sinusoidal source (e.g., a trumpet playing 440 Hz), the pressure field takes the form:
    • p(x,t) = A_m cos(ωt − βx + ψ)
    • ω = 2πf is the angular frequency (rad/s)
    • β is the wavenumber (rad/m)
    • A_m is the amplitude (determined by source strength)
    • ψ is the initial phase (determined by when and where the source started)
  • Two key observations justify this form:
    • At any fixed position x, p varies sinusoidally in time (because the source is sinusoidal and the wave equation is linear and time-invariant).
    • At any fixed time t, p varies sinusoidally in space (because the wave propagates with a phase shift proportional to distance).

🔢 Wavenumber β

Wavenumber β (rad/m) is the rate at which the phase of a sinusoidal wave progresses with distance.

  • β is also called the phase propagation constant or spatial frequency.
  • It plays the same role for space (x) as ω plays for time (t).
  • From the wave equation, β = ω/c_s (for sound) or β = ω/c (for EM waves, where c is the speed of light).
  • Example: for 440 Hz sound, β ≈ 8.13 rad/m, meaning the phase changes by 8.13 radians per meter of distance.

📏 Wavelength λ

Wavelength λ = 2π/β is the distance required for the phase of a sinusoidal wave to increase by one complete cycle (2π rad) at any given time.

  • Wavelength is the spatial period of the wave.
  • It is inversely related to wavenumber: λ = 2π/β.
  • Example: for 440 Hz sound with β ≈ 8.13 rad/m, λ ≈ 77.3 cm.
  • Don't confuse: wavelength is a distance (meters), wavenumber is a rate (rad/m).

🚀 Phase velocity v_p

Phase velocity v_p = λf is the speed at which a point of constant phase in a sinusoidal waveform travels.

  • Phase velocity is the speed at which the wave pattern moves through space.
  • It can be derived by observing how far a point of constant phase shifts in a given time.
  • Example: in the 440 Hz sound wave, a point of constant phase shifts λ/4 in time 1/(4f), so v_p = λf ≈ 340 m/s (the speed of sound).
  • The excerpt notes that "velocity" technically means speed in a direction, but here it is used as a scalar (speed only).
ParameterSymbolUnitsMeaningRelation
FrequencyfHzCycles per secondω = 2πf
Angular frequencyωrad/sPhase change per secondω = 2πf
Wavenumberβrad/mPhase change per meterβ = ω/c_s
WavelengthλmDistance per cycleλ = 2π/β
Phase velocityv_pm/sSpeed of wave patternv_p = λf

🧭 Direction of propagation

🧭 How the sign of βx determines direction

  • The general form p(x,t) = A_m cos(ωt − βx + ψ) has "−βx" in the argument.
  • This means the phase decreases as x increases (for fixed t).
  • To keep the phase constant (i.e., to follow a point of constant phase), x must increase as t increases → the wave travels in the +x direction.

🔄 Reversing the sign

  • If we write p(x,t) = A_m cos(ωt + βx + ψ), the phase increases as x increases.
  • To keep the phase constant, x must decrease as t increases → the wave travels in the −x direction.
  • Both forms satisfy the wave equation; the choice depends on the physical situation.
  • Example: the trumpet sound travels away from the source (+x direction), but an echo from a wall would be a wave traveling back (−x direction, with "+βx").

🎯 Rule of thumb

  • Phase decreasing with βx (i.e., "−βx") → wave propagates in the +x direction.
  • Phase increasing with βx (i.e., "+βx") → wave propagates in the −x direction.

🔬 Electromagnetic waves vs sound waves

🔬 Similarities

  • Electromagnetic waves obey the same wave equation as sound waves (Equation 1.1).
  • All the concepts—wavenumber, wavelength, phase velocity, direction of propagation—apply identically to EM waves.
  • Example: an EM wave with λ = 77.3 cm (same as the 440 Hz sound wave) lies in the radio portion of the spectrum.

⚡ Key differences

  • Nature of the disturbance:
    • Sound waves are variations in pressure (a scalar quantity).
    • Electromagnetic waves are variations in electric and magnetic fields (vector quantities with direction and magnitude).
  • Phase velocity:
    • Sound: ~340 m/s at sea level.
    • EM waves: much greater (speed of light, ~3×10⁸ m/s in vacuum).
  • Frequency and wavelength:
    • EM waves have much higher frequencies than sound, but because the phase velocity is also much higher, wavelengths can be similar in order of magnitude.
  • Multiple fields:
    • To fully describe an EM wave, we often need to consider both the electric field and the magnetic field.
    • Despite this complexity, the same wave equation and parameters (β, λ, v_p) apply.

🌐 Practical implication

  • The excerpt emphasizes that the analogy between sound and EM waves is strong enough that understanding sound waves provides useful insight into EM waves.
  • Don't confuse: sound and EM waves are completely distinct phenomena (pressure vs fields), but they share the same mathematical structure.
4

Guided and Unguided Waves

1.4 Guided and Unguided Waves

🧭 Overview

🧠 One-sentence thesis

Waves are classified as either guided (constrained to follow a structure like transmission lines or optical fibers) or unguided (propagating freely after emission from antennas or unintentional radiators until redirected or dissipated).

📌 Key points (3–5)

  • The fundamental distinction: guided waves are constrained by physical structures; unguided waves propagate freely in an uncontrolled manner.
  • Examples of unguided waves: radiated by antennas or unintentionally radiated sources; they continue until scattered or dissipated by material losses.
  • Examples of guided waves: exist within transmission lines, waveguides, and optical fibers; they follow the path defined by the structure.
  • Common confusion: "guided" does not mean "controlled remotely"—it means physically constrained to a path by a structure.

🌊 The two categories of wave propagation

🌐 Unguided waves

Unguided waves: waves that propagate in an uncontrolled manner until they are redirected by scattering or dissipated by losses associated with materials.

  • These waves are initiated (radiated) and then travel freely through space.
  • Two main sources:
    • Antennas: intentionally emit electromagnetic waves into free space.
    • Unintentional radiators: emit waves as a side effect (not the primary purpose).
  • Once initiated, the wave's path is not predetermined—it spreads outward and may be affected by obstacles, reflections, or absorption.
  • Example: a radio antenna broadcasts a signal; the wave travels through the air until it hits a building (scattering) or is absorbed by materials.

🛤️ Guided waves

Guided waves: waves that exist within structures such as transmission lines, waveguides, and optical fibers; they are constrained to follow the path defined by the structure.

  • The physical structure acts as a "guide" that confines the wave to a specific path.
  • Three main types of guiding structures:
    • Transmission lines: carry electromagnetic waves along conductors (e.g., coaxial cables).
    • Waveguides: hollow metal tubes or dielectric structures that channel waves.
    • Optical fibers: thin strands of glass or plastic that guide light waves via total internal reflection.
  • The wave cannot freely propagate in all directions—it is physically bound to the structure.
  • Example: an optical fiber carries a light signal from a transmitter to a receiver; the light stays inside the fiber and follows its bends and turns.

🔍 How to distinguish guided from unguided

🔍 Key criterion: structural constraint

  • Guided: the wave's path is determined by a physical structure; the wave cannot leave the structure without special conditions (e.g., leakage or intentional coupling).
  • Unguided: no structure confines the wave; it spreads according to the medium and any obstacles encountered.

🔍 Behavior after initiation

TypeAfter initiationPath controlExamples
UnguidedPropagates freely; affected by scattering, absorption, or reflectionNo predetermined pathAntenna radiation, unintentional emissions
GuidedConstrained to follow the structurePath defined by the structureTransmission lines, waveguides, optical fibers
  • Don't confuse: "guided" does not mean the wave is steered remotely or actively controlled—it simply means the wave is physically confined to a structure.
  • Example: a waveguide does not "steer" the wave in real time; it passively constrains the wave to travel along its interior.

📡 Practical implications

📡 Unguided wave applications

  • Used when signals must travel through open space without physical connections.
  • Wireless communication (radio, TV, cellular) relies on unguided waves radiated by antennas.
  • Challenges: waves spread out, lose energy over distance, and are subject to interference and scattering.

📡 Guided wave applications

  • Used when signals must travel along a specific route with minimal loss and interference.
  • Wired communication (internet cables, telephone lines) and fiber-optic networks rely on guided waves.
  • Advantages: the structure protects the wave from external interference and confines energy to the intended path.
  • Challenges: requires physical infrastructure; the structure itself may introduce losses or distortions.
5

Phasors

1.5 Phasors

🧭 Overview

🧠 One-sentence thesis

Phasor representation simplifies the analysis of sinusoidal signals by encoding magnitude and phase as a single complex number, enabling easier mathematical operations and extending to arbitrary signals through Fourier analysis.

📌 Key points (3–5)

  • What a phasor is: a complex-valued number that represents a real-valued sinusoidal waveform by capturing its magnitude and phase.
  • Why phasors simplify analysis: operations on sinusoids (differentiation, integration, finding peak values) become much easier when performed on phasors instead of time-domain signals.
  • How to convert: multiply the phasor by e^(jωt) and take the real part to recover the physical sinusoidal signal.
  • Common confusion: a phasor is not the signal itself—it is a simplified mathematical representation; the actual physical signal is real-valued, while the phasor is complex-valued.
  • Broader applicability: phasor analysis applies not only to pure sinusoids but also to narrowband signals and arbitrary-bandwidth signals via Fourier analysis and superposition.

🔄 What phasors represent

🔄 Definition and core idea

A phasor is a complex-valued number that represents a real-valued sinusoidal waveform. Specifically, a phasor has the magnitude and phase of the sinusoid it represents.

  • A phasor is not the physical signal; it is a simplified mathematical representation.
  • The actual physical signal is real-valued; the phasor is complex-valued and constant (does not vary with time).
  • Example: A sinusoid with amplitude A_m and phase ψ is represented by the phasor C = A_m e^(jψ).

📐 Mathematical form

A general physical sinusoidal quantity varying with angular frequency ω = 2πf is:

A(t; ω) = A_m(ω) cos(ωt + ψ(ω))

where:

  • A_m(ω) is the magnitude at the specified frequency
  • ψ(ω) is the phase at the specified frequency
  • t is time
  • The magnitude A_m does not vary with time (∂A_m/∂t = 0); all time variation is in the cosine function

The equivalent phasor representation is:

C(ω) = A_m(ω) e^(jψ(ω))

🔁 Converting back to the physical signal

To recover the physical signal from a phasor:

  1. Restore time dependence by multiplying by e^(jωt)
  2. Take the real part of the result

In notation: A(t; ω) = Re{C(ω) e^(jωt)}

  • Commonly, the explicit frequency dependence is dropped and written as C = A_m e^(jψ), with the understanding that C and ψ apply at whatever frequency is represented.

📊 Examples of phasor representations

Physical signal A(t)Phasor C
A_m cos(ωt)A_m
A_m cos(ωt + ψ)A_m e^(jψ)
A_m sin(ωt) = A_m cos(ωt - π/2)-jA_m
A_m cos(ωt) + B_m sin(ωt)A_m - jB_m

(A_m and B_m are real-valued and constant with respect to t)

⚡ Why phasors simplify analysis

⚡ Easier extraction of signal properties

Without phasors:

  • Finding the peak value requires a time-domain search over one period of the sinusoid
  • Finding the phase ψ requires correlation (integration) over one period

With phasors:

  • The peak value is simply the magnitude |C|—no search required
  • The phase ψ is simply the phase of C—no integration required

🧮 Operations on phasors replace operations on signals

Mathematical operations applied to the time-domain signal A(t; ω) can be equivalently performed as operations on the phasor C, and the latter are typically much easier.

Two key claims justify this:

Claim 1: If two phasors are equal, then the sinusoidal waveforms they represent are also equal.

  • Formally: if Re{C₁ e^(jωt)} = Re{C₂ e^(jωt)} for all t, then C₁ = C₂.
  • This means phasor equality implies signal equality.

Claim 2: For any real-valued linear operator T (addition, multiplication by a constant, differentiation, integration, etc.) and complex-valued quantity C: T(Re{C}) = Re{T(C)}

  • Example with differentiation: Re{∂C/∂ω} = ∂(Re{C})/∂ω
  • This means you can differentiate (or integrate, etc.) the phasor directly without transforming back to the time-domain signal.

🎯 Summary of the simplification

Claims 1 and 2 together entitle us to perform operations on phasors as surrogates for the physical, real-valued, sinusoidal waveforms they represent.

  • Once operations are done, you can transform the resulting phasor back to the physical waveform using Re{C e^(jωt)}, if desired.
  • However, a final transformation back to the time domain is usually not desired, since the phasor tells us everything we can know about the corresponding sinusoid.

🌐 Applicability beyond pure sinusoids

🌐 Context: LTI systems and superposition

  • Many engineering signals are well-modeled as sinusoids.
  • Devices that process these signals are often well-modeled as linear time-invariant (LTI) systems.
  • The response of an LTI system to any linear combination of sinusoids is another linear combination of sinusoids having the same frequencies.

This means:

  1. Sinusoidal signals processed by LTI systems remain sinusoids (not transformed into square waves or other waveforms).
  2. You can calculate the response for one sinusoid at a time, then add the results to find the response when multiple sinusoids are applied simultaneously.

This property of LTI systems is known as superposition.

📡 Narrowband signals

Many signals, although not strictly sinusoidal, are "narrowband" and therefore well-modeled as sinusoidal.

  • Example: A cellular telecommunications signal might have a bandwidth of about 10 MHz and a center frequency of about 2 GHz.
  • The difference in frequency between the band edges is just 0.5% of the center frequency.
  • The frequency response associated with signal propagation or hardware can often be assumed constant over this range.
  • With some caveats, doing phasor analysis at the center frequency and assuming the results apply equally well over the bandwidth of interest is often a pretty good approximation.

🔬 Arbitrary-bandwidth signals via Fourier analysis

Phasor analysis is easily extensible to any physical signal, regardless of bandwidth.

  • Any physical signal can be decomposed into a linear combination of sinusoids—this is known as Fourier analysis.
  • The way to find this linear combination is by computing the Fourier series (if the signal is periodic) or the Fourier Transform (otherwise).
  • Phasor analysis applies to each frequency independently.
  • Invoking superposition, the results can be added together to obtain the result for the complete signal.

The process of combining results after phasor analysis is integration over frequency:

∫ A(t; ω) dω from -∞ to +∞

Using the phasor representation A(t; ω) = Re{C(ω) e^(jωt)}, this becomes:

∫ Re{C(ω) e^(jωt)} dω from -∞ to +∞

Using Claim 2, this can be rewritten as:

Re{∫ C(ω) e^(jωt) dω from -∞ to +∞}

The quantity in the curly braces is simply the Fourier transform of C(ω).

🎓 Conclusion on applicability

Phasor analysis does not limit us to sinusoidal waveforms. Phasor analysis is not only applicable to sinusoids and signals that are sufficiently narrowband, but is also applicable to signals of arbitrary bandwidth via Fourier analysis.

  • You can analyze a signal of arbitrarily-large bandwidth by keeping ω as an independent variable while doing phasor analysis.
  • If you ever need the physical signal, just take the real part of the Fourier transform of the phasor.
  • Not only is it possible to analyze any time-domain signal using phasor analysis, it is also often far easier than doing the same analysis on the time-domain signal directly.
6

Units

1.6 Units

🧭 Overview

🧠 One-sentence thesis

Units are essential for expressing physical quantities unambiguously, and including them in all expressions prevents errors and enables dimensional error-checking.

📌 Key points (3–5)

  • What units are: the measure used to express a physical quantity (e.g., meters for distance).
  • Why prefixes matter: standard prefixes (kilo-, mega-, nano-, etc.) make large or small numbers easier to write and read.
  • Critical practice: always indicate units explicitly to avoid ambiguity and errors in calculations.
  • Common confusion: the same number can mean different things depending on units (e.g., "3" could be 3 m/s or 3 km/h).
  • Error-checking tool: dimensional correctness helps verify whether an expression can possibly be correct.

📏 What units express

📏 Definition and purpose

Unit: the measure used to express a physical quantity.

  • A unit gives meaning to a number by specifying what is being measured.
  • Example: "6,371,000 meters" describes Earth's mean radius—the number alone is meaningless without "meters."
  • Without units, expressions become ambiguous and prone to misinterpretation.

🔤 Standard abbreviations

  • Writing full unit names repeatedly is tedious, so standard abbreviations are used.
  • Example: "km" for kilometer, "m" for meter, "s" for second.
  • The excerpt provides tables of common prefixes and base units used in electromagnetics.

🔢 Prefixes and scaling

🔢 Why prefixes exist

  • Large or small numbers become cumbersome (e.g., "6,371,000 meters").
  • Prefixes modify units to yield values in a manageable range (0.001 to 10,000).
  • Example: 6,371,000 meters = 6371 kilometers (using the "kilo-" prefix for ×1000).

📊 Common prefixes

PrefixAbbreviationMultiply by
exaE10¹⁸
petaP10¹⁵
teraT10¹²
gigaG10⁹
megaM10⁶
kilok10³
millim10⁻³
microμ10⁻⁶
nanon10⁻⁹
picop10⁻¹²
femtof10⁻¹⁵
attoa10⁻¹⁸

⚠️ Ambiguity and errors

⚠️ The danger of omitting units

  • Failing to indicate units is a common source of error and misunderstanding.
  • The excerpt emphasizes: "it is important to always indicate the units of a quantity."

🧮 Example of ambiguity

Consider the expression: l = 3t (where l is length and t is time).

Problem: What does "3" mean?

  • If l is in meters and t is in seconds, then "3" means "3 m/s."
  • If l is in kilometers and t is in hours, then "3" means "3 km/h."
  • The equation is literally different depending on the intended units.

Poor solutions:

  • Writing l = 3t m/s still doesn't clarify what "3" represents.
  • Writing l = 3t where l is in meters and t is in seconds is unambiguous but awkward for complex expressions.

Better solutions:

  • l = (3 m/s) t — units are explicit.
  • l = at where a = 3 m/s — separates the proportionality from the units and highlights the constant.

🌍 Unit systems

🌍 SI (International System of Units)

  • The excerpt uses SI units exclusively.
  • Also known informally as the "metric system."
  • SI defines seven fundamental units: meter (m), second (s), ampere (A), kilogram (kg), kelvin (K), mole (mol), candela (cd).
  • All other units (e.g., coulombs for charge, volts for potential) are derived from these fundamental units.

🔄 Other systems

  • English system: uses miles, pounds, etc.; still common in some regions and applications.
  • CGS system (centimeter-gram-second): similar to SI but with significant differences (e.g., energy is measured in "ergs" instead of joules; some physical constants become unitless).
  • Don't confuse: the same physical quantity can have different numerical values and units in different systems, reinforcing the need to always specify units.

✅ Dimensional error-checking

✅ How units help catch mistakes

An expression is dimensionally correct if its units match the expected units for the quantity being calculated.

  • Example: electric field intensity is measured in volts per meter (V/m).
  • If an expression for electric field yields V/m, it is dimensionally correct (though not necessarily fully correct).
  • If an expression cannot be reduced to V/m, it cannot be correct.

🛠️ Practical benefit

  • Units act as a built-in error-checking tool.
  • Dimensional analysis can reveal mistakes in derivations or calculations before numerical evaluation.
7

Notation

1.7 Notation

🧭 Overview

🧠 One-sentence thesis

This section establishes a consistent set of mathematical symbols and typographic conventions that distinguish vectors from scalars, exact from approximate equality, and different coordinate systems, enabling precise communication of electromagnetic concepts throughout the book.

📌 Key points (3–5)

  • Typographic distinctions: boldface indicates vectors, circumflex indicates unit vectors, tilde indicates phasors, and script indicates geometric entities like curves and surfaces.
  • Three coordinate systems: position can be expressed in Cartesian (x, y, z), cylindrical (ρ, φ, z), or spherical (r, θ, φ) coordinates, or coordinate-independently as r.
  • Three levels of approximation: the symbols ∼=, ≈, and ∼ represent different degrees of equality—exact but rounded, approximate even when precise, and order-of-magnitude, respectively.
  • Common confusion: distinguishing ∼= (approximately equal with rounding) from ≈ (inherently approximate) from ∼ (within a factor of 10).
  • Integration notation: single integrals with subscripts denote open paths/surfaces/volumes, while a circle on the integral sign indicates closed curves or surfaces.

✍️ Typographic conventions for quantities

✍️ Vectors vs scalars

Vector: indicated by boldface; e.g., E for electric field intensity vector.

  • Quantities not in boldface are scalars (single numbers without direction).
  • When handwriting, common alternatives are "E" with an arrow above or "E" with an underline.
  • Example: E is a vector field, but E (not bold) would be its scalar magnitude.

🎩 Unit vectors

Unit vector: indicated by a circumflex (hat symbol); a vector with magnitude equal to one.

  • Example: x̂ is the unit vector pointing in the +x direction.
  • Spoken aloud as "x hat."
  • Purpose: unit vectors show direction without magnitude, making them building blocks for expressing other vectors.

🌊 Phasors

Phasor: indicated by a tilde; e.g., Ṽ for a voltage phasor, Ẽ for the phasor representation of E.

  • Phasors are a mathematical tool for representing oscillating quantities.
  • The tilde distinguishes the phasor representation from the time-domain vector.

📐 Geometric entities

Curves, surfaces, and volumes: indicated in script font; e.g., 𝒮 for a surface, 𝒞 for a curve, 𝒱 for a volume.

  • Example: an open surface might be 𝒮, and the curve bounding it might be 𝒞.
  • A closed surface 𝒮 encloses a volume 𝒱.
  • Don't confuse: script notation is reserved for these geometric objects, not for field quantities.

📍 Position and coordinate systems

📍 Three coordinate systems

SystemSymbolsDescription
Cartesian(x, y, z)Rectangular coordinates
Cylindrical(ρ, φ, z)Radial distance, angle, height
Spherical(r, θ, φ)Radial distance, two angles
  • Coordinate-independent position: the symbol r expresses position without choosing a coordinate system.
  • Example: in Cartesian coordinates, r = x̂ x + ŷ y + ẑ z.
  • Why it matters: some physical laws are easier to state without committing to a particular coordinate system.

⏰ Time

  • The symbol t indicates time.
  • Standard across all coordinate systems and contexts.

🔢 Equality and approximation symbols

🔢 Three levels of approximation

SymbolMeaningWhen to useExample from excerpt
∼=Approximately equal toEquality exists but not expressed with exact numerical precisionπ ∼= 3.14
Approximately equal toThe two quantities are unequal even if expressed exactlye^x ≈ 1 + x for x ≪ 1
On the order ofWithin a factor of 10 or soμ ∼ 10^5 for iron alloys (exact values vary by factor of 5)

🔍 How to distinguish the three symbols

  • ∼=: The underlying value is exact (like π), but you're rounding for convenience.
    • Example: e^0.1 ∼= 1.1052 (the actual computed value, rounded).
  • : The equation itself is an approximation; even infinite precision won't make them equal.
    • Example: e^x = 1 + x + x²/2 + ... (infinite series), but e^x ≈ 1 + x (truncated approximation).
    • Using this approximation, e^0.1 ≈ 1.1, which is close to but not equal to the actual value.
  • : A weak statement; the value is in the right ballpark, possibly off by a factor of several.
    • Example: μ ∼ 10^5 means μ could be 5×10^4 or 5×10^5.

≝ Definition symbol

  • The symbol "≝" means "is defined as" or "is equal as the result of a definition."
  • Distinguishes definitions from derived equalities.

∫ Integration notation

∫ Open vs closed integrals

  • Single integral sign with subscript: integration over an open curve, surface, or volume.
    • ∫_𝒞 ... dl: integral over the curve 𝒞.
    • ∫_𝒮 ... ds: integral over the surface 𝒮.
    • ∫_𝒱 ... dv: integral over the volume 𝒱.
  • Circle superimposed on integral sign: integration over a closed curve or surface.
    • ∮_𝒞 ... dl: integral over the closed curve 𝒞.
    • ∮_𝒮 ... ds: integral over the closed surface 𝒮.

🔒 What "closed" means

Closed curve: one which forms an unbroken loop; e.g., a circle.

Closed surface: one which encloses a volume with no openings; e.g., a sphere.

  • Don't confuse: an open surface (like a disk) has a boundary curve; a closed surface (like a sphere) has no boundary.
  • Why it matters: many fundamental theorems in electromagnetics (like Gauss's law) apply specifically to closed surfaces.

🔧 Other conventions

🔧 Complex numbers

  • The symbol j represents the imaginary unit, √−1.
  • (Note: engineering texts often use j instead of i to avoid confusion with current.)

🔧 Physical constants

  • The excerpt notes that Appendix C contains notation for commonly-used physical constants.
  • This section does not define those constants, only points to where they are documented.
8

What is a Field?

2.1 What is a Field?

🧭 Overview

🧠 One-sentence thesis

Fields describe how physical quantities vary continuously across space and time, and waves are time-varying fields that transport energy independently of their source.

📌 Key points (3–5)

  • What a field is: a continuum of values of a quantity (scalar or vector, real or complex) as a function of position and time.
  • Fields in electromagnetics: electric field intensity E is a real-valued vector field; electric potential V is a scalar field; phasor representations are complex-valued but time-independent.
  • What a wave is: a time-varying field that continues to exist without its source and can transport energy.
  • Common confusion: field notation—E(x, y, z, t) or E(r, t) shows time dependence; phasor Ẽ(r) or is complex-valued but has no explicit time dependence.

🌐 Understanding fields

🌐 The field concept

A field is the continuum of values of a quantity as a function of position and time.

  • A field assigns a value to every point in space (and possibly time).
  • The value may be:
    • Scalar (a single number) or vector (magnitude + direction).
    • Real-valued or complex-valued.
  • Fields describe how a physical quantity is distributed and how it changes.

📐 Notation and representation

The excerpt uses several notations to express fields depending on their properties:

RepresentationMeaningExample
E(x, y, z, t) or E(r, t)Real-valued vector field varying with position and timeElectric field intensity in time domain
E (simplified)Same field, context implies position/time dependenceShorthand when dependencies are understood
Ẽ(r) or Complex-valued phasor with no time dependenceElectric field intensity as a phasor
V(r, t)Scalar field varying with position and timeElectric potential
  • Don't confuse: phasor notation (tilde, e.g., ) indicates complex values but removes explicit time dependence; time-domain notation includes (t).

⚡ Fields in electromagnetics

⚡ Electric field intensity E

  • E is a real-valued vector field.
  • It may vary as a function of position and time.
  • When expressed as a phasor, it becomes complex-valued but exhibits no time dependence.
  • Example: in time domain, E(r, t) describes how the electric field changes everywhere over time; in phasor form, Ẽ(r) captures spatial variation with implicit sinusoidal time behavior.

🔋 Electric potential V

  • V is a scalar field in electromagnetics.
  • Notation: V(r, t) indicates it varies with position and time.
  • Unlike E, which is a vector, V has magnitude only (no direction).

🌊 Waves as special fields

🌊 What makes a wave

A wave is a time-varying field that continues to exist in the absence of the source that created it and is therefore able to transport energy.

  • Key property: the wave persists even after the source stops or is removed.
  • Energy transport: waves carry energy from one location to another.
  • Example: an electromagnetic wave generated by an antenna continues to propagate through space, delivering energy to a receiver far from the source.

🔄 Wave vs static field

  • A static field (e.g., the electric field around a stationary charge) does not vary with time and does not transport energy away from the source.
  • A wave is time-varying and self-sustaining; it moves energy through space.
  • Don't confuse: not all time-varying fields are waves—only those that propagate independently and transport energy qualify.
9

Electric Field Intensity

2.2 Electric Field Intensity

🧭 Overview

🧠 One-sentence thesis

Electric field intensity E quantifies the force per unit charge at every point in space, and it points in the direction where electric potential decreases most rapidly, with magnitude equal to the rate of that decrease.

📌 Key points (3–5)

  • What E measures: the force experienced by a vanishingly small test charge, divided by that charge; units are N/C or equivalently V/m.
  • Two ways to understand E: (1) as force per charge on a test particle, or (2) as the rate of change of electric potential with distance.
  • Direction and magnitude: E points where potential decreases fastest; its magnitude is how quickly potential drops per meter.
  • Common confusion: E is expressed in V/m (not N/C) because 1 V/m = 1 N/C, and the voltage-per-distance interpretation is more intuitive for circuits and engineering.
  • What E does vs. what E is: classical physics defines E by its effects (force on charge); quantum mechanics reveals it as part of the electromagnetic force.

🔋 What is a field and the electric field

🌐 Field definition

A field is the continuum of values of a quantity as a function of position and time.

  • The quantity may be scalar or vector, real- or complex-valued.
  • Electric field intensity E is a real-valued vector field that varies with position and time: E(x, y, z, t) or E(r, t).
  • When expressed as a phasor (complex-valued, no time dependence), it is written as Ẽ(r).
  • Example of a scalar field in electromagnetics: electric potential V(r, t).

⚡ The broader electric field concept

  • Imagine a single positively-charged particle in an otherwise empty universe.
  • A second positive charge appears nearby → like charges repel → the first particle exerts force on the second.
  • If the second particle is free to move, it gains kinetic energy; if held in place, it has potential energy (equally "real" because releasing it converts potential to kinetic).
  • You can map every position in space to a vector describing the force a test particle with charge q would experience there → this map is the electric field.

🧲 Defining electric field intensity E

📐 The formal definition

Electric field intensity E(r) is the force F(r) experienced by a test particle with charge q, divided by q, in the limit as q approaches zero:

E(r) = limit (q → 0) of F(r) / q

  • Why the limit? The test charge itself contributes to the total field (since charge is the source of electric fields). To measure the field of interest without perturbing it, the test charge must be vanishingly small.
  • This definition is awkward from an engineering perspective (addressed later via the voltage interpretation).

🔢 Units: N/C vs V/m

  • According to the definition, E has units of force divided by charge: newtons per coulomb (N/C).
  • In practice, E is expressed in volts per meter (V/m).
  • These are equivalent: 1 V/m = 1 N/C.
    • Derivation: N/C = (N·m) / (C·m) = J / (C·m) = V/m, using 1 N·m = 1 joule (J) and 1 J/C = 1 volt (V).
  • Don't confuse: the two units are interchangeable; V/m is preferred because it connects directly to electric potential and circuit intuition.

🔌 The voltage interpretation (why V/m is preferred)

🧪 Parallel-plate capacitor thought experiment

  • A circuit with a 9 V battery connected to a parallel-plate capacitor (plates 1 mm apart).
  • The battery forces positive charge on the upper plate, negative on the lower plate.
  • Traveling from position A (lower plate) through the battery to position B (upper plate): potential increases by +9 V.
  • Traveling from B back to A through the capacitor: potential decreases by −9 V (closed-loop voltage sum is zero).
  • Rate of potential change between the plates: 9 V / 1 mm = 9000 V/m.
  • This is literally the electric field intensity between the plates.
  • Placing an infinitesimally small charge at point C between the plates: the ratio of force to charge is 9000 N/C, pointing toward A (lower potential).

🎯 The remarkable conclusion

E points in the direction in which electric potential is most rapidly decreasing, and the magnitude of E is the rate of change in electric potential with distance in this direction.

  • This interpretation is more intuitive for engineering: E tells you how quickly voltage drops per meter.
  • Example: in the capacitor, E = 9000 V/m means potential drops 9000 volts for every meter traveled in the direction of E.

🌌 What E does vs. what E is

🛠️ Classical physics perspective

  • Classical physics defines the electric field by what it does: it exerts force on charged particles.
  • This is adequate for most engineering applications.
  • We have not directly addressed what the electric field is at a fundamental level.

⚛️ Quantum mechanics perspective

  • Electric and magnetic fields are manifestations of the same fundamental force: the electromagnetic force.
  • The electromagnetic force is one of four fundamental forces (the others: gravity, strong nuclear force, weak nuclear force).
  • Quantum mechanics provides deeper insight into electric charge and the photon (the fundamental constituent of electromagnetic waves).
  • A wave is a time-varying field that exists independently of its source and can transport energy.

🔍 Don't confuse

  • Classical definition (force per charge) vs. quantum understanding (electromagnetic force manifestation) are not contradictory; the latter is a deeper explanation of the former.
  • For engineering purposes, the classical "what it does" definition is sufficient.
10

Permittivity

2.3 Permittivity

🧭 Overview

🧠 One-sentence thesis

Permittivity captures how material affects the strength of the electric field produced by a given charge, with most materials weakening the field compared to a vacuum.

📌 Key points (3–5)

  • What permittivity describes: the effect of material in determining the electric field in response to electric charge.
  • How charge creates a field: a charge q produces an electric field that increases with charge, decreases with distance squared (inverse square law), and is scaled by 1/ε (the permittivity factor).
  • Free space baseline: in a vacuum, permittivity equals ε₀ ≈ 8.854 × 10⁻¹² F/m; air is nearly the same; most other materials have higher permittivity (weaker field for the same charge).
  • Relative permittivity: materials are often described by εᵣ = ε / ε₀, which ranges from 1 (vacuum) to about 60 in common engineering applications.
  • Common confusion: "dielectric constant" is sometimes used for relative permittivity, but permittivity applies to many materials that are not dielectrics.

🔬 What permittivity does

🔬 The role of material

Permittivity (ε, F/m) describes the effect of material in determining the electric field intensity in response to charge.

  • Permittivity is the constant of proportionality that captures how the surrounding material influences the electric field.
  • It appears in the formula for the electric field produced by a point charge.
  • The excerpt emphasizes that ε depends on the material: different substances yield different values.

⚡ The electric field formula

The excerpt presents the laboratory-observed relationship:

  • Electric field E equals (unit vector R-hat) times (charge q) times (1 / 4πR²) times (1 / ε).
  • Breaking down the factors:
    • E increases with q: more charge → stronger field (charge is the source).
    • E decreases with 4πR²: the field spreads over the area of a sphere around the charge (inverse square law).
    • 1/ε scales the result: this is where material enters; higher ε → weaker E for the same charge.
  • Units: E is in V/m, q is in C, so ε must be in farads per meter (F/m). The excerpt notes that 1 F = 1 C/V.

📐 Inverse square law

  • The electric field is inversely proportional to 4πR², meaning it decreases in proportion to the area of a sphere surrounding the charge.
  • This principle is commonly known as the inverse square law.
  • Example: doubling the distance from the charge reduces the field strength by a factor of four (since area grows as R²).

🌌 Permittivity in different materials

🌌 Free space (vacuum)

  • In a perfect vacuum, permittivity equals ε₀.
  • The value is approximately 8.854 × 10⁻¹² F/m.
  • Air has only slightly greater permittivity and is usually assumed equal to free space.

🧱 Other materials

  • In most other materials, permittivity is significantly greater than ε₀.
  • Higher permittivity means the same charge results in a weaker electric field intensity.
  • The excerpt states that this is a general observation from laboratory experiments.

📊 Relative permittivity

📊 Definition and use

Relative permittivity: εᵣ = ε / ε₀

  • This is the permittivity of a material divided by the permittivity of free space.
  • It provides a dimensionless way to compare materials to the vacuum baseline.
  • Common practice: describe materials by their relative permittivity rather than absolute ε.

📊 Typical values

Material typeRelative permittivity (εᵣ)
Perfect vacuum1
Air≈ 1 (slightly greater)
Common engineering materialsUp to about 60
  • The excerpt notes that values for representative materials are given in Appendix A.1.
  • The range is from 1 (vacuum) to roughly 60 in typical applications.

⚠️ Terminology caution

  • Relative permittivity is sometimes called "dielectric constant."
  • Don't confuse: this term is misleading because permittivity is a meaningful concept for many materials that are not dielectrics.
  • The excerpt warns that "dielectric constant" suggests a narrower scope than the actual applicability of permittivity.
11

Electric Flux Density

2.4 Electric Flux Density

🧭 Overview

🧠 One-sentence thesis

Electric flux density D provides an alternative description of electric fields that remains constant with distance and simplifies analysis at boundaries between different materials, especially conductors.

📌 Key points (3–5)

  • What flux density is: the quantity D = ε E, measured in C/m², describes the electric field in terms of flux rather than force or potential.
  • Why flux matters: the flux of E through a sphere (integral of E over the surface) equals q/ε and stays constant regardless of sphere radius, even though E itself decreases with distance.
  • When D becomes essential: at boundaries between materials with different permittivities, D constrains the perpendicular component of the field; with perfect conductors, D alone determines the entire field.
  • Common confusion: D has units of C/m² but does not describe actual surface charge density—it represents an equivalent charge density that would produce the observed field.
  • Practical advantage: using D greatly simplifies problems involving conductors and capacitors.

🔄 From electric field to flux

🔄 The inverse square law and integration

  • Electric field intensity E from a point charge q is inversely proportional to 4πR² (the area of a sphere at distance R).
  • This is the inverse square law: E decreases as the surrounding sphere's area increases.
  • When you integrate E over the entire sphere, the result is q/ε, which does not depend on R.

🌊 What flux means

Flux: the integral of a vector field over a specified surface.

  • The flux of E through a sphere surrounding charge q equals q/ε.
  • Key insight: flux stays constant with distance, even though E itself weakens.
  • Example: a larger sphere has weaker E at each point, but the total flux through the sphere remains the same.

📐 Defining electric flux density

📐 The definition

Electric flux density: D = ε E, with units of C/m², describes the electric field in terms of flux rather than force or change in potential.

  • By multiplying E by permittivity ε, the flux of D through a sphere equals the enclosed charge q directly (without the 1/ε factor).
  • Mathematically: the integral of D over sphere S equals q.

🔍 Why introduce D when we already have E?

  • In homogeneous media (same material everywhere), D is redundant—you can always compute it from E and ε.
  • D becomes crucial at boundaries between materials with different permittivities.
  • Don't confuse: D is not just a rescaled E; it reveals different boundary behavior.

🧱 Boundaries and boundary conditions

🧱 Perpendicular vs tangential components

  • At a boundary between two materials, D constrains the perpendicular (normal) component of the electric field.
  • If you only consider E, you constrain only the tangential component.
  • Both constraints are needed to fully determine the field at a boundary.

⚡ Perfect conductors

  • When one material is a perfect conductor (e.g., many metals), the boundary condition on D completely determines the electric field.
  • This makes analysis much simpler: you don't need to solve for both components separately.
  • Example: analyzing capacitors (devices with conductor plates) becomes straightforward using D.
ScenarioWhat D providesBenefit
Boundary between two dielectricsConstraint on perpendicular field componentComplements tangential constraint from E
Boundary with perfect conductorComplete determination of the fieldGreatly simplifies finding the electric field

⚠️ Common confusion: D vs surface charge

⚠️ Units do not imply meaning

  • D has units of C/m², the same units used for surface charge density.
  • Warning: D describes an electric field, not actual surface charge.
  • There is no implication that the field source is a distribution of surface charge.

🔄 Equivalent vs actual charge

  • D can be interpreted as an equivalent surface charge density that would produce the observed field.
  • In some cases, this equivalent density happens to match the actual charge density.
  • In other cases, it does not—D is a field description, not a charge inventory.
  • Example: a region with D present may have no actual charges on its surfaces; D simply represents the field configuration.
12

Magnetic Flux Density

2.5 Magnetic Flux Density

🧭 Overview

🧠 One-sentence thesis

Magnetic flux density B quantifies the magnetic field through the force it exerts on moving charged particles and currents, with the key insight that this force acts perpendicular to both the particle's motion and the field direction.

📌 Key points (3–5)

  • What B represents: a vector field describing the magnetic field, measured in tesla (T) or Wb/m², defined through the force equation on charged particles.
  • When magnetic force appears: a magnetic field exerts force only on charged particles in motion—stationary charges experience no magnetic force.
  • Direction is counterintuitive: the force is perpendicular to both the velocity and the magnetic field direction (via the cross product), not aligned with B.
  • Common confusion: "flux density" terminology comes from engineering practice (integrating B over surfaces), not from B literally being a density of something physical at every point.
  • Closed-loop property: magnetic field lines always form closed loops, unlike electric field lines.

🧲 What is the magnetic field?

🧲 Sources of magnetic fields

Magnetic fields arise from two main sources:

  • Permanent magnets: the field originates from atomic-scale mechanisms in the material (details deferred in the excerpt).
  • Electric currents: a current-bearing coil generates a magnetic field identical in nature to that of a permanent magnet.

The excerpt emphasizes that both sources produce the same underlying phenomenon—the same vector field.

🔄 Reciprocal interaction

The magnetic field describes the force exerted on permanent magnets and currents in the presence of other permanent magnets and currents.

  • Permanent magnets affect each other (N attracts S, N repels N).
  • Currents influence permanent magnets, and vice versa.
  • This symmetry shows that current is both a source of magnetic fields and subject to magnetic forces.

⚡ How magnetic fields exert force

⚡ Force on a moving charged particle

The force applied to a particle bearing charge q is:

F = q v × B

where v is the velocity and × denotes the cross product.

  • Key requirement: the particle must be in motion. A stationary charged particle in a magnetic field experiences no force.
  • Example: Figure 2.6 shows a charged particle motionless (top)—no force; then moving with velocity v perpendicular to the page (bottom)—suddenly a force F appears.

🔀 Why perpendicular?

  • The cross product means the force is perpendicular to both the direction of motion v and the magnetic field B.
  • The excerpt acknowledges this is counterintuitive: "The reader would be well-justified in wondering why the force exerted by the magnetic field should be perpendicular to B."
  • Classical physics provides no obvious answer; a full explanation requires quantum mechanics and the concept of the electromagnetic force.
  • For engineering purposes: accept this as experimental fact and proceed accordingly.

🔌 Connection to current

  • A single charged particle in motion is the simplest form of current.
  • Since motion is required for the magnetic field to influence the particle, the magnetic field exerts force on current in general.
  • This explains why current-bearing coils and permanent magnets interact.

📏 Defining and measuring B

📏 Units and terminology

  • SI unit: tesla (T), which equals (N·s)/(C·m).
  • Alternative unit: Wb/m² (weber per square meter), where 1 Wb/m² = 1 T.
  • The weber (Wb) is the SI unit for magnetic flux (the integral of B over a surface).

Magnetic flux density (B, T or Wb/m²) is a description of the magnetic field that can be defined as the solution to Equation 2.9.

🤔 Why "flux density"?

  • The terminology is "somewhat arbitrary" and "not even uniformly accepted."
  • In engineering electromagnetics, B is called a flux density because:
    • Engineers frequently integrate B over mathematical surfaces.
    • Any quantity obtained by integration over a surface is called "flux."
    • Therefore B is naturally thought of as flux per unit area.
  • Don't confuse: "flux density" does not mean B is a density of some physical substance at each point; it is a field quantity whose integral over a surface yields flux.

🔁 Magnetic field lines

🔁 Definition and visualization

A magnetic field line is the curve in space traced out by following the direction in which the magnetic field vector points.

  • Figures 2.7 and 2.8 illustrate field lines for a bar magnet and a current-bearing coil, respectively.
  • Field lines provide a visual representation of the direction and structure of the magnetic field.

🔁 Closed-loop property

A magnetic field line always forms a closed loop.

  • This is a fundamental property distinguishing magnetic fields from electric fields (electric field lines can start and end on charges).
  • The closed-loop nature reflects the absence of magnetic monopoles (no isolated "north" or "south" charges).
13

Permeability

2.6 Permeability

🧭 Overview

🧠 One-sentence thesis

Permeability (μ) is the material property that determines how strongly a material amplifies magnetic flux density, with most materials having values close to free space except for a small class of magnetic materials that can amplify the field by factors up to a million.

📌 Key points (3–5)

  • What permeability measures: the effect of material in determining magnetic flux density—all else equal, B increases in proportion to μ.
  • Free space baseline: permeability in free space (μ₀) equals 4π × 10⁻⁷ H/m and represents the minimum possible value.
  • Relative permeability: μᵣ = μ / μ₀ compares a material's permeability to free space; most materials have μᵣ ≈ 1.
  • Magnetic materials exception: a small class of materials (e.g., ferromagnetic materials like iron) can have μᵣ as large as ~10⁶.
  • Common confusion: permeability is not about the field itself but about how material amplifies or modifies the magnetic flux density produced by a given source.

🧲 What permeability is and how it works

🧲 Definition and role

Permeability (μ, H/m): describes the effect of material in determining the magnetic flux density.

  • Permeability is the constant of proportionality that captures how material affects the magnetic field.
  • It appears in the equation for magnetic flux density from a moving charged particle:
    • B(r) = (μ q v) / (4π R²) × R̂
    • Where q is charge, v is velocity, R̂ is the unit vector from particle to field point, R is distance, and × is the cross product.
  • Why it matters: B increases with charge and speed (the source), decreases with distance (inverse square law), and scales directly with μ (the material effect).

📏 Units and dimensional analysis

  • Permeability has units of henries per meter (H/m).
  • This comes from the fact that B is in Wb/m², v is in m/s, and 1 H = 1 Wb/A.
  • The unit captures how material modifies the relationship between moving charge and resulting magnetic flux density.

🌌 Free space and relative permeability

🌌 Free space permeability (μ₀)

  • In free space (vacuum), permeability equals μ₀:
    • μ₀ = 4π × 10⁻⁷ H/m
  • This is the minimum possible value of permeability.
  • It represents the baseline against which all materials are compared.

📊 Relative permeability (μᵣ)

Relative permeability: μᵣ = μ / μ₀

  • Gives permeability relative to the free space value.
  • For most materials: μᵣ ≈ 1 (approximately equal to free space).
  • This means most materials do not significantly amplify or reduce magnetic flux density compared to vacuum.

Example: If a material has μᵣ = 1.0002, its permeability is only 0.02% higher than free space—essentially no difference for most engineering purposes.

⚠️ Don't confuse

  • Relative permeability is dimensionless (a ratio), while permeability μ has units of H/m.
  • μᵣ ≈ 1 does not mean "no magnetic field"; it means "the material behaves like free space."

🧲 Magnetic materials: the exception

🧲 What makes a material "magnetic"

Magnetic materials: a small class of materials with μᵣ significantly different from 1, potentially as large as ~10⁶.

  • Most materials have μᵣ ≈ 1, but magnetic materials are the exception.
  • These materials can amplify magnetic flux density by factors of up to a million compared to free space.
  • Ferromagnetic materials are a commonly-encountered category; the best-known example is iron.

📈 Practical implications

Material typeRelative permeability (μᵣ)Effect on B
Most materials≈ 1Essentially same as free space
Magnetic materialsUp to ~10⁶Can amplify B by up to a million times

Example: If a moving charge produces a certain B in free space, placing iron (a ferromagnetic material) in the field could increase B by a factor of thousands or more, depending on the specific iron alloy.

⚠️ Don't confuse

  • "Magnetic material" does not mean "any material that experiences magnetic force."
  • It specifically refers to materials with μᵣ significantly greater than 1—those that strongly amplify magnetic flux density.
  • The vast majority of materials (including most metals, plastics, air, water) are not magnetic materials in this technical sense.
14

Magnetic Field Intensity

2.7 Magnetic Field Intensity

🧭 Overview

🧠 One-sentence thesis

Magnetic field intensity H factors out material effects from the magnetic flux density B, making it essential for analyzing boundaries between different materials and nonlinear magnetic behavior.

📌 Key points (3–5)

  • What H represents: an alternative description of the magnetic field where material permeability is separated out from B.
  • How B and H relate: B equals permeability μ times H; in homogeneous media, H does not depend on μ.
  • Units and interpretation: H has units of amperes per meter (A/m) but represents an equivalent (not actual) current description of the magnetic field, not a surface current density.
  • Common confusion: H may seem redundant when you know B and μ in homogeneous media, but it becomes crucial at boundaries between materials with different permeabilities.
  • Why it matters: boundary conditions constrain the tangent component of H (whereas B constrains only the perpendicular component), and H is useful when permeability is nonlinear.

🔗 Relationship between B and H

🔗 Factoring out material effects

The excerpt shows that magnetic flux density B can be rewritten to separate material permeability:

  • The Biot-Savart Law gives B for a point charge q moving at velocity v: B equals μ times (q times v) divided by (4 π R squared), cross product with the unit vector R-hat.
  • This can be rewritten as B equals μ times H, where H equals (q times v) divided by (4 π R squared), cross product with R-hat.
  • Key insight: in homogeneous media, H does not depend on μ—the material property is isolated in the μ factor.

Magnetic field intensity H (A/m), defined using B equals μ H, is a description of the magnetic field independent from material properties.

📏 Units and what they mean

  • Dimensional analysis shows H has units of amperes per meter (A/m).
  • Don't confuse: although A/m can describe surface current density (current distributed over a linear cross-section), H does not represent an actual current distribution.
  • Instead, H is "associated with the magnetic field" and can be viewed as describing the field "in terms of an equivalent (but not actual) current."

Example: A current distribution might physically have units A/m, but H uses the same units to describe the magnetic field itself, not the source current.

🧱 Why H is not redundant

🧱 Homogeneous media vs boundaries

  • In homogeneous media (uniform permeability), H might appear redundant: if you know B and μ, you can calculate H.
  • However, the excerpt emphasizes that H "becomes important—and decidedly not redundant—when we encounter boundaries between media having different permeabilities."

🔀 Boundary conditions

The excerpt explains how H and B constrain different components at boundaries:

FieldWhat is constrained at a boundary
HTangent component (parallel to the boundary)
BPerpendicular component (normal to the boundary)
  • If you only consider B, you constrain only the perpendicular component.
  • Boundary conditions on H constrain the tangent component, which is essential for solving problems at material interfaces.
  • Don't confuse: ignoring H means missing the tangent-component constraint, which matters when permeabilities differ across a boundary.

Example: At the boundary between two materials with different μ, the tangent component of H must satisfy a specific condition; using only B would not capture this.

🔄 Nonlinear magnetic behavior

🔄 When permeability is not constant

  • The excerpt notes that H is also useful "in certain problems in which μ is not a constant, but rather is a function of magnetic field strength."
  • In such cases, the magnetic behavior is called nonlinear.
  • H provides a clearer framework when permeability varies with field strength, because it separates the field description from the material response.

Example: In a material where μ changes as the field gets stronger, expressing the field as H (independent of μ) simplifies analysis compared to working directly with B (which mixes field and material).

🧲 Context: Magnetic materials

🧲 Permeability values

The excerpt provides context from the preceding section:

  • For most materials, relative permeability μ_r is approximately 1.
  • A small class of magnetic materials can have μ_r as large as approximately 10 to the power of 6.
  • Ferromagnetic materials (best-known example: iron) are a commonly-encountered category of magnetic materials.

Why this matters for H: The large variation in permeability across materials makes the separation of H from μ especially valuable at boundaries and in nonlinear cases.

15

Electromagnetic Properties of Materials

2.8 Electromagnetic Properties of Materials

🧭 Overview

🧠 One-sentence thesis

Electromagnetic analysis depends on three constitutive parameters—permittivity, permeability, and conductivity—and most practical materials can be treated as homogeneous, isotropic, linear, and time-invariant to simplify engineering calculations.

📌 Key points (3–5)

  • Three principal constitutive parameters: permittivity (ε), permeability (μ), and conductivity (σ) quantify how materials respond to electric charge, current, and electric fields respectively.
  • Four idealizing properties: homogeneity, isotropy, linearity, and time-invariance simplify electromagnetic theory; most materials approximate these properties well enough for practical use.
  • Linearity and time-invariance enable superposition: when both properties hold, fields from multiple sources can be added together; nonlinear or time-varying materials break this rule.
  • Common confusion: no real material is perfectly ideal, but deviations are usually small enough not to affect engineering analysis; some materials deviate significantly in one property but can still be analyzed with modifications.
  • Nonlinearity complicates analysis: materials whose parameters depend on field strength (e.g., ferromagnetic materials) do not follow elementary circuit theory and require more difficult analysis.

📐 The Three Constitutive Parameters

⚡ Permittivity (ε)

Permittivity (ε, F/m): quantifies the effect of matter in determining the electric field in response to electric charge.

  • Units: farads per meter (F/m).
  • Describes how a material influences the electric field when charge is present.
  • The excerpt notes this is addressed in Section 2.3 (not included here).

🧲 Permeability (μ)

Permeability (μ, H/m): quantifies the effect of matter in determining the magnetic field in response to current.

  • Units: henries per meter (H/m).
  • Describes how a material influences the magnetic field when current flows.
  • The excerpt notes this is addressed in Section 2.6 (not included here).
  • Important note: permeability can be nonlinear in ferromagnetic materials, meaning its value depends on the magnetic field strength itself.

🔌 Conductivity (σ)

Conductivity (σ, S/m): quantifies the effect of matter in determining the flow of current in response to an electric field.

  • Units: siemens per meter (S/m).
  • Describes how easily current flows through a material when an electric field is applied.
  • The excerpt notes this is addressed in Section 6.3 (not included here).

🧱 Four Idealizing Properties of Materials

🟦 Homogeneity

A material that is homogeneous is uniform over the space it occupies; that is, the values of its constitutive parameters are constant at all locations within the material.

  • Homogeneous means the material's properties do not vary from place to place.
  • The constitutive parameters (ε, μ, σ) have the same values everywhere within the material.
  • Counter-example: soil is not homogeneous because it contains multiple chemically-distinct compounds that are not thoroughly mixed.
  • Don't confuse: homogeneity is about spatial uniformity, not about how the material responds to fields.

🔄 Isotropy

A material that is isotropic behaves in precisely the same way regardless of how it is oriented with respect to sources and fields occupying the same space.

  • Isotropic means the material's response does not depend on direction or orientation.
  • Rotating an isotropic material does not change its electromagnetic properties.
  • Counter-example: quartz has atoms arranged in a uniformly-spaced crystalline lattice, so its electromagnetic properties change when you rotate it relative to applied sources and fields.
  • Don't confuse: isotropy is about directional uniformity, not spatial uniformity (that's homogeneity).

📏 Linearity

A material is said to be linear if its properties are constant and independent of the magnitude of the sources and fields applied to the material.

  • Linear means the constitutive parameters do not change as field or source strength changes.
  • The material's response scales proportionally with the applied field or source.
  • Example: capacitor dielectric material is approximately linear below the rated working voltage—permittivity ε is constant, so capacitance does not vary significantly with voltage V. Above the working voltage, ε depends on V, making capacitance a function of V (nonlinear behavior).
  • Example: ferromagnetic materials have nonlinear permeability μ, meaning the precise value of μ depends on the magnitude of the magnetic field.
  • Don't confuse: linearity is about whether parameters change with field strength, not whether they change with position (homogeneity) or direction (isotropy).

⏱️ Time-invariance

  • A time-invariant material has electromagnetic properties that do not change over time.
  • Counter-example: piezoelectric materials have electromagnetic properties that vary significantly depending on the mechanical forces applied to them; this property is exploited to make sensors and transducers.
  • Don't confuse: time-invariance means properties don't change with time or external conditions, not that the material doesn't respond to fields.

🔗 Why Linearity and Time-Invariance Matter

➕ Superposition principle

  • Linearity and time-invariance (LTI) together are requirements for superposition.
  • Superposition means you can calculate the effect of each source separately and then add the results together.

How superposition works in LTI materials:

  • Calculate field E₁ due to point charge q₁ at location r₁.
  • Calculate field E₂ due to point charge q₂ at location r₂.
  • When both charges are present simultaneously, the total field is E₁ + E₂.

What happens without LTI:

  • The same superposition rule does not necessarily hold for materials that are not LTI.
  • You cannot simply add fields calculated separately.

⚠️ Consequences of nonlinearity

  • Devices that are nonlinear (and therefore not LTI) do not necessarily follow the rules of elementary circuit theory.
  • Elementary circuit theory presumes that superposition applies.
  • Nonlinearity makes analysis and design much more difficult.

🌍 Real-World Materials vs Ideal Assumptions

🎯 Practical approximations

  • No practical material is truly homogeneous, isotropic, linear, and time-invariant.
  • However, for most materials in most applications, the deviation from this ideal condition is not large enough to significantly affect engineering analysis and design.
  • The excerpt emphasizes that these idealizations are used "to keep electromagnetic theory from becoming too complex."

🔧 When deviations matter

  • In some cases, materials may be significantly non-ideal in one of these respects.
  • Such materials may still be analyzed with appropriate modifications to the theory.
  • Example: ferromagnetic materials with nonlinear permeability require special treatment but can still be analyzed.
  • Don't confuse: "non-ideal" does not mean "unusable"—it means the standard simplified theory needs adjustments.
PropertyWhat it meansCounter-example from excerpt
HomogeneityUniform properties throughout spaceSoil (multiple unmixed compounds)
IsotropySame behavior in all directionsQuartz (crystalline lattice structure)
LinearityParameters independent of field strengthCapacitor dielectric above working voltage; ferromagnetic materials
Time-invarianceProperties do not change over timePiezoelectric materials (vary with mechanical force)
16

Introduction to Transmission Lines

3.1 Introduction to Transmission Lines

🧭 Overview

🧠 One-sentence thesis

Transmission lines are designed to support guided electromagnetic waves with controlled impedance, low loss, and immunity from electromagnetic interference, becoming essential when operating frequencies or physical dimensions approach a significant fraction of a wavelength.

📌 Key points (3–5)

  • When random wires fail: at very low frequencies and short lengths, simple wires work; but when length or frequency increases so that dimensions become a significant fraction of a wavelength, proper transmission lines become necessary.
  • What makes a guided wave: an electromagnetic wave contained within or bound to the line that does not radiate away, normally achieved when dimensions are small relative to wavelength (e.g., λ/100 or 1% of wavelength).
  • The reflection problem: beyond very small phase shifts, reflected signals traveling in the opposite direction create impedance variations along the line, making the line no longer "transparent" to source and destination.
  • Common confusion: the same wire arrangement can work fine at 10 MHz over 3 cm (0.1% of wavelength) but fail at 1 GHz or over 3 m (10% of wavelength)—it's the ratio of physical size to wavelength that matters.
  • Design goals: transmission lines control impedance, minimize loss, and provide electromagnetic interference immunity through specific cross-sectional geometry and materials.

📏 When transmission lines become necessary

📏 The wavelength fraction rule

A guided wave is an electromagnetic wave that is contained within or bound to the line, and which does not radiate away from the line.

  • This condition is normally met if the length and cross-sectional dimensions are small relative to a wavelength—say λ/100 (1% of the wavelength).
  • Below this threshold, randomly-arranged wires may serve well enough.
  • Above this threshold, unintended radiation becomes a concern and proper transmission line design is needed.

🔢 Concrete example: 10 MHz vs 1 GHz

The excerpt provides a numerical comparison:

ScenarioFrequencyLengthWavelengthRatioSuitable?
Low frequency, short10 MHz3 cm30 m0.1%Yes, random wires OK
Higher frequency1 GHz3 cm0.3 m10%No, need proper line
Longer length10 MHz3 m30 m10%No, need proper line
  • When length is only 0.1% of wavelength, the phase shift is roughly 0.001 × 360° = 0.36°, negligible.
  • When length reaches 10% of wavelength, the round-trip phase shift becomes 72°, no longer negligible.
  • Don't confuse: it's not just frequency or just length—it's the ratio of physical dimension to wavelength that determines whether a proper transmission line is needed.

🌊 The transparency problem

🌊 What "transparent" means

  • At very low frequencies and short lengths, the entire transmission line is at approximately the same electrical potential.
  • The line appears "transparent" to both the source and the destination—it doesn't modify the signal significantly.
  • Loss, reactance, and electromagnetic interference are not a concern in this regime.

🔄 Reflection and impedance variation

When dimensions become a significant fraction of wavelength:

  • A reflected signal traveling in the opposite direction adds to the forward signal.
  • The total electrical potential now varies in both magnitude and phase with position along the line.
  • The impedance looking toward the destination via the transmission line becomes different from the impedance looking toward the destination directly.
  • This modified impedance depends on cross-sectional geometry, materials, and length of the line.

Example: With a round-trip phase shift of 72° (at 1 GHz over 3 m, or 10 MHz over 3 m), reflected waves interfere constructively or destructively at different positions, creating standing wave patterns and impedance mismatches.

Don't confuse: The problem is not just that signals take time to travel; it's that reflections create position-dependent interference, making the line behave as an active circuit element rather than a passive connection.

🎯 Design objectives

🎯 Three key goals

Transmission lines are designed to achieve:

  1. Controlled impedance: predictable impedance characteristics determined by geometry and materials, not random variations.
  2. Low loss: minimize signal attenuation over the length of the line.
  3. EMI immunity: degree of protection from electromagnetic interference.

🔧 How geometry and materials matter

  • Cross-sectional geometry determines impedance characteristics and EMI immunity.
  • Materials affect both loss and impedance.
  • Length influences the total phase shift and reflection behavior.

These factors are controlled through specific transmission line designs rather than left to chance with random wire arrangements.

🛠️ Common transmission line types

🛠️ TEM transmission lines

TEM (transverse electromagnetic) transmission lines employ a single electromagnetic wave "mode" having electric and magnetic field vectors in directions perpendicular to the axis of the line.

Two common examples mentioned:

TypeDescriptionApplication
Coaxial line(Figure 3.2 in excerpt)Radio frequency applications
Microstrip line(Figure 3.3 in excerpt)Radio frequency applications

📐 TEM field structure

  • Electric and magnetic field vectors are perpendicular to the direction of propagation.
  • The wave propagates along the length of the transmission line.
  • This is the primary structure for radio frequency applications.
  • Note: The excerpt mentions that not all transmission lines exhibit TEM field structure, but details on non-TEM lines are not provided in this section.
17

Types of Transmission Lines

3.2 Types of Transmission Lines

🧭 Overview

🧠 One-sentence thesis

Transmission lines are designed to guide electromagnetic waves with controlled impedance, low loss, and EMI immunity, and they differ in field structure—TEM lines (coaxial, microstrip) have simpler perpendicular fields, while non-TEM lines (waveguides, multimode optical fiber) exhibit more complex field patterns.

📌 Key points (3–5)

  • Purpose of transmission lines: support guided waves with controlled impedance, low loss, and immunity from electromagnetic interference (EMI).
  • TEM vs non-TEM distinction: TEM lines have electric and magnetic fields perpendicular to the propagation axis; non-TEM lines have fields that are not necessarily perpendicular and are more complex.
  • Common TEM examples: coaxial line and microstrip line, used primarily in radio frequency applications.
  • Common non-TEM examples: waveguides (for very low loss or high power) and multimode optical fiber (complex fields due to small wavelength relative to fiber cross-section).
  • Common confusion: not all transmission lines have the same field structure—TEM is simpler and more common at RF, while non-TEM appears when loss/power requirements are extreme or wavelength is very small.

📡 What transmission lines do

📡 Core function

Transmission lines are designed to support guided waves with controlled impedance, low loss, and a degree of immunity from EMI.

  • They guide electromagnetic waves from source to destination.
  • Controlled impedance: the cross-sectional geometry, materials, and length determine the impedance seen by the source.
  • Low loss: materials and geometry minimize energy dissipation.
  • EMI immunity: design reduces susceptibility to external electromagnetic interference.

⚡ Why impedance matters

  • When the transmission line is electrically long (round-trip phase shift becomes significant, e.g., 72°), reflected signals combine with forward signals.
  • The total electrical potential varies in magnitude and phase along the line.
  • The impedance looking toward the destination via the line differs from the direct impedance.
  • Example: at 10 MHz and 3 cm length, round-trip phase shift is only ~0.72°, so the line is transparent; at 1 GHz or 3 m length, round-trip phase shift is 72°, and impedance modification becomes important.

🔌 TEM transmission lines

🔌 What TEM means

TEM (transverse electromagnetic) transmission line: employs a single electromagnetic wave mode with electric and magnetic field vectors in directions perpendicular to the axis of the line.

  • "Transverse" means the fields are perpendicular to the direction of propagation.
  • The wave propagates along the length of the line; fields point sideways.
  • TEM lines appear primarily in radio frequency applications.

🧲 Field structure in TEM lines

  • Both electric and magnetic field vectors are perpendicular to the propagation axis.
  • The field structure is simpler compared to non-TEM lines.
  • Example: in coaxial and microstrip lines, the wave propagates along the line axis, and fields are oriented in the cross-sectional plane.

🔩 Coaxial line

  • Structure: inner conductor surrounded by outer conductor (see Figure 3.2 in excerpt).
  • Fields are confined between the conductors.
  • Common TEM transmission line type.

🛤️ Microstrip line

  • Structure: metallic trace on top of a dielectric slab over a ground plane (see Figure 3.3 in excerpt).
  • Fields exist both inside and outside the line; outside fields are possibly significant and complicated (not fully shown in diagrams).
  • Common TEM transmission line type.

🌊 Non-TEM transmission lines

🌊 What non-TEM means

  • Electric and magnetic field vectors are not necessarily perpendicular to the axis of the line.
  • Field structure is complex relative to TEM lines.
  • These lines are designed to guide waves with relatively complex structure.

📻 Waveguides

  • Example of non-TEM transmission line.
  • Most prevalent at radio frequencies.
  • Used in applications requiring very low loss or handling very high power levels.
  • Example: radio frequency waveguides in air traffic control radar (see Figure 3.6 in excerpt).

💡 Multimode optical fiber

  • Another example of non-TEM transmission line.
  • Exhibits complex field structure because the wavelength of light is very small compared to the cross-section of the fiber.
  • This makes excitation and propagation of non-TEM waves difficult to avoid.
  • Don't confuse with: single mode fiber, which overcomes this issue but is much more difficult and expensive to manufacture.

🔍 Comparison of transmission line types

TypeField structureTypical applicationsKey characteristics
Coaxial lineTEM (fields perpendicular to axis)Radio frequencySimple field structure, confined fields
Microstrip lineTEM (fields perpendicular to axis)Radio frequencySimple field structure, some external fields
WaveguideNon-TEM (fields not necessarily perpendicular)RF, very low loss or high powerComplex field structure
Multimode optical fiberNon-TEM (fields not necessarily perpendicular)Optical transmissionComplex fields due to small wavelength vs cross-section

🧩 When to use which

  • TEM lines (coaxial, microstrip): standard radio frequency applications where simplicity and controlled impedance are priorities.
  • Waveguides: when very low loss or very high power handling is critical at radio frequencies.
  • Multimode optical fiber: optical transmission where wavelength is very small relative to fiber size (single mode fiber is preferred when feasible but is more expensive).
18

Transmission Lines as Two-Port Devices

3.3 Transmission Lines as Two-Port Devices

🧭 Overview

🧠 One-sentence thesis

A transmission line can be modeled as a two-port device whose behavior is fully described by its length and three key parameters—phase propagation constant, attenuation constant, and characteristic impedance—and it is typically not transparent to the source and load, meaning the impedances seen at each end may differ from the actual source and load impedances.

📌 Key points (3–5)

  • Two-port representation: transmission lines connect a source (generator) to a load and are described by length l and parameters β (phase propagation), α (attenuation), and Z₀ (characteristic impedance).
  • Not transparent: the load impedance Z_L may not equal the impedance the source sees, and vice versa; the transmission line transforms impedances depending on its parameters.
  • Lumped-element model: a transmission line can be analyzed by dividing it into small segments, each modeled as a circuit with resistance R′Δz, conductance G′Δz, capacitance C′Δz, and inductance L′Δz.
  • Common confusion: G′ (conductance per unit length) is not equal to 1/R′; they describe entirely different physical mechanisms—R′ is series resistance of conductors, G′ is leakage between conductors.
  • Parameter units and meaning: β (rad/m) relates to wavelength λ = 2π/β; α (1/m) quantifies loss; Z₀ (Ω) is the voltage-to-current ratio under perfect matching.

🔌 Two-port circuit representation

🔌 Standard circuit symbols

The excerpt shows three common ways to represent transmission lines in circuit diagrams:

  • As a generic two-conductor direct connection (top symbol).
  • As a generic two-port "black box" (middle symbol).
  • As a coaxial cable (bottom symbol).

In each case:

  • The source (also called the generator) is represented by a Thévenin equivalent circuit: a voltage source V_S in series with an impedance Z_S.
  • The termination (also called the load) is represented as an impedance Z_L, which can be any circuit exhibiting that input impedance.

📐 Complete description parameters

A two-port representation of a transmission line is completely described by its length l along with some combination of the following parameters:

ParameterSymbolUnitsWhat it represents
Phase propagation constantβrad/mRelates to wavelength in the line: λ = 2π/β
Attenuation constantα1/mQuantifies the effect of loss in the line
Characteristic impedanceZ₀ΩRatio of voltage to current when the line is perfectly impedance-matched at both ends
  • These parameters depend on the materials and geometry of the line.
  • Example: a longer line with higher α will exhibit more loss; a line with specific Z₀ will only present that impedance ratio under perfect matching conditions.

🔄 Impedance transformation effect

🔄 Non-transparency of transmission lines

The excerpt emphasizes that a transmission line is typically not transparent to the source and load:

  • The load impedance may be Z_L, but the impedance presented to the source may or may not equal Z_L.
  • Similarly, the source impedance may be Z_S, but the impedance presented to the load may or may not equal Z_S.

Why this matters:

  • The transmission line transforms impedances based on its parameters (β, α, Z₀) and length l.
  • You cannot assume the source "sees" the load directly, or vice versa.
  • Example: even if the load is 50 Ω, the source might see a different impedance depending on the line's characteristics and length.

🔍 What determines the transformation

The effect of the transmission line on source and load impedances depends on:

  • The phase propagation constant β.
  • The attenuation constant α.
  • The characteristic impedance Z₀.
  • The length l of the line.

Don't confuse: the load impedance Z_L is a property of the termination circuit, but the impedance the source experiences is modified by the transmission line itself.

🧩 Lumped-element model

🧩 Concept and structure

It is possible to ascertain the relevant behaviors of a transmission line using elementary circuit theory applied to a differential-length lumped-element model of the transmission line.

The approach:

  • Divide the transmission line into segments of small but finite length Δz.
  • Each segment is modeled as an identical two-port with an equivalent circuit consisting of four components.
  • This allows analysis using standard circuit theory.

⚡ Four circuit components

Each segment of length Δz contains:

  1. Resistance R′Δz

    • Represents the series-combined ohmic resistance of both conductors.
    • R′ is resistance per unit length (Ω/m); multiplying by Δz gives resistance in Ω.
    • Accounts for resistive loss as current flows through the conductors.
  2. Conductance G′Δz

    • Represents leakage of current directly from one conductor to the other.
    • When G′Δz > 0, the resistance between conductors is less than infinite, allowing current to flow between them.
    • G′ has units of S/m (siemens per meter).
    • This is a loss mechanism separate from R′.
  3. Capacitance C′Δz

    • Represents the capacitance of the transmission line structure.
    • Capacitance is the tendency to store energy in electric fields.
    • Depends on cross-sectional geometry and the media separating the conductors.
    • C′ has units of F/m (farads per meter).
  4. Inductance L′Δz

    • Represents the inductance of the transmission line structure.
    • Inductance is the tendency to store energy in magnetic fields.
    • Like capacitance, depends on geometry and materials.
    • (The excerpt cuts off before completing the description of L′.)

⚠️ Common confusion: G′ vs R′

Important distinction:

  • G′ is not equal to 1/R′.
  • They describe entirely different physical mechanisms:
    • R′: series resistance of the conductors themselves (current flowing through the conductors).
    • G′: leakage conductance between the conductors (current flowing across from one conductor to the other).
  • In principle, either could be defined as resistance or conductance; the choice reflects the physical mechanism being modeled.

Example: a transmission line could have high R′ (lossy conductors) but low G′ (good insulation between conductors), or vice versa.

19

Lumped-Element Model

3.4 Lumped-Element Model

🧭 Overview

🧠 One-sentence thesis

A transmission line can be analyzed using elementary circuit theory by dividing it into small segments, each modeled as a two-port circuit with four lumped elements (resistance, conductance, capacitance, and inductance) that together capture the line's physical behavior.

📌 Key points (3–5)

  • Core modeling approach: divide the transmission line into small segments of length Δz, each represented by an identical lumped-element equivalent circuit.
  • Four fundamental parameters: R′ (series resistance per unit length), G′ (leakage conductance per unit length), C′ (capacitance per unit length), and L′ (inductance per unit length).
  • Common confusion: G′ is not equal to 1/R′—they describe entirely different physical mechanisms (leakage between conductors vs. ohmic resistance along conductors).
  • Per-unit-length notation: the prime (′) notation indicates parameters are specified per meter; multiply by length Δz to get actual component values in each segment.
  • Foundation for analysis: this model enables deriving the telegrapher's equations that govern voltage and current behavior along the line.

🔧 The segmentation concept

🔧 How the model works

  • The transmission line is aligned along the z-axis and divided into segments of small but finite length Δz.
  • Each segment is modeled as an identical two-port network with the same equivalent circuit.
  • The segments are cascaded (series-connected) to represent the entire transmission line.
  • This approach allows using elementary circuit theory (Kirchhoff's laws) to analyze transmission line behavior.

📐 Why small segments matter

  • The model treats a distributed structure (the transmission line) as a cascade of discrete lumped elements.
  • As Δz becomes smaller, the model more accurately represents the continuous nature of the actual line.
  • The telegrapher's equations are derived by taking the limit as Δz → 0.

⚡ The four lumped elements

🔴 Series resistance R′Δz

The resistance R′Δz represents the series-combined ohmic resistance of the two conductors.

  • What it captures: resistive losses as current flows through both conductors.
  • Units: R′ has units of Ω/m (ohms per meter); multiply by Δz to get resistance in Ω for one segment.
  • Important: this accounts for both conductors, since current must flow through both in the actual transmission line.
  • Physical meaning: energy dissipated as heat due to finite conductivity of the conductor material.

🟢 Shunt conductance G′Δz

The conductance G′Δz represents the leakage of current directly from one conductor to the other.

  • What it captures: current that flows between the conductors (not along them).
  • Units: G′ has units of S/m (siemens per meter).
  • When G′Δz > 0: the resistance between conductors is less than infinite, allowing leakage current.
  • Physical meaning: power loss separate from the ohmic loss R′; represents imperfect insulation between conductors.
  • Don't confuse: G′ is not equal to 1/R′—they describe entirely different physical mechanisms. R′ is resistance along the conductors; G′ is conductance between the conductors. Either could in principle be defined as resistance or conductance.

🔵 Shunt capacitance C′Δz

The capacitance C′Δz represents the capacitance of the transmission line structure.

  • What it captures: the tendency to store energy in electric fields between the conductors.
  • Units: C′ has units of F/m (farads per meter).
  • What it depends on: cross-sectional geometry and the media (dielectric material) separating the conductors.
  • Physical meaning: the ability of the structure to hold charge at a given voltage difference.

🟣 Series inductance L′Δz

The inductance L′Δz represents the inductance of the transmission line structure.

  • What it captures: the tendency to store energy in magnetic fields around the conductors.
  • Units: L′ has units of H/m (henries per meter).
  • What it depends on: cross-sectional geometry and the media separating the conductors (like capacitance).
  • Physical meaning: the opposition to changes in current due to magnetic field energy storage.

📊 Parameter summary and usage

📊 The four parameters compared

ParameterSymbolUnitsPhysical mechanismLoss or storage?
ResistanceR′Ω/mOhmic resistance along both conductorsLoss (heat)
ConductanceG′S/mLeakage current between conductorsLoss (leakage)
CapacitanceC′F/mElectric field energy storageStorage
InductanceL′H/mMagnetic field energy storageStorage

🔢 How to use the model

  • Obtaining values: methods for computing R′, G′, C′, and L′ are addressed elsewhere in the book (not in this excerpt).
  • Building the circuit: for each segment of length Δz, use component values R′Δz, G′Δz, C′Δz, and L′Δz.
  • Analysis: apply Kirchhoff's voltage and current laws to the lumped-element circuit.
  • Deriving equations: the telegrapher's equations (Section 3.5) are derived by applying circuit laws and taking the limit as Δz → 0.

🔗 Connection to telegrapher's equations

🔗 From lumped model to governing equations

  • The lumped-element model serves as the foundation for deriving the telegrapher's equations.
  • These equations govern the potential v(z,t) and current i(z,t) along the transmission line oriented along the z-axis.
  • The derivation uses Kirchhoff's voltage law applied to one segment, then takes the limit as Δz → 0.

📝 Example derivation step (from the excerpt)

  • Assign v(z,t) and i(z,t) to the left side of a segment, and v(z+Δz,t) and i(z+Δz,t) to the right side.
  • Apply Kirchhoff's voltage law through R′Δz and L′Δz:
    • v(z,t) − (R′Δz)i(z,t) − (L′Δz)(∂/∂t)i(z,t) − v(z+Δz,t) = 0
  • Rearrange and divide by Δz, then take the limit as Δz → 0:
    • −(∂/∂z)v(z,t) = R′i(z,t) + L′(∂/∂t)i(z,t)
  • This is one of the telegrapher's equations (the voltage equation).
  • A similar process using Kirchhoff's current law yields the current equation (not fully shown in this excerpt).
20

Telegrapher's Equations

3.5 Telegrapher’s Equations

🧭 Overview

🧠 One-sentence thesis

The telegrapher's equations are coupled differential equations that govern voltage and current along a transmission line and can be simplified using phasor representation to eliminate time derivatives, leaving only spatial variation.

📌 Key points (3–5)

  • What they are: two coupled differential equations (Equations 3.3 and 3.6) that describe how voltage v(z,t) and current i(z,t) vary along a transmission line, derived from Kirchhoff's laws applied to the lumped-element model.
  • How they're derived: apply Kirchhoff's voltage law and current law to a differential segment Δz of the transmission line, then take the limit as Δz approaches zero.
  • Phasor simplification: converting to phasor representation (Equations 3.9 and 3.10) eliminates time derivatives, leaving only spatial derivatives—considerably simpler to solve.
  • Common confusion: time-domain telegrapher's equations are "usually more than we need or want"; for sinusoidal stimuli, phasor form is preferred because it removes time complexity.
  • Four parameters: the equations depend on R′, G′, L′, and C′ (resistance, conductance, capacitance, and inductance per unit length) from the lumped-element model.

📐 Deriving the time-domain equations

📐 Setup and sign conventions

  • The transmission line is oriented along the z axis and divided into segments of small length Δz.
  • Each segment uses the lumped-element model from Section 3.4 with four components: R′Δz, L′Δz, G′Δz, and C′Δz.
  • Voltage and current are defined at the left side as v(z,t) and i(z,t), and at the right side as v(z+Δz,t) and i(z+Δz,t), with reference polarity and direction as shown in Figure 3.11.

⚡ Kirchhoff's voltage law (KVL)

  • Apply KVL from the left port through R′Δz and L′Δz, returning via the right port:
    • v(z,t) − (R′Δz)i(z,t) − (L′Δz)(∂i(z,t)/∂t) − v(z+Δz,t) = 0
  • Rearrange to isolate current terms on the right, divide by Δz, then take the limit as Δz → 0:
    • Result: −∂v(z,t)/∂z = R′i(z,t) + L′(∂i(z,t)/∂t) (Equation 3.3)

🔄 Kirchhoff's current law (KCL)

  • Apply KCL at the right port:
    • i(z,t) − (G′Δz)v(z+Δz,t) − (C′Δz)(∂v(z+Δz,t)/∂t) − i(z+Δz,t) = 0
  • Rearrange to isolate voltage terms on the right, divide by Δz, then take the limit as Δz → 0:
    • Result: −∂i(z,t)/∂z = G′v(z,t) + C′(∂v(z,t)/∂t) (Equation 3.6)

📝 Time-domain telegrapher's equations

Equations 3.3 and 3.6 are the telegrapher's equations: coupled (simultaneous) differential equations that can be solved for v(z,t) and i(z,t) given R′, G′, L′, C′ and suitable boundary conditions.

  • They describe how voltage and current change with both position (z) and time (t).
  • "Coupled" means each equation involves both voltage and current, so you must solve them together.

🌊 Phasor representation simplification

🌊 Why phasor form is preferred

  • The time-domain equations are "usually more than we need or want."
  • If we are only interested in the response to a sinusoidal stimulus, phasor representation allows "considerable simplification."
  • Don't confuse: phasor form is not a different physical model—it's a mathematical tool for sinusoidal steady-state analysis (see Section 1.5 for a refresher).

🔢 Converting to phasors

  • Define phasors Ṽ(z) and Ĩ(z) through the usual relationship:
    • v(z,t) = Re{Ṽ(z) exp(jωt)}
    • i(z,t) = Re{Ĩ(z) exp(jωt)}
  • Key transformations:
    • Spatial derivative: ∂v(z,t)/∂z becomes ∂Ṽ(z)/∂z (no change in form)
    • Time derivative: ∂i(z,t)/∂t becomes jωĨ(z) (time derivative replaced by multiplication by jω)

⚙️ Phasor telegrapher's equations

Equations 3.9 and 3.10 are the telegrapher's equations in phasor representation.

  • Equation 3.9: −∂Ṽ(z)/∂z = [R′ + jωL′]Ĩ(z)
  • Equation 3.10: −∂Ĩ(z)/∂z = [G′ + jωC′]Ṽ(z)
  • Principal advantage: "we no longer need to contend with derivatives with respect to time—only derivatives with respect to distance remain. This considerably simplifies the equations."

🔗 Relationship to the lumped-element model

🔗 Four parameters from the model

The telegrapher's equations depend on the four per-unit-length parameters from Section 3.4:

ParameterMeaningUnits
R′Series-combined ohmic resistance of both conductorsΩ/m
G′Leakage conductance between conductors (not 1/R′)S/m
L′Inductance (energy storage in magnetic fields)H/m
C′Capacitance (energy storage in electric fields)F/m
  • These parameters capture the physical properties of the transmission line (geometry and materials).
  • Methods for computing these parameters are addressed elsewhere in the book.

🧩 How the equations connect to circuit laws

  • The telegrapher's equations are not arbitrary; they are derived directly from Kirchhoff's voltage and current laws applied to the lumped-element model.
  • Example: the R′i(z,t) term in Equation 3.3 comes from the voltage drop across the series resistance R′Δz; the L′(∂i(z,t)/∂t) term comes from the voltage drop across the series inductance L′Δz.
  • Similarly, the G′v(z,t) and C′(∂v(z,t)/∂t) terms in Equation 3.6 come from the shunt conductance and capacitance.
21

Wave Equation for a TEM Transmission Line

3.6 Wave Equation for a TEM Transmission Line

🧭 Overview

🧠 One-sentence thesis

The wave equation reduces the two coupled telegrapher's equations into a single equation for each of voltage and current, revealing that both quantities propagate as waves along the transmission line and differ only by a multiplicative constant (the characteristic impedance).

📌 Key points (3–5)

  • Why derive a wave equation: The telegrapher's equations require solving for both voltage and current simultaneously; the wave equation allows solving for each independently.
  • What the propagation constant captures: γ encodes how materials, geometry, and frequency determine how voltage and current vary with distance along the line.
  • Key insight from identical equations: voltage and current satisfy the same differential equation, meaning they differ only by a multiplicative constant (an impedance).
  • Common confusion: "same equation" does not mean voltage equals current; it means their spatial behavior is linked by a constant ratio.
  • General solution structure: both voltage and current are sums of two exponential terms representing waves traveling in opposite directions (+ z and − z).

🔧 Deriving the wave equation

🔧 Starting from telegrapher's equations

The phasor-form telegrapher's equations are:

  • Negative derivative of voltage with respect to z equals (R′ + jωL′) times current
  • Negative derivative of current with respect to z equals (G′ + jωC′) times voltage

Where R′, L′, G′, C′ are the lumped-element equivalent circuit parameters per unit length.

🔧 Elimination procedure

Step 1: Differentiate the voltage equation with respect to z again (take the second derivative of voltage).

Step 2: Use the current equation to eliminate the current derivative term.

Result: The second derivative of voltage with respect to z equals γ² times voltage, where γ² = (R′ + jωL′)(G′ + jωC′).

Same process for current: Following the same procedure starting from the current equation yields an identical form for current.

🔧 Standard wave equation form

The final equations are written as:

  • Second derivative of voltage phasor with respect to z, minus γ² times voltage phasor, equals zero
  • Second derivative of current phasor with respect to z, minus γ² times current phasor, equals zero

This is a linear homogeneous differential equation.

📐 The propagation constant

📐 Definition and meaning

Propagation constant γ (units: per meter): the principal square root of (R′ + jωL′)(G′ + jωC′).

  • It captures the combined effect of:

    • Materials (resistance, conductance, capacitance, inductance per unit length)
    • Geometry (cross-sectional shape, which determines R′, L′, G′, C′)
    • Frequency (ω appears in the formula)
  • It determines how potential and current vary with distance on a TEM transmission line.

📐 Why it matters

The propagation constant is the single parameter that governs wave behavior along the line. Once γ is known, the spatial variation of both voltage and current is determined.

🌊 General solutions and physical interpretation

🌊 Form of the solutions

The general solutions are:

  • Voltage phasor = V⁺₀ times e^(−γz) plus V⁻₀ times e^(+γz)
  • Current phasor = I⁺₀ times e^(−γz) plus I⁻₀ times e^(+γz)

Where V⁺₀, V⁻₀, I⁺₀, I⁻₀ are complex-valued constants.

🌊 Two-wave interpretation

Each solution is a sum of two terms:

  • The e^(−γz) term represents a wave propagating in the + z direction
  • The e^(+γz) term represents a wave propagating in the − z direction

Example: On a transmission line, the forward wave might come from a source at one end, while the backward wave arises from reflections at the other end or from discontinuities.

🌊 What determines the constants

The constants V⁺₀, V⁻₀, I⁺₀, I⁻₀ are determined by boundary conditions:

  • Constraints on voltage and current at specific positions along the line
  • May represent sources, loads, or discontinuities in materials/geometry

The excerpt encourages verifying that these solutions satisfy the wave equations for any values of the constants.

🔗 Relationship between voltage and current

🔗 Same equation, different quantities

Both voltage and current satisfy the same linear homogeneous differential equation (the wave equation with the same γ²).

Don't confuse: This does not mean voltage equals current.

What it means: Voltage and current can differ by no more than a multiplicative constant.

🔗 The constant must be an impedance

  • Voltage has units of potential
  • Current has units of current
  • Therefore, the multiplicative constant linking them must have units of impedance (voltage divided by current)

This impedance is known as the characteristic impedance.

The characteristic impedance is determined separately (mentioned as covered in Section 3.7) and depends only on the transmission line's properties, not on boundary conditions or position.

22

Characteristic Impedance

3.7 Characteristic Impedance

🧭 Overview

🧠 One-sentence thesis

Characteristic impedance is the ratio of voltage to current for a wave traveling in a single direction on a transmission line, determined solely by the line's materials and geometry, not by its length, excitation, or termination.

📌 Key points (3–5)

  • What it is: the ratio of voltage to current for a wave propagating in a single direction along a transmission line.
  • What determines it: only the materials and cross-sectional geometry of the transmission line (the things that determine γ), not length, excitation, termination, or position.
  • Common confusion: Z₀ is not the ratio of total voltage to total current at a point; it relates only the voltage and current waves traveling in the same direction.
  • Why real-valued matters: transmission lines are normally designed so Z₀ has no imaginary component, because imaginary impedance represents energy storage (capacitors/inductors) whereas transmission lines are for energy transfer.
  • How it's derived: from the wave equations and telegrapher's equations using the equivalent circuit parameters R′, L′, G′, and C′.

🔍 Core concept and definition

🔍 What characteristic impedance means

Characteristic impedance Z₀ (Ω): the ratio of potential to current in a wave traveling in a single direction along the transmission line.

  • Both voltage phasor Ṽ(z) and current phasor Ĩ(z) satisfy the same linear homogeneous differential equation (the wave equation).
  • Because they satisfy the same equation, they can differ by no more than a multiplicative constant.
  • Since Ṽ(z) is potential and Ĩ(z) is current, that constant must have units of impedance.
  • This impedance is the characteristic impedance.

⚠️ Critical distinction

Don't confuse: Z₀ is not the ratio of Ṽ(z) to Ĩ(z) in general.

  • The general solutions contain waves traveling in both +z and −z directions:
    • Ṽ(z) = V⁺₀ e^(−γz) + V⁻₀ e^(+γz)
    • Ĩ(z) = I⁺₀ e^(−γz) + I⁻₀ e^(+γz)
  • Z₀ relates only the potential and current waves traveling in the same direction.
  • Example: for the +z direction wave, V⁺₀/I⁺₀ = Z₀; for the −z direction wave, −V⁻₀/I⁻₀ = Z₀.

🧮 Mathematical derivation

🧮 Starting equations

The derivation begins with:

  1. Wave equations (same form for voltage and current):

    • ∂²/∂z² Ṽ(z) − γ² Ṽ(z) = 0
    • ∂²/∂z² Ĩ(z) − γ² Ĩ(z) = 0
    • Where γ = √[(R′ + jωL′)(G′ + jωC′)]
  2. Telegrapher's equations:

    • −∂/∂z Ṽ(z) = [R′ + jωL′] Ĩ(z)
    • −∂/∂z Ĩ(z) = [G′ + jωC′] Ṽ(z)
  3. General solutions:

    • Ṽ(z) = V⁺₀ e^(−γz) + V⁻₀ e^(+γz)
    • Ĩ(z) = I⁺₀ e^(−γz) + I⁻₀ e^(+γz)

🔧 Derivation steps

The excerpt derives Z₀ by:

  1. Differentiating the voltage equation with respect to z:

    • ∂/∂z Ṽ(z) = −γ [V⁺₀ e^(−γz) − V⁻₀ e^(+γz)]
  2. Substituting into the first telegrapher's equation:

    • γ [V⁺₀ e^(−γz) − V⁻₀ e^(+γz)] = [R′ + jωL′] Ĩ(z)
  3. Solving for Ĩ(z):

    • Ĩ(z) = [γ/(R′ + jωL′)] [V⁺₀ e^(−γz) − V⁻₀ e^(+γz)]
  4. Comparing with the general solution for Ĩ(z) to find:

    • I⁺₀ = [γ/(R′ + jωL′)] V⁺₀
    • I⁻₀ = −[γ/(R′ + jωL′)] V⁻₀
  5. Defining Z₀ = (R′ + jωL′)/γ and observing:

    • V⁺₀/I⁺₀ = −V⁻₀/I⁻₀ = Z₀

📐 Final expressions

Two equivalent forms for characteristic impedance:

FormExpressionWhat it shows
In terms of γZ₀ = (R′ + jωL′)/γRelationship to propagation constant
In terms of circuit parametersZ₀ = √[(R′ + jωL′)/(G′ + jωC′)]Depends only on equivalent circuit parameters

Where:

  • R′ = resistance per unit length
  • L′ = inductance per unit length
  • G′ = conductance per unit length
  • C′ = capacitance per unit length

🎯 Key properties and design considerations

🎯 What determines Z₀

Characteristic impedance is so-named because it depends only on:

  • The materials of the transmission line
  • The cross-sectional geometry of the transmission line
  • These are the things that determine γ

It does not depend on:

  • Length of the line
  • Excitation (source)
  • Termination (load)
  • Position along the line

⚡ Real-valued vs complex-valued impedance

Design principle: Transmission lines are normally designed to have a characteristic impedance that is completely real-valued (no imaginary component).

Why this matters:

  • The imaginary component of an impedance represents energy storage (like capacitors and inductors).
  • The purpose of a transmission line is energy transfer, not storage.
  • A real-valued Z₀ means the line efficiently transfers energy without storing it.

Example: If Z₀ had a large imaginary part, the line would act like a reactive component (capacitor or inductor) rather than a pure transmission medium.

23

Wave Propagation on a TEM Transmission Line

3.8 Wave Propagation on a TEM Transmission Line

🧭 Overview

🧠 One-sentence thesis

Voltage and current on a transmission line propagate as damped sinusoidal waves traveling in opposite directions, each characterized by phase velocity, wavelength, and attenuation, with the characteristic impedance relating the potential and current of waves traveling in the same direction.

📌 Key points (3–5)

  • Wave structure: The phasor expressions represent two waves—one traveling in the +z direction (e^(−γz)) and one in the −z direction (e^(+γz))—each with exponential damping and sinusoidal oscillation.
  • Propagation constant decomposition: γ = α + jβ, where α (attenuation constant) controls magnitude decay and β (phase propagation constant) controls phase change with distance.
  • Phase velocity: The speed at which constant-phase points move is v_p = ω/β = λf, derived by tracking how phase changes over time and distance.
  • Common confusion: Z₀ is not the ratio of total voltage to total current at a point; it relates only the voltage and current of a wave traveling in the same direction.
  • Lossless vs lossy lines: When α = 0 the line is lossless; when α > 0 the line is lossy, with greater α meaning faster magnitude decay with distance.

🌊 Wave structure and propagation constant

🌊 Decomposing the propagation constant

The excerpt defines:

  • α ≡ Re{γ} (real part of γ)
  • β ≡ Im{γ} (imaginary part of γ)
  • Therefore γ = α + jβ

This decomposition separates two physical effects:

  • α (attenuation constant): controls how quickly the wave's magnitude decreases with distance (units: 1/m or Np/m, "nepers per meter")
  • β (phase propagation constant): controls how quickly the wave's phase changes with distance (units: rad/m)

🔄 Forward and backward waves

The general phasor expressions are:

  • Voltage: Ṽ(z) = V₀⁺ e^(−γz) + V₀⁻ e^(+γz)
  • Current: Ĩ(z) = I₀⁺ e^(−γz) + I₀⁻ e^(+γz)

Using e^(±γz) = e^(±αz) e^(±jβz), the excerpt shows:

  • e^(−γz) represents a damped sinusoidal wave traveling in the +z direction
  • e^(+γz) represents a damped sinusoidal wave traveling in the −z direction

In the time domain:

  • Re{e^(±γz) e^(jωt)} = e^(±αz) cos(ωt ± βz)

Example: For a wave traveling in +z, the time-domain voltage is v⁺(z,t) = |V₀⁺| e^(−αz) cos(ωt − βz + ψ), where ψ is the phase of V₀⁺.

🔍 Lossless vs lossy transmission lines

A transmission line is lossless if α = 0 (no magnitude decay with distance).

A transmission line is lossy (or "low loss") if α > 0 (magnitude decreases with distance).

  • Greater α → faster decay
  • The excerpt notes that even "low loss" lines may have significant total loss over long distances, but still satisfy certain approximations

📏 Wavelength and phase velocity

📏 Wavelength

Wavelength λ = 2π/β

This is the spatial period of the sinusoidal wave—the distance over which the phase changes by 2π radians.

🚀 Phase velocity

Phase velocity v_p = ω/β = λf is the speed at which a point of constant phase travels along the line.

How it's derived:

  • Consider a point of constant phase: ωt − βz + φ = constant
  • At time t, the phase is at position z
  • At time t + Δt, the same phase value is at position z + Δz
  • Setting ωt − βz + φ = ω(t + Δt) − β(z + Δz) + φ
  • Solving for Δz/Δt gives v_p = ω/β

Don't confuse: Phase velocity is the speed of a phase point, not necessarily the speed of energy or information transfer.

⚡ Characteristic impedance and wave direction

⚡ Characteristic impedance definition

The excerpt earlier defined:

Characteristic impedance Z₀ (Ω) is the ratio of potential to current in a wave traveling in a single direction along the transmission line.

In general form:

  • Z₀ = √[(R′ + jωL′)/(G′ + jωC′)]

where R′, L′, G′, C′ are the per-unit-length equivalent circuit parameters.

🔀 Relating voltage and current by direction

For the +z-traveling wave:

  • Ĩ⁺(z) = (V₀⁺/Z₀) e^(−γz)
  • The current is in phase with the voltage (same sign)

For the −z-traveling wave:

  • Ĩ⁻(z) = −(V₀⁻/Z₀) e^(+γz)
  • Note the negative sign

Physical meaning of the negative sign: Where the potential of the −z-traveling wave is positive, the current flows in the −z direction (opposite to the +z reference direction used in the telegrapher's equations).

⚠️ Common confusion about Z₀

The excerpt emphasizes:

  • Z₀ is not the ratio of Ṽ(z) to Ĩ(z) in general
  • Z₀ relates only the potential and current waves traveling in the same direction
  • When waves travel in both directions simultaneously, the total voltage and current are sums: Ṽ(z) = Ṽ⁺(z) + Ṽ⁻(z) and Ĩ(z) = Ĩ⁺(z) + Ĩ⁻(z)

🎯 Design consideration

The excerpt notes that transmission lines are normally designed so Z₀ is completely real-valued (no imaginary component):

  • Imaginary impedance represents energy storage (like capacitors and inductors)
  • The purpose of a transmission line is energy transfer, not storage
  • Real Z₀ minimizes reactive effects

🧮 Low-loss and lossless approximations

🧮 Low-loss conditions

The excerpt defines "low loss" as meeting:

  • R′ ≪ ωL′
  • G′ ≪ ωC′

where R′ represents conductor resistance and G′ represents leakage current through the spacer material.

📐 Simplified expressions under low-loss approximation

When the low-loss conditions hold:

ParameterGeneral expressionLow-loss approximation
γ√[(R′ + jωL′)(G′ + jωC′)]jω√(L′C′)
αRe{γ}≈ 0
βIm{γ}ω√(L′C′)
v_pω/β1/√(L′C′)
Z₀√[(R′ + jωL′)/(G′ + jωC′)]√(L′/C′)

Key observations:

  • If the line is strictly lossless (R′ = G′ = 0), these are exact, not approximations
  • Z₀ and v_p become approximately independent of frequency under low-loss conditions
  • The low-loss approximation for β is commonly used even when α is significant enough to matter for total loss calculations

Example: A coaxial cable connecting an antenna to a radio may have important loss to consider in design, but still satisfies the low-loss inequalities, so β ≈ ω√(L′C′) is used while α is not approximated as zero.

🔁 Bidirectional propagation and standing waves

🔁 Simultaneous forward and backward waves

The excerpt notes it is frequently necessary to consider waves traveling in both directions simultaneously, especially when there is reflection from a discontinuity (e.g., a termination that is not perfectly impedance-matched).

The total voltage and current are:

  • Ṽ(z) = Ṽ⁺(z) + Ṽ⁻(z) (incident + reflected)
  • Ĩ(z) = Ĩ⁺(z) + Ĩ⁻(z)

🌀 Standing waves

The existence of waves propagating simultaneously in both directions gives rise to a phenomenon known as a standing wave.

The excerpt indicates that:

  • Standing waves are addressed in Section 3.13
  • Calculation of the reflection coefficients V₀⁻ and I₀⁻ is addressed in Section 3.12

(These sections are not included in the provided excerpt, so no further detail is given here.)

24

Lossless and Low-Loss Transmission Lines

3.9 Lossless and Low-Loss Transmission Lines

🧭 Overview

🧠 One-sentence thesis

Low-loss transmission lines allow significant simplification of transmission line theory by neglecting small resistive and conductive losses, yielding frequency-independent characteristic impedance and phase velocity.

📌 Key points (3–5)

  • What "low loss" means: the resistive loss R′ is much smaller than the inductive reactance ωL′, and the conductive loss G′ is much smaller than the capacitive susceptance ωC′.
  • Key simplifications: under low-loss conditions, the attenuation constant α ≈ 0, and both characteristic impedance Z₀ and phase velocity vₚ become approximately independent of frequency.
  • Common confusion: "low loss" does not mean "no loss"—a line can satisfy the low-loss inequalities yet still have significant loss that matters in practice (e.g., coaxial cables for antennas).
  • Lossless vs low-loss: when R′ = G′ = 0 exactly, the simplified expressions are exact, not approximations.
  • Practical importance: these approximations are widely used because most practical transmission lines meet the low-loss conditions, and the resulting expressions are much simpler.

📐 Defining loss in transmission lines

📐 What loss represents

"Loss" refers to the reduction of magnitude as a wave propagates through space.

  • In the lumped-element equivalent circuit model, two parameters represent loss mechanisms:
    • R′: resistance of the conductors themselves.
    • G′: undesirable current induced between conductors through the spacing material (leakage).
  • Both parameters cause wave amplitude to decrease as it travels along the line.

📐 The general propagation constant

  • The propagation constant γ in general form is:
    • γ = square root of (R′ + jωL′)(G′ + jωC′)
  • This expression includes both loss (R′, G′) and reactive (L′, C′) components.
  • The real part of γ is the attenuation constant α; the imaginary part is the phase constant β.

🔧 Low-loss conditions and simplifications

🔧 The two low-loss inequalities

The excerpt defines "low loss" as meeting both:

  • R′ ≪ ωL′: resistive loss is much smaller than inductive reactance.
  • G′ ≪ ωC′: conductive loss is much smaller than capacitive susceptance.

When these hold, the loss terms become negligible compared to the reactive terms.

🔧 Simplified propagation constant

Under low-loss conditions:

  • γ ≈ square root of (jωL′)(jωC′) = jω times square root of L′C′
  • This yields:
    • α ≈ 0: attenuation is approximately zero (low-loss approximation).
    • β ≈ ω times square root of L′C′: phase constant depends only on frequency and the reactive parameters.
    • vₚ = ω/β ≈ 1 / square root of L′C′: phase velocity is approximately constant with frequency.

🔧 Simplified characteristic impedance

  • The general expression Z₀ = square root of (R′ + jωL′)/(G′ + jωC′) simplifies to:
    • Z₀ ≈ square root of L′/C′ (low-loss approximation).
  • This simplified Z₀ is approximately independent of frequency when low-loss conditions hold.
  • Example: a coaxial cable connecting an antenna to a radio may have significant loss in practice, but still satisfies the low-loss inequalities, so the simplified β expression is used while α might not be approximated as zero.

🔧 Lossless vs low-loss distinction

ConditionR′ and G′Simplified expressionsStatus
Strictly losslessR′ = G′ = 0 exactlyExact, not approximationsIdeal case
Low-lossR′ ≪ ωL′ and G′ ≪ ωC′ApproximationsCommon in practice
  • Don't confuse: if the line is strictly lossless, the simplified expressions are exact; if low-loss, they are approximations.
  • The excerpt emphasizes that practical transmission lines typically meet the low-loss conditions, making these approximations widely applicable.

🔌 Coaxial transmission line example

🔌 Structure and characteristics

  • Coaxial lines consist of metallic inner and outer conductors separated by a spacer material (dielectric).
  • The outer conductor is also called the "shield" and provides isolation from nearby objects and fields.
  • The spacer material typically has permeability μ ≈ μ₀ and permittivity εₛ ranging from near ε₀ (air-filled) to 2–3 times ε₀.
  • Coaxial line is "single-ended": the geometry is asymmetric and the shield is normally grounded at both ends.
  • These characteristics make coaxial line attractive for connecting single-ended circuits in widely-separated locations and for connecting antennas to receivers and transmitters.

🔌 Equivalent circuit parameters

The excerpt provides expressions for the per-unit-length parameters:

  • C′ = 2πεₛ / ln(b/a): capacitance per unit length, where a and b are the radii of inner and outer conductors.
  • L′ = (μ₀/2π) ln(b/a): inductance per unit length.
  • G′ = 2πσₛ / ln(b/a): loss conductance, depending on the conductance σₛ of the spacer material.
  • R′: resistance per unit length is relatively difficult to quantify because:
    • Inner and outer conductors may consist of different materials or compositions.
    • The outer conductor may be a mesh, braid, or composite.
    • Resistance varies significantly with frequency, unlike C′, L′, and G′.

🔌 Low-loss approximation for coaxial line

  • The low-loss conditions R′ ≪ ωL′ and G′ ≪ ωC′ are often applicable for coaxial lines.
  • This means R′ and G′ are important only if it is necessary to compute loss.
  • Under low-loss conditions, the characteristic impedance simplifies to:
    • Z₀ ≈ (1/2π) times square root of (μ₀/εₛ) times ln(b/a) (low-loss).
  • Example: a coaxial cable used to connect an antenna on a tower to a radio near the ground typically has loss that is important to consider in analysis and design, but nevertheless satisfies the low-loss inequalities—so the low-loss expression for β is used, but α might not be approximated as zero.

🌊 Standing waves and reflections (context)

🌊 When reflections occur

  • The excerpt mentions that reflections arise from a discontinuity, e.g., a termination that is not perfectly impedance-matched.
  • In this case, the total potential and current are the sum of incident (+z-traveling) and reflected (−z-traveling) waves.
  • The existence of waves propagating simultaneously in both directions gives rise to a phenomenon known as a standing wave.
  • Don't confuse: this section (3.9) focuses on low-loss simplifications; standing waves and reflection coefficients are addressed in later sections (3.12 and 3.13).
25

Coaxial Line

3.10 Coaxial Line

🧭 Overview

🧠 One-sentence thesis

Coaxial transmission lines use a simple geometry of inner and outer conductors separated by a dielectric spacer, and under common low-loss conditions their characteristic impedance and phase velocity can be expressed directly in terms of geometry and material properties without computing lumped-element circuit parameters.

📌 Key points (3–5)

  • Structure and use: coaxial line consists of metallic inner and outer conductors separated by a low-loss dielectric spacer; it is single-ended and widely used to connect antennas, receivers, and transmitters.
  • Low-loss simplification: when R′ ≪ ωL′ and G′ ≪ ωC′, characteristic impedance Z₀ and phase velocity vₚ can be expressed directly in terms of geometry (radii a and b) and spacer permittivity (εᵣ), without first computing L′ and C′.
  • Common confusion: "low loss" does not mean "no loss"—even cables with significant loss (important for design) often satisfy the low-loss conditions, so the low-loss β expression is used but α may not be approximated as zero.
  • Frequency independence: under low-loss conditions, Z₀ and vₚ are approximately independent of frequency, though R′ varies significantly with frequency.
  • Why it matters: the outer conductor (shield) provides isolation from nearby objects and fields, making coaxial line attractive for connecting single-ended circuits over long distances.

🔌 Physical structure and characteristics

🔌 Geometry and materials

  • The cross-section shows an inner conductor of radius a and an outer conductor of radius b, separated by a spacer material.
  • The spacer is typically a low-loss dielectric with permeability μ ≈ μ₀ and permittivity εₛ ranging from very near ε₀ (air-filled) to 2–3 times ε₀.
  • The outer conductor is also called the "shield" because it provides a high degree of isolation from nearby objects and electromagnetic fields.

📡 Single-ended nature

Single-ended: the conductor geometry is asymmetric and the shield is normally attached to ground at both ends.

  • This asymmetry makes coaxial line suitable for connecting single-ended circuits in widely-separated locations.
  • It is also used for connecting antennas to receivers and transmitters.

🌊 Field structure

  • Coaxial lines exhibit TEM (transverse electromagnetic) field structure.
  • The electric and magnetic fields are confined between the inner and outer conductors, propagating along the line.
  • Example: the wave propagates away from the viewer in the figure, with fields perpendicular to the direction of propagation.

📐 Equivalent circuit parameters

📐 Capacitance per unit length

  • The capacitance per unit length is given by:
    • C′ = 2πεₛ / ln(b/a)
  • It depends on the spacer permittivity εₛ and the ratio of outer to inner radii.

📐 Inductance per unit length

  • The inductance per unit length is:
    • L′ = (μ₀ / 2π) × ln(b/a)
  • It depends on the permeability of free space and the geometry ratio.

📐 Loss conductance

  • The loss conductance per unit length is:
    • G′ = 2πσₛ / ln(b/a)
  • It depends on the conductance σₛ of the spacer material.

📐 Resistance per unit length (difficult to quantify)

  • R′ is relatively difficult to quantify for several reasons:
    • Inner and outer conductors may consist of different materials or compositions (trading off conductivity, strength, weight, and cost).
    • The outer conductor may be a metal mesh, braid, or composite rather than homogeneous material.
    • Resistance varies significantly with frequency, whereas C′, L′, and G′ exhibit relatively little variation from their electrostatic and magnetostatic values.
  • These factors make it hard to devise a single simple and generally applicable expression for R′.

🧮 Low-loss approximations

🧮 When low-loss conditions apply

  • The low-loss conditions are:
    • R′ ≪ ωL′
    • G′ ≪ ωC′
  • These conditions are often applicable in practice, so R′ and G′ are important only if it is necessary to compute loss.
  • Don't confuse: "low loss" does not mean "no loss." A coaxial cable (e.g., connecting an antenna on a tower to a radio) may have loss that is important to consider in analysis and design, but still satisfies the low-loss conditions. In this case, the low-loss expression for β is used, but α might not be approximated as zero.

🧮 Characteristic impedance under low-loss

  • Under low-loss conditions, the characteristic impedance is:
    • Z₀ ≈ √(L′/C′)
  • Substituting the expressions for L′ and C′:
    • Z₀ = (1 / 2π) × √(μ₀/εₛ) × ln(b/a)
  • Expressing the spacer permittivity as εₛ = εᵣ ε₀ and using √(μ₀/ε₀) as a constant:
    • Z₀ ≈ (60 Ω / √εᵣ) × ln(b/a)
  • This allows Z₀ to be expressed directly in terms of geometry (a and b) and material (εᵣ) without first computing L′ and C′.

🧮 Phase velocity under low-loss

  • Under low-loss conditions, the phase velocity is:
    • vₚ ≈ 1 / √(L′C′)
  • This simplifies to:
    • vₚ = c / √εᵣ
  • where c = 1 / √(μ₀ε₀) is the speed of light in free space.
  • In other words, the phase velocity in a low-loss coaxial line is approximately the speed of electromagnetic propagation in free space divided by the square root of the relative permittivity of the spacer material.
  • Example: an air-filled coaxial line (εᵣ ≈ 1) has phase velocity approximately equal to c, but a dielectric spacer reduces the phase velocity.

🧮 Frequency independence

  • Under low-loss conditions, Z₀ and vₚ are approximately independent of frequency.
  • However, R′ varies significantly with frequency (e.g., increasing approximately in proportion to the square root of frequency).

📦 Example: RG-59 coaxial cable

📦 Physical parameters

  • RG-59 is a very common type of coaxial line.
  • Radii: a ≈ 0.292 mm, b ≈ 1.855 mm (mean).
  • Spacer material: polyethylene with εᵣ ≈ 2.25 and conductivity σₛ ≈ 5.9 × 10⁻⁵ S/m.

📦 Computed parameters

  • From the radii: L′ ≈ 370 nH/m.
  • From the spacer permittivity: C′ ≈ 67.7 pF/m.
  • From the spacer conductivity: G′ ≈ 200 μS/m.
  • Typical resistance: R′ is on the order of 0.1 Ω/m near DC, increasing approximately in proportion to the square root of frequency.

📦 Low-loss criteria

  • RG-59 satisfies the low-loss criteria:
    • R′ ≪ ωL′ for f ≫ 43 kHz
    • G′ ≪ ωC′ for f ≫ 470 kHz
  • Under these conditions:
    • Z₀ ≈ √(L′/C′) ≈ 74 Ω
    • This means the ratio of potential to current in a wave traveling in a single direction on RG-59 is about 74 Ω.

📦 Phase velocity and wavelength

  • The phase velocity is:
    • vₚ ≈ 1 / √(L′C′) ≈ 2 × 10⁸ m/s
  • This is about 67% of c.
  • Example: a signal that takes 1 ns to traverse a distance l in free space requires about 1.5 ns to traverse a length-l section of RG-59.
  • Since vₚ = λf, a wavelength in RG-59 is 67% of a wavelength in free space.

📦 Loss (attenuation)

  • Using the full expression for the propagation constant γ = √[(R′ + jωL′)(G′ + jωC′)] with R′ = 0.1 Ω/m, and taking the real part to obtain α:
    • α ≈ 0.01 m⁻¹ at relatively low frequencies, increasing with frequency.
  • Example: the magnitude of the potential or current is decreased by about 50% by traveling a distance of about 70 m (i.e., e⁻ᵅˡ = 0.5 for l ≈ 70 m).
26

Microstrip Line

3.11 Microstrip Line

🧭 Overview

🧠 One-sentence thesis

Microstrip transmission lines—implemented as a narrow metallic trace above a ground plane separated by dielectric—are widely used on printed circuit boards and exhibit quasi-TEM field structure with effective permittivity determining their characteristic impedance and wave propagation speed.

📌 Key points (3–5)

  • What microstrip is: a narrow metallic trace separated from a metallic ground plane by a dielectric slab, naturally implemented on printed circuit boards.
  • Single-ended and asymmetric: the conductor geometry is asymmetric, with the ground plane serving as both a conductor and ground for source and load.
  • Complex field structure: electric and magnetic fields exist both in the dielectric and in the air above it, making simple lumped-element descriptions difficult.
  • Common confusion: microstrip is distinct from stripline, which is a very different transmission line type.
  • Effective permittivity concept: because fields exist in both dielectric and air, an effective relative permittivity (roughly the average of the two media) determines phase velocity and wavelength.

🏗️ Physical structure and geometry

🏗️ Basic construction

A microstrip transmission line consists of a narrow metallic trace separated from a metallic ground plane by a slab of dielectric material.

  • The metallic trace sits on top of a dielectric slab.
  • Below the dielectric is a metallic ground plane.
  • This structure is a natural way to implement transmission lines on printed circuit boards.

📐 Design parameters

The key geometric parameters (shown in Figure 3.16):

ParameterMeaning
WWidth of the metallic trace
hThickness of the dielectric slab
tThickness of the metallic trace
εᵣRelative permittivity of the dielectric
  • Typically t ≪ W and t ≪ h, so the trace thickness can often be neglected (W′ ≈ W).
  • The spacer material is typically low-loss dielectric with permeability approximately equal to free space (μ ≈ μ₀) and relative permittivity εᵣ in the range 2 to about 10.

🔌 Single-ended nature

  • The conductor geometry is asymmetric.
  • The ground plane normally serves as ground for both the source and the load.
  • This distinguishes microstrip from other transmission line types.

Don't confuse: Microstrip is distinct from stripline, which is a very different type of transmission line structure.

⚡ Field structure and TEM approximation

⚡ Quasi-TEM fields

  • Microstrip line nominally exhibits TEM (transverse electromagnetic) field structure.
  • Electric and magnetic fields exist both in the dielectric and in the space above the dielectric (typically air).
  • Figure 3.17 shows this structure, with the wave propagating perpendicular to the field directions.

🌊 Why field structure matters

  • This complex field structure makes it difficult to describe microstrip concisely in terms of lumped-element equivalent circuit parameters (R′, L′, G′, C′).
  • Instead, expressions for characteristic impedance Z₀ directly in terms of h/W and εᵣ are typically used.
  • The fields outside the line are "possibly significant, complicated, and not shown" in simplified diagrams.

🔧 Characteristic impedance

🔧 General expression

A widely-accepted and broadly-applicable expression for characteristic impedance is:

Z₀ ≈ (42.4 Ω / √(εᵣ + 1)) × ln[1 + (4h/W′) × (Φ + √(Φ² + (1 + 1/εᵣ)/(2π²)))]

where:

  • Φ = ((14 + 8/εᵣ)/11) × (4h/W′)
  • W′ is W adjusted to account for trace thickness t
  • When t ≪ W and t ≪ h, then W′ ≈ W

This expression is from Wheeler 1977 and represents a careful approximation accounting for the complex field distribution.

📝 Simpler approximations

  • Simpler approximations for Z₀ are also commonly employed for quick "back of the envelope" calculations.
  • These are limited in the range of h/W for which they are valid.
  • They can usually be shown to be special cases or approximations of the general expression above.

🌐 Wave propagation parameters

🌐 Effective relative permittivity concept

Because fields exist in both the dielectric (εᵣ) and air (εᵣ = 1), an effective relative permittivity εᵣ,eff is used:

εᵣ,eff ≈ (εᵣ + 1)/2

  • This is roughly the average of the relative permittivity of the dielectric slab and free space.
  • The assumption: some fraction of the power in the guided wave is in the dielectric, and the rest is above the dielectric.
  • In practice, manufacturing variations in εᵣ typically make more precise estimates irrelevant.

📏 Phase propagation constant

For low-loss microstrip:

β ≈ ω√(μ₀εᵣ,effε₀) = β₀√εᵣ,eff

where β₀ = ω√(μ₀ε₀) is the free-space phase propagation constant.

  • In other words, the phase propagation constant in microstrip is the free-space value times a correction factor √εᵣ,eff.
  • This comes from approximating the field structure as a uniform plane wave in unbounded media with permeability μ₀ but effective permittivity εᵣ,eff.

📏 Wavelength

λ = 2π/β = λ₀/√εᵣ,eff

where λ₀ is the free-space wavelength c/f.

  • The wavelength in microstrip is shorter than in free space by a factor of √εᵣ,eff.

🚀 Phase velocity

vₚ = ω/β = c/√εᵣ,eff

  • The phase velocity in microstrip is slower than c by a factor of √εᵣ,eff.
  • This is analogous to wave propagation in dielectric media, but uses the effective permittivity instead of the actual dielectric permittivity.

📋 Practical example: 50 Ω microstrip in FR4

📋 FR4 material properties

FR4 is a low-loss fiberglass epoxy dielectric commonly used to make printed circuit boards:

  • Typical slab thickness: h ≈ 1.575 mm
  • Relative permittivity: εᵣ ≈ 4.5

🎯 Design for 50 Ω

To implement a 50 Ω microstrip line in FR4:

  • Since h and εᵣ are fixed, only W remains to set Z₀.
  • Experimentation with the characteristic impedance equation reveals that h/W ≈ 1/2 yields Z₀ ≈ 50 Ω for εᵣ = 4.5.
  • Therefore, W should be about 3.15 mm.

🎯 Propagation characteristics

For this 50 Ω FR4 microstrip:

  • Effective relative permittivity: εᵣ,eff ≈ (4.5 + 1)/2 = 2.75
  • Phase velocity: vₚ ≈ c/√2.75 ≈ 60% of c
  • Wavelength: λ ≈ 60% of the free-space wavelength

Example: A signal that takes 1 ns to traverse a distance in free space would take longer in the microstrip because the phase velocity is reduced.

27

Voltage Reflection Coefficient

3.12 Voltage Reflection Coefficient

🧭 Overview

🧠 One-sentence thesis

The voltage reflection coefficient Γ determines the magnitude and phase of the reflected wave when an incident wave encounters a termination whose impedance differs from the transmission line's characteristic impedance.

📌 Key points (3–5)

  • What Γ measures: the ratio of reflected wave amplitude to incident wave amplitude, determined by the mismatch between terminating impedance Z_L and characteristic impedance Z_0.
  • When reflection occurs: reflection happens whenever Z_L ≠ Z_0; when Z_L = Z_0 (matched load), Γ = 0 and there is no reflection.
  • Range of Γ: the magnitude |Γ| is always between 0 and 1; Γ can be real, imaginary, or complex-valued depending on Z_L.
  • Common confusion: Γ is the voltage reflection coefficient; the current reflection coefficient is −Γ (opposite sign).
  • Standing waves result: when incident and reflected waves combine, they create a standing wave pattern whose magnitude varies sinusoidally along the line but does not change with time.

🔍 The reflection problem setup

🔍 The scenario

  • A wave traveling from the left along a lossless transmission line with characteristic impedance Z_0 arrives at a termination located at z = 0.
  • The termination has impedance Z_L, which may be real, imaginary, or complex-valued.
  • The questions are: when does reflection occur, and what is the reflected wave?

📐 Wave expressions

The incident wave (traveling rightward, in the +z direction) is expressed as:

  • Voltage: Ṽ⁺(z) = V⁺₀ exp(−jβz)
  • V⁺₀ is determined by the source and is effectively "given."

Any reflected wave (traveling leftward, in the −z direction) must have the form:

  • Voltage: Ṽ⁻(z) = V⁻₀ exp(+jβz)
  • The problem is to find V⁻₀ given V⁺₀, Z_0, and Z_L.

🔗 Why voltage is sufficient

  • The potential and current of the incident wave are related by the constant Z_0.
  • Similarly, the potential and current of the reflected wave are related by Z_0.
  • Therefore, it suffices to consider either potential or current; the excerpt chooses potential.

🧮 Deriving the reflection coefficient

🧮 Boundary conditions at z = 0

At the termination interface (z = 0), three conditions must hold:

  1. Definition of terminating impedance:

    Z_L ≜ Ṽ_L / Ĩ_L where Ṽ_L and Ĩ_L are the potential across and current through the termination.

  2. Voltage continuity: Ṽ⁺(0) + Ṽ⁻(0) = Ṽ_L

  3. Current continuity: Ĩ⁺(0) + Ĩ⁻(0) = Ĩ_L

🧮 Solving for the reflected amplitude

Since voltage and current are related by Z_0, the current equation can be rewritten:

  • Ṽ⁺(0)/Z_0 − Ṽ⁻(0)/Z_0 = Ĩ_L
  • Note the minus sign: the reflected wave travels in the opposite direction.

Evaluating at z = 0:

  • V⁺₀ + V⁻₀ = Ṽ_L
  • V⁺₀/Z_0 − V⁻₀/Z_0 = Ĩ_L

Substituting into the definition of Z_L and solving for V⁻₀:

  • V⁻₀ = [(Z_L − Z_0)/(Z_L + Z_0)] V⁺₀

🎯 The voltage reflection coefficient

Γ ≜ (Z_L − Z_0)/(Z_L + Z_0)

The reflected wave amplitude is:

  • V⁻₀ = Γ V⁺₀

Key insight: Γ determines both the magnitude and phase of the reflected wave relative to the incident wave.

⚠️ Current reflection coefficient

  • Don't confuse: Γ is not the ratio of current coefficients.
  • The ratio of current coefficients I⁻₀ to I⁺₀ is actually −Γ (opposite sign).
  • This follows from the opposite direction of travel for the reflected wave.

📊 Common termination cases

📊 Matched load (Z_L = Z_0)

  • When the terminating impedance equals the characteristic impedance, Γ = 0.
  • Result: there is no reflection; V⁻₀ = 0.
  • The termination may be a device with impedance Z_0 or another transmission line with the same characteristic impedance.

📊 Open circuit (Z_L → ∞)

  • An "open circuit" is the absence of a termination.
  • As Z_L → ∞, Γ → +1.
  • Voltage: the reflected voltage wave is in phase with the incident wave.
  • Current: since the current reflection coefficient is −Γ = −1, the reflected current wave is 180° out of phase with the incident current wave, making the total current at the open circuit equal to zero, as expected.

📊 Short circuit (Z_L = 0)

  • "Short circuit" means Z_L = 0, so Γ = −1.
  • Voltage: the phase of Γ is 180°, so the reflected voltage wave cancels the incident voltage wave at the short circuit, making the total potential zero, as it must be.
  • Current: the current reflection coefficient is −Γ = +1, so the reflected current wave is in phase with the incident current wave, and the total current magnitude is non-zero, as expected.

📊 Purely reactive load (Z_L = jX)

  • A purely reactive load (capacitor or inductor) has Z_L = jX, where X is reactance.
  • Inductor: X > 0; capacitor: X < 0.
  • For this case: Γ = (−Z_0 + jX)/(+Z_0 + jX)
  • Magnitude: the numerator and denominator have the same magnitude, so |Γ| = 1.
  • Phase: if φ is the phase of the denominator (+Z_0 + jX), then the phase of Γ is π − 2φ.
  • Result: the phase of Γ is no longer limited to 0° or 180°, but can be any value in between; the reflected wave is phase-shifted by this amount.

📊 General terminations

| Termination type | Z_L | Γ | |Γ| | |------------------|-----|---|-----| | Matched load | Z_0 | 0 | 0 | | Open circuit | ∞ | +1 | 1 | | Short circuit | 0 | −1 | 1 | | Purely reactive | jX | complex | 1 | | General | complex | complex | 0 to 1 |

📊 Range of |Γ|

The magnitude of Γ is always limited to the range 0 to 1:

  • Minimum (|Γ| = 0): occurs when the numerator is zero, i.e., when Z_L = Z_0.
  • Maximum (|Γ| = 1): occurs when Z_L/Z_0 → ∞ (open circuit) or Z_L/Z_0 = 0 (short circuit).
  • General bound: 0 ≤ |Γ| ≤ 1

Any termination (series/parallel combinations of devices) can be expressed as a complex-valued Z_L, and the associated |Γ| falls within this range.

🌊 Standing waves

🌊 What is a standing wave

A standing wave consists of waves moving in opposite directions; these waves add to make a distinct magnitude variation as a function of distance that does not vary in time.

  • The incident wave V⁺₀ exp(−jβz) travels in the +z direction.
  • The reflected wave Γ V⁺₀ exp(+jβz) travels in the −z direction.
  • These waves add to make the total potential: Ṽ(z) = V⁺₀ [exp(−jβz) + Γ exp(+jβz)]

🌊 Magnitude of the standing wave

Through algebraic manipulation (finding |Ṽ(z)|² and taking the square root), the magnitude of the total voltage is:

  • |Ṽ(z)| = |V⁺₀| √[1 + |Γ|² + 2|Γ| cos(2βz + φ)]
  • where φ is the phase of Γ (i.e., Γ = |Γ| exp(jφ)).

Similarly, for current:

  • |Ĩ(z)| = (|V⁺₀|/Z_0) √[1 + |Γ|² − 2|Γ| cos(2βz + φ)]
  • Note the minus sign in the current expression.

Key property: the magnitude varies sinusoidally along the line (as a function of z) but does not vary with time—this is why it is called a "standing" wave.

🌊 Special cases of standing waves

🌊 Matched load

  • When Z_L = Z_0, Γ = 0 and there is no reflection.
  • The expressions reduce to |Ṽ(z)| = |V⁺₀| and |Ĩ(z)| = |V⁺₀|/Z_0, as expected.
  • Result: no standing wave pattern; magnitude is constant along the line.

🌊 Open or short circuit

  • When Γ = ±1 (open circuit: φ = 0; short circuit: φ = π):
    • |Ṽ(z)| = |V⁺₀| √[2 + 2 cos(2βz + φ)]
    • |Ĩ(z)| = (|V⁺₀|/Z_0) √[2 − 2 cos(2βz + φ)]
  • Result: maximum standing wave pattern with |Γ| = 1.

Example (open circuit at z = 0):

  • The voltage magnitude varies from 0 to 2|V⁺₀| along the line.
  • The current magnitude also varies sinusoidally, with nulls and peaks at different locations than the voltage.
  • The pattern is stationary in space (does not move with time).

🌊 Why "standing"

  • Don't confuse: a traveling wave has a magnitude that is constant in space but moves with time.
  • A standing wave has a magnitude that varies in space but is stationary (does not move) in time.
  • The standing wave results from the interference (superposition) of the incident and reflected waves.
28

Standing Waves

3.13 Standing Waves

🧭 Overview

🧠 One-sentence thesis

Standing waves arise when incident and reflected waves traveling in opposite directions on a transmission line interfere to create a magnitude pattern that varies with position but not with time, and the degree of mismatch between the line and its termination determines the characteristics of this pattern.

📌 Key points (3–5)

  • What standing waves are: waves moving in opposite directions that add to produce a magnitude variation along the line that does not change with time.
  • How they form: an incident wave and a reflected wave (determined by the reflection coefficient Γ) combine, creating a sinusoidal magnitude pattern along the transmission line.
  • Key property: the period of the standing wave is always half a wavelength (λ/2), regardless of the reflection coefficient value.
  • Common confusion: voltage maxima correspond to current minima, and vice versa—the two standing wave patterns are out of phase.
  • Why it matters: standing wave ratio (SWR) quantifies the quality of the match between the transmission line and its termination, ranging from 1 (perfect match) to infinity (open or short circuit).

🌊 Formation and mathematics of standing waves

🌊 How incident and reflected waves combine

  • An incident wave traveling in the +z direction is written as V₀⁺ times e to the power of negative j times beta times z.
  • A reflected wave traveling in the opposite direction is Γ times V₀⁺ times e to the power of positive j times beta times z, where Γ is the voltage reflection coefficient.
  • The total potential is the sum of these two waves.
  • The magnitude of the total potential varies sinusoidally along the line, but this variation does not change with time—hence "standing" wave.

📐 Magnitude expression for voltage

The magnitude of the total voltage is: |V(z)| = |V₀⁺| times the square root of [1 + |Γ|² + 2|Γ| cos(2βz + φ)], where φ is the phase of Γ.

  • The cosine term creates the sinusoidal variation along the line.
  • The phase φ depends on the termination impedance.
  • Example: as you move along the line (changing z), the magnitude oscillates between maximum and minimum values.

📐 Magnitude expression for current

The magnitude of the total current is: |I(z)| = (|V₀⁺|/Z₀) times the square root of [1 + |Γ|² − 2|Γ| cos(2βz + φ)].

  • Note the minus sign in front of the cosine term (voltage has a plus sign).
  • This means voltage and current standing waves are out of phase: where voltage is maximum, current is minimum, and vice versa.
  • Don't confuse: both are standing waves, but their patterns are shifted relative to each other.

🔧 Special cases and termination types

🔧 Matched load (Γ = 0)

  • When the termination impedance Z_L equals the characteristic impedance Z₀, the reflection coefficient Γ = 0.
  • No reflection occurs.
  • The voltage magnitude is simply |V₀⁺| everywhere, and the current magnitude is |V₀⁺|/Z₀ everywhere.
  • No standing wave pattern appears—the magnitude is constant along the line.

🔧 Open or short circuit (|Γ| = 1)

  • For an open circuit: Γ = +1 (φ = 0).
  • For a short circuit: Γ = −1 (φ = π).
  • The voltage magnitude becomes |V₀⁺| times the square root of [2 + 2 cos(2βz + φ)].
  • The current magnitude becomes (|V₀⁺|/Z₀) times the square root of [2 − 2 cos(2βz + φ)].
  • The roles of voltage and current are reversed between open and short circuits: where one has a maximum, the other has a minimum.

🔧 Mismatched loads (0 < |Γ| < 1)

  • Most practical situations fall between perfect match and open/short circuit.
  • The standing wave pattern is present but less extreme than for open/short circuits.
  • The magnitude of Γ determines how pronounced the standing wave pattern is.

📏 Period and spatial characteristics

📏 Half-wavelength period

The period of the standing wave is λ/2 (one-half of a wavelength).

  • The frequency argument of the cosine function is 2βz.
  • This can be rewritten as 2π times (β/π) times z, so the frequency of variation is β/π.
  • The period of variation is π/β.
  • Since β = 2π/λ, the period is λ/2.
  • Important: this λ/2 period is true regardless of the value of Γ—it depends only on the wavelength, not on the degree of mismatch.

📏 Voltage maxima and current minima

  • Voltage maxima occur where the cosine term equals +1.
  • Current minima occur at the same locations (because of the minus sign in the current expression).
  • Voltage minima occur where the cosine term equals −1.
  • Current maxima occur at these same locations.
  • Example: if voltage peaks at a certain position, current will be at a minimum there, and vice versa.

📊 Standing wave ratio (SWR)

📊 Definition and meaning

Standing wave ratio (SWR) is defined as the ratio of the maximum magnitude of the standing wave to the minimum magnitude of the standing wave.

  • For voltage: SWR = (maximum |V|) / (minimum |V|).
  • SWR quantifies the quality of the match between the transmission line and its termination.
  • A lower SWR indicates a better match; a higher SWR indicates more reflection and mismatch.

📊 Relationship to reflection coefficient

SWR = (1 + |Γ|) / (1 − |Γ|).

  • The maximum voltage magnitude is |V₀⁺| times (1 + |Γ|), occurring when the cosine equals +1.
  • The minimum voltage magnitude is |V₀⁺| times (1 − |Γ|), occurring when the cosine equals −1.
  • Taking the ratio gives the SWR formula above.
  • The inverse relationship: |Γ| = (SWR − 1) / (SWR + 1).

📊 Range and interpretation

| Termination condition | |Γ| value | SWR value | Interpretation | |---|---|---|---| | Perfectly matched (Z_L = Z₀) | 0 | 1 | No reflection, ideal match | | Typical mismatch | 0 to 1 | 1 to ∞ | Partial reflection | | Open or short circuit | 1 | ∞ | Total reflection, worst mismatch |

  • SWR ranges from 1 (perfect match) to infinity (open/short circuit).
  • SWR < 2 is usually considered a "good match."
  • Some applications require SWR < 1.1 or better.
  • Other applications tolerate SWR of 3 or greater.

📊 Practical examples

Example: What reflection coefficient corresponds to various SWR values?

  • SWR = 1.1 corresponds to |Γ| = 0.0476 (about 5% reflection).
  • SWR = 2.0 corresponds to |Γ| = 1/3 (about 33% reflection).
  • SWR = 3.0 corresponds to |Γ| = 1/2 (50% reflection).

📊 Voltage vs current SWR

  • The term "voltage standing wave ratio" (VSWR) is sometimes used.
  • However, repeating the analysis for current reveals that current SWR equals voltage SWR.
  • Therefore, the term "SWR" alone is sufficient—no need to specify voltage or current.
29

Standing Wave Ratio

3.14 Standing Wave Ratio

🧭 Overview

🧠 One-sentence thesis

Standing wave ratio (SWR) quantifies the quality of impedance matching between a transmission line and its termination by comparing the maximum and minimum magnitudes of the resulting standing wave.

📌 Key points (3–5)

  • What SWR measures: the ratio of maximum to minimum voltage magnitude in the standing wave pattern created by mismatch.
  • How SWR relates to reflection: SWR can be calculated directly from the magnitude of the reflection coefficient Γ, ranging from 1 (perfect match) to infinity (total reflection).
  • Common confusion: SWR is sometimes called "voltage standing wave ratio" (VSWR), but current SWR equals voltage SWR, so the simpler term suffices.
  • Practical interpretation: SWR less than 2 is usually considered a "good match," though requirements vary by application (some need SWR < 1.1, others tolerate SWR ≥ 3).
  • Bidirectional calculation: you can find SWR from the reflection coefficient or work backward to find the reflection coefficient from a measured SWR.

📏 Definition and formula

📏 What SWR is

Standing wave ratio (SWR) is defined as the ratio of the maximum magnitude of the standing wave to minimum magnitude of the standing wave.

  • In terms of voltage: SWR = (maximum |Ṽ|) / (minimum |Ṽ|)
  • This ratio captures how "uneven" the standing wave pattern is along the transmission line.
  • A perfectly matched line has uniform voltage magnitude (no standing wave), so maximum equals minimum and SWR = 1.

🧮 Deriving the SWR formula

The excerpt derives SWR from the standing wave voltage magnitude expression:

  • The magnitude of voltage is: |Ṽ(z)| = |V₀⁺| times the square root of (1 + |Γ|² + 2|Γ| cos(2βz + φ))
  • Maximum occurs when cosine = +1: max |Ṽ| = |V₀⁺|(1 + |Γ|)
  • Minimum occurs when cosine = −1: min |Ṽ| = |V₀⁺|(1 − |Γ|)
  • Taking the ratio: SWR = (1 + |Γ|) / (1 − |Γ|)

This simple formula depends only on the magnitude of the reflection coefficient, not on position or phase.

🔄 Relationship between SWR and reflection coefficient

🔄 From reflection coefficient to SWR

| Reflection coefficient magnitude |Γ| | SWR | Interpretation | |---|---|---| | 0 | 1 | Perfect match (no reflection) | | Between 0 and 1 | Between 1 and ∞ | Partial mismatch | | 1 | ∞ | Total reflection (open or short circuit) |

  • The relationship is shown graphically in Figure 3.21 (referenced in the excerpt).
  • As |Γ| increases from 0 to 1, SWR increases from 1 to infinity in a nonlinear fashion.

🔄 From SWR to reflection coefficient

The excerpt also provides the inverse formula:

  • |Γ| = (SWR − 1) / (SWR + 1)
  • This is useful when you measure SWR and need to know the reflection coefficient.

Example from the excerpt:

  • SWR = 1.1 corresponds to |Γ| = 0.0476
  • SWR = 2.0 corresponds to |Γ| = 1/3
  • SWR = 3.0 corresponds to |Γ| = 1/2

🎯 Practical interpretation

🎯 What counts as a "good match"

The excerpt provides application-dependent guidelines:

  • SWR < 2: usually considered a "good match" in general applications
  • SWR < 1.1: required for some demanding applications (better match needed)
  • SWR ≥ 3: tolerable in other applications (more mismatch acceptable)

Don't confuse: there is no universal "acceptable" SWR—it depends on the specific system requirements and tolerance for reflected power.

🎯 Why voltage vs current SWR doesn't matter

  • The excerpt notes that SWR is sometimes called "voltage standing wave ratio" (VSWR).
  • However, repeating the analysis for current reveals that current SWR equals voltage SWR.
  • Therefore, the simpler term "SWR" is sufficient—no need to specify voltage or current.

🔗 Context: when SWR matters

🔗 Imperfect matching scenarios

The excerpt explains the practical motivation:

  • Precise matching of transmission lines to terminations is often not practical or possible.
  • Whenever significant mismatch exists, a standing wave appears (from Section 3.13).
  • SWR provides a single number to express the "quality of the match."

🔗 Range of mismatch conditions

The excerpt references different termination scenarios:

  • Perfectly matched (Γ = 0): SWR = 1, no standing wave
  • Mismatched loads (0 < |Γ| < 1): SWR between 1 and infinity, partial standing wave
  • Open or short circuit (|Γ| = 1): SWR = ∞, maximum standing wave

Example: if you measure SWR = 2 on a line, you immediately know |Γ| = 1/3, meaning one-third of the incident wave is reflected back.

30

Input Impedance of a Terminated Lossless Transmission Line

3.15 Input Impedance of a Terminated Lossless Transmission Line

🧭 Overview

🧠 One-sentence thesis

The input impedance of a terminated lossless transmission line depends on the load impedance, characteristic impedance, length, and phase propagation constant, and varies periodically with a period of half a wavelength.

📌 Key points (3–5)

  • What determines input impedance: load impedance Z_L, characteristic impedance Z_0, phase propagation constant β, and line length l.
  • Periodic behavior: input impedance repeats every half wavelength (λ/2), meaning all possible impedance values occur within this length.
  • Matched vs unmatched: when Z_L equals Z_0, input impedance equals Z_L; otherwise it depends on transmission line characteristics.
  • Common confusion: changing line length by more than λ/2 produces impedances already achievable with shorter lengths—the pattern repeats.
  • Special cases: short-circuit and open-circuit terminations produce purely imaginary input impedances that can transform one into the other at quarter-wavelength distances.

🔌 General input impedance formula

🔌 Basic definition and setup

Input impedance Z_in: the ratio of voltage to current at the input of the transmission line (at position z = -l).

  • The transmission line is driven from the left and terminated by impedance Z_L on the right.
  • A coordinate system places the source-transmission line interface at z = -l and the load at z = 0.
  • The formula uses expressions for voltage and current from standing wave analysis.

📐 The main equation

The input impedance is given by:

  • Z_in(l) = Z_0 times (1 + Γ times e^(-j2βl)) divided by (1 - Γ times e^(-j2βl))
  • Where Γ (reflection coefficient) = (Z_L - Z_0) divided by (Z_L + Z_0)
  • This can also be written entirely in terms of Z_L and Z_0: Z_in(l) = Z_0 times [(Z_L + jZ_0 tan(βl)) divided by (Z_0 + jZ_L tan(βl))]

Example: If the load matches the line (Z_L = Z_0), then Γ = 0, and Z_in = Z_0 regardless of length.

⚡ Electrical length concept

  • The term βl appears in the equations and is called electrical length.
  • It has units of radians.
  • Why it matters: electrical length expresses physical length relative to wavelength, making analysis independent of frequency.
  • This allows engineers to think in terms of "fractions of a wavelength" rather than absolute distances.

🔄 Periodicity and wavelength dependence

🔄 Half-wavelength period

  • The input impedance is periodic in the length of the transmission line.
  • Period = λ/2 (half a wavelength).
  • This occurs because the complex exponential factors contain 2βl, and since β = 2π/λ, the frequency of variation is β/π.

Don't confuse: This is not about the full wavelength λ, but half of it—λ/2 is the fundamental repeat distance.

🌊 Connection to standing waves

  • The λ/2 period matches the period of the standing wave pattern.
  • Why: both phenomena result from interference between incident and reflected waves.
  • All possible input impedance values are achieved by varying length over just λ/2.
  • Changing length by more than λ/2 produces impedances that could have been obtained with a smaller length change.

⚙️ Short-circuit and open-circuit terminations

⚙️ Short-circuit stub (Z_L = 0)

Stub: a transmission line terminated in an open- or short-circuit.

For a short-circuit termination:

  • Z_L = 0, so Γ = -1
  • Input impedance: Z_in(l) = +jZ_0 tan(βl)
  • Purely imaginary: the real part is always zero
  • At l = 0: Z_in = 0 (as expected—just a short circuit)
  • At l = λ/4: Z_in → ∞ (transforms short into open!)

🔓 Open-circuit stub (Z_L → ∞)

For an open-circuit termination:

  • Z_L → ∞, so Γ = +1
  • Input impedance: Z_in(l) = -jZ_0 cot(βl)
  • Purely imaginary: the real part is always zero
  • At l = 0: Z_in → ∞ (as expected—just an open circuit)
  • At l = λ/4: Z_in = 0 (transforms open into short!)

🔀 Quarter-wavelength transformation

Termination typeAt l = 0At l = λ/4Transformation
Short-circuit (Z_L = 0)Z_in = 0Z_in → ∞Short → Open
Open-circuit (Z_L → ∞)Z_in → ∞Z_in = 0Open → Short

Why this matters: From a lumped-element circuit perspective, this seems unusual, but the length-dependent impedance enables very useful applications in transmission line design.

Example: A short-circuited stub exactly λ/4 long behaves as an open circuit at its input, even though it is physically shorted at the far end.

📊 Periodic variation with length

  • Both short- and open-circuit stubs show λ/2 periodicity.
  • The input impedance cycles through all possible purely imaginary values as length increases.
  • The imaginary part oscillates between positive and negative infinity.
  • Don't confuse: The impedance is always purely reactive (imaginary)—there is no resistive (real) component for lossless stubs.
31

Input Impedance for Open- and Short-Circuit Terminations

3.16 Input Impedance for Open- and Short-Circuit Terminations

🧭 Overview

🧠 One-sentence thesis

A transmission line terminated in a short or open circuit (called a stub) exhibits a purely imaginary input impedance that alternates between open and short conditions every quarter-wavelength, enabling it to replace discrete inductors and capacitors.

📌 Key points (3–5)

  • What a stub is: a transmission line terminated in either an open circuit or a short circuit; its input impedance depends on its physical length.
  • Key behavior: the input impedance is completely imaginary (purely reactive, no resistance) and alternates between zero and infinity every λ/4 increase in length.
  • Transformation property: a short-circuit termination can look like an open circuit at λ/4 length, and vice versa.
  • Common confusion: length matters—the same stub can behave as an inductor (positive reactance) or capacitor (negative reactance) depending on whether it is shorter or longer than λ/4.
  • Practical use: stubs can replace discrete inductors and capacitors at the design frequency, especially useful at high frequencies where discrete components suffer performance limitations.

🔌 What is a stub and why it matters

🔌 Definition and motivation

Stub: a transmission line that is terminated in an open- or short-circuit.

  • From a lumped-element circuit perspective, a short or open termination seems trivial.
  • However, because the input impedance depends on length (as discussed in Section 3.15), stubs enable very useful applications.
  • The key insight: the structure's impedance changes with physical length, creating design flexibility.

🎯 Why study this

  • Stubs exhibit input impedance that varies with length, not just with the termination type.
  • This length-dependent behavior allows stubs to mimic inductors or capacitors.
  • At high frequencies, transmission line stubs outperform discrete components and are less expensive.

⚡ Short-circuit termination behavior

⚡ Deriving the input impedance

  • For a short circuit, the load impedance Z_L = 0, and the reflection coefficient Γ = −1.
  • Starting from the general input impedance formula and substituting these values, then applying trigonometric identities for cosine and sine in terms of exponentials, the result is:
    • Input impedance for short-circuit stub: Z_in(l) = +j Z_0 tan(βl)
  • The derivation uses the identities:
    • cos θ = (1/2)[e^(+jθ) + e^(−jθ)]
    • sin θ = (1/j2)[e^(+jθ) − e^(−jθ)]

📐 What the formula tells us

  • At l = 0: Z_in = 0, as expected (a short circuit with no line).
  • As l approaches λ/4: Z_in approaches infinity—the short circuit has been transformed into an open circuit.
  • Periodicity: Z_in varies periodically with period λ/2, consistent with standing wave theory.
  • The impedance is purely imaginary (the real part is always zero).

🔄 Physical interpretation

  • The short-circuit stub's reactance oscillates between zero and infinity.
  • Example: at exactly λ/4 length, a short-circuit termination looks like an open circuit to the input.
  • This transformation is due to the interference of incident and reflected waves.

🔓 Open-circuit termination behavior

🔓 Deriving the input impedance

  • For an open circuit, Z_L → ∞, and the reflection coefficient Γ = +1.
  • Following the same mathematical procedure as for the short-circuit case, the result is:
    • Input impedance for open-circuit stub: Z_in(l) = −j Z_0 cot(βl)

📐 What the formula tells us

  • At l = 0: Z_in → ∞, as expected (an open circuit with no line).
  • At l = λ/4: Z_in = 0—the open circuit has been transformed into a short circuit.
  • Periodicity: same λ/2 periodicity as the short-circuit case.
  • Again, the impedance is purely imaginary (real part is always zero).

🔄 Physical interpretation

  • The open-circuit stub's reactance oscillates between infinity and zero.
  • Example: at exactly λ/4 length, an open-circuit termination looks like a short circuit to the input.
  • The transformation is the opposite of the short-circuit case.

📊 Summary of stub behavior

📊 Key properties

PropertyShort-circuit stubOpen-circuit stub
FormulaZ_in = +j Z_0 tan(βl)Z_in = −j Z_0 cot(βl)
At l = 0Z_in = 0 (short)Z_in → ∞ (open)
At l = λ/4Z_in → ∞ (open)Z_in = 0 (short)
Real partAlways zeroAlways zero
Periodicityλ/2λ/2

🔁 Alternating behavior

The input impedance of a short- or open-circuited lossless transmission line alternates between open-circuit (Z_in → ∞) and short-circuit (Z_in = 0) conditions with each λ/4-increase in length.

  • Every quarter-wavelength, the impedance flips between the two extremes.
  • This alternation is a direct consequence of standing wave patterns.

🧮 Electrical length

  • The argument βl in the formulas is called electrical length.
  • It has units of radians.
  • Electrical length expresses physical length with respect to wavelength, making analysis independent of frequency.

🛠️ Practical applications of stubs

🛠️ Replacing inductors and capacitors

  • Short-circuit stub as inductor: for lengths less than λ/4, the reactance is positive, so the stub "looks" like an inductor.
  • Short-circuit stub as capacitor: for lengths between λ/4 and λ/2, the reactance is negative, so the stub "looks" like a capacitor.
  • Open-circuit stubs can be used similarly.
  • The input impedance at the design frequency is identical to that of the discrete component.

⚠️ Frequency response consideration

  • The variation of reactance with frequency will not be identical to a discrete inductor or capacitor.
  • This may or may not be a concern depending on bandwidth and frequency response requirements.
  • Don't confuse: stubs match the impedance at one frequency, but their behavior across a range of frequencies differs from discrete components.

✅ Advantages and drawbacks

AspectAdvantageDrawback
High-frequency performanceStubs do not suffer the performance limitations of discrete devices at high frequencies
CostLess expensive than discrete components
SizeStubs are typically much larger than the discrete devices they replace

🎛️ Design applications

  • Stubs enable implementation of:
    • Impedance matching circuits
    • Filters
    • Other devices entirely from transmission lines, with fewer or no discrete inductors or capacitors required.
  • Example mentioned: emitter induction in low-noise amplifiers using bipolar transistors in common-emitter configuration—a small inductance between emitter and ground can be implemented with a short-circuited stub.
32

Applications of Open- and Short-Circuited Transmission Line Stubs

3.17 Applications of Open- and Short-Circuited Transmission Line Stubs

🧭 Overview

🧠 One-sentence thesis

Open- and short-circuited transmission line stubs can replace discrete inductors and capacitors by providing equivalent reactances at specific lengths, enabling transmission-line-only circuit designs that perform better at high frequencies.

📌 Key points (3–5)

  • Core capability: Stubs produce purely imaginary input impedance that mimics inductors (positive reactance) or capacitors (negative reactance) depending on length.
  • Length-to-reactance mapping: Short-circuited lines shorter than λ/4 behave like inductors; lengths between λ/4 and λ/2 behave like capacitors.
  • Practical advantage: Transmission lines avoid performance limitations of discrete components at high frequencies and cost less.
  • Common confusion: The reactance matches the discrete component only at the design frequency—frequency response differs, which matters for bandwidth-sensitive applications.
  • Trade-off: Stubs are physically much larger than the discrete devices they replace.

🔌 How stubs emulate reactive components

🔌 Input impedance characteristics

From Section 3.16, the excerpt establishes two key formulas:

  • Short-circuited stub: Input impedance = +jZ₀ tan(βl)
  • Open-circuited stub: Input impedance = -jZ₀ cot(βl)

Both produce completely imaginary-valued impedance (no real part), meaning they dissipate no power—exactly like ideal inductors and capacitors.

📏 Length determines reactance type

The excerpt explains the mapping for short-circuited lines:

Length rangeReactance signBehaves like
0 < l < λ/4Positive (inductive)Inductor
λ/4 < l < λ/2Negative (capacitive)Capacitor
  • Why this works: The tangent function in the impedance formula changes sign at λ/4, flipping from positive to negative reactance.
  • Periodicity: The pattern repeats every λ/2 due to standing wave behavior.

Example: To replace a specific inductor, choose a short-circuited stub length less than λ/4 that produces the same positive reactance at the operating frequency.

🔄 Open-circuited lines work similarly

Open-circuited stubs use the cotangent function, which has opposite behavior but provides the same flexibility to emulate either component type by adjusting length.

🛠️ Practical applications and design

🛠️ Where stubs are used

The excerpt lists three main applications:

  1. Impedance matching circuits (referenced in Section 3.23)
  2. Filters
  3. Other devices requiring reactive elements

All can be "implemented entirely from transmission lines, with fewer or no discrete inductors or capacitors required."

⚡ High-frequency advantages

Transmission lines do not suffer the performance limitations of discrete devices at high frequencies and are less expensive.

  • Discrete inductors and capacitors degrade at high frequencies due to parasitic effects.
  • Transmission line stubs maintain predictable behavior into UHF and higher bands.
  • Cost savings come from eliminating specialized discrete components.

📐 Design procedure (from Example 3.4)

The excerpt provides a worked example for emitter induction in a low-noise amplifier:

Goal: Replace a 2.2 nH inductor at 6 GHz using 50 Ω microstrip with phase velocity 0.6c.

Steps:

  1. Calculate target impedance: +jωL = +j(2πfL) = +j82.9 Ω
  2. Set equal to stub formula: +jZ₀ tan(βl) = +j82.9 Ω
  3. Solve for electrical length: βl ≈ 1.028 rad
  4. Calculate phase constant: β = 2πf/(0.6c) ≈ 209.4 rad/m
  5. Find physical length: l = (βl)/β ≈ 4.9 mm

This demonstrates the practical calculation flow from component specification to physical stub dimensions.

⚠️ Limitations and trade-offs

⚠️ Frequency response mismatch

The excerpt warns:

The variation of reactance with respect to frequency will not be identical, which may or may not be a concern depending on the bandwidth and frequency response requirements of the application.

  • What this means: A discrete inductor's reactance increases linearly with frequency (ωL), but a stub's reactance follows tan(βl), where β itself depends on frequency.
  • When it matters: Wideband circuits or applications requiring specific frequency roll-off characteristics may not tolerate this difference.
  • Don't confuse: Matching at one frequency ≠ matching across all frequencies.

📦 Size disadvantage

A drawback of transmission line stubs in this application is that the lines are typically much larger than the discrete devices they are intended to replace.

  • Physical length relates to wavelength, which can be centimeters or more at lower microwave frequencies.
  • Discrete components are often sub-millimeter in size.
  • Trade-off: Better high-frequency performance vs. larger circuit footprint.

🔬 Measurement technique preview

🔬 Characterizing unknown transmission lines

The excerpt briefly introduces Section 3.18's measurement method, which uses stub theory in reverse:

Measurements needed:

  • Input impedance with short-circuit termination: Z^(SC)_in
  • Input impedance with open-circuit termination: Z^(OC)_in

What can be determined:

  1. Characteristic impedance: Z₀ = √(Z^(SC)_in · Z^(OC)_in)
  2. Electrical length: tan(βl) = √(-Z^(SC)_in / Z^(OC)_in)
  3. Phase velocity: v_p = ω/β (after finding β from βl and known length)

This demonstrates how the stub impedance formulas enable practical transmission line characterization without direct access to internal properties.

33

Measurement of Transmission Line Characteristics

3.18 Measurement of Transmission Line Characteristics

🧭 Overview

🧠 One-sentence thesis

A simple two-measurement technique using short-circuit and open-circuit input impedances allows determination of a lossless transmission line's characteristic impedance, electrical length, and phase velocity.

📌 Key points (3–5)

  • What the technique measures: characteristic impedance Z₀, electrical length βl, and phase velocity vₚ of a lossless transmission line.
  • How it works: measure input impedance when the line is short-circuited and when it is open-circuited, then combine the results mathematically.
  • Key formulas: multiply the two measurements to get Z₀²; divide them to get tan²(βl).
  • Common confusion: if the line is longer than λ/2, the tangent function's periodicity creates ambiguity—you may not be able to determine βl unambiguously from a single frequency.
  • Why it matters: once you know βl and the physical length l, you can calculate β and then phase velocity vₚ = ω/β.

📏 The two required measurements

📏 Short-circuit input impedance

Z⁽ˢᶜ⁾ᵢₙ = +jZ₀ tan(βl): the input impedance when the transmission line is terminated in a short circuit.

  • This result comes from Section 3.16 (referenced in the excerpt).
  • The impedance is purely imaginary (reactive) and depends on both the characteristic impedance Z₀ and the electrical length βl.

📏 Open-circuit input impedance

Z⁽ᴼᶜ⁾ᵢₙ = −jZ₀ cot(βl): the input impedance when the transmission line is terminated in an open circuit.

  • Also derived in Section 3.16.
  • Again purely imaginary, but with a negative sign and using cotangent instead of tangent.
  • Example: if you measure these two impedances at the input of an unknown line, you can extract the line's properties without needing to know Z₀ or βl in advance.

🔢 Extracting characteristic impedance

🔢 Multiplying the measurements

The excerpt shows that when you multiply the short-circuit and open-circuit input impedances:

Z⁽ˢᶜ⁾ᵢₙ · Z⁽ᴼᶜ⁾ᵢₙ = Z₀²

  • The product is simply the square of the characteristic impedance.
  • The imaginary units and the tangent/cotangent terms cancel out.

🔢 Formula for Z₀

From the product above:

Z₀ = √(Z⁽ˢᶜ⁾ᵢₙ · Z⁽ᴼᶜ⁾ᵢₙ)

  • This is Equation 3.121 in the excerpt.
  • You take the square root of the product of your two measurements.
  • Example: if Z⁽ˢᶜ⁾ᵢₙ = +j100 Ω and Z⁽ᴼᶜ⁾ᵢₙ = −j25 Ω, then Z₀² = (+j100)(−j25) = 2500, so Z₀ = 50 Ω.

🔢 Extracting electrical length

🔢 Dividing the measurements

If you instead divide the short-circuit impedance by the open-circuit impedance:

Z⁽ˢᶜ⁾ᵢₙ / Z⁽ᴼᶜ⁾ᵢₙ = −tan²(βl)

  • The characteristic impedance Z₀ cancels out.
  • You are left with the negative square of the tangent of the electrical length.

🔢 Formula for βl

Rearranging:

tan(βl) = [−Z⁽ˢᶜ⁾ᵢₙ / Z⁽ᴼᶜ⁾ᵢₙ]^(1/2)

  • This is Equation 3.122 in the excerpt.
  • Take the square root of the negative ratio of the two measurements.
  • Then take the inverse tangent (arctan) to find βl.

⚠️ Length ambiguity

Don't confuse: the tangent function repeats every π radians (half-wavelength).

  • If you know in advance that the line length l is less than λ/2, you can unambiguously determine βl by taking the inverse tangent.
  • If l is unknown or longer than λ/2, the periodicity of tangent creates ambiguity—multiple values of βl can give the same tangent.
  • The excerpt notes that making multiple measurements over a range of frequencies can resolve this ambiguity, but the method is not presented.

🚀 Determining phase velocity

🚀 From βl to β

Once you have determined the electrical length βl and you know the physical length l:

β = (βl) / l

  • β is the phase propagation constant (radians per meter).
  • Example: if βl ≈ 1.028 rad and l = 4.9 mm, then β ≈ 1.028 / 0.0049 ≈ 209.4 rad/m (as shown in the earlier example in the excerpt).

🚀 From β to vₚ

Once you know β, you can calculate the phase velocity:

vₚ = ω / β

  • ω is the angular frequency (2πf).
  • This follows from the definition of β in Section 3.8 (referenced in the excerpt).
  • Example: if you measure at a known frequency f, calculate ω = 2πf, then divide by β to get vₚ.

🚀 Summary workflow

StepWhat you doWhat you get
1. MeasureZ⁽ˢᶜ⁾ᵢₙ and Z⁽ᴼᶜ⁾ᵢₙTwo complex impedances
2. Multiply√(Z⁽ˢᶜ⁾ᵢₙ · Z⁽ᴼᶜ⁾ᵢₙ)Characteristic impedance Z₀
3. Divide[−Z⁽ˢᶜ⁾ᵢₙ / Z⁽ᴼᶜ⁾ᵢₙ]^(1/2) then arctanElectrical length βl
4. Calculate(βl) / lPhase constant β
5. Calculateω / βPhase velocity vₚ
34

Quarter-Wavelength Transmission Line

3.19 Quarter-Wavelength Transmission Line

🧭 Overview

🧠 One-sentence thesis

A quarter-wavelength transmission line transforms impedance inversely (input impedance equals characteristic impedance squared divided by load impedance), making it a powerful tool for impedance matching and RF decoupling despite its relatively large physical size.

📌 Key points (3–5)

  • Core property: At length λ/4, the input impedance is Z₀²/Z_L—inversely proportional to the load impedance.
  • Impedance matching: Quarter-wave lines can match any two real-valued impedances by choosing Z₀ = √(Z_in · Z_L).
  • Complex loads: When the load has significant imaginary component, a two-stage approach is needed: first eliminate the imaginary part with a line of length l₁ < λ/4, then apply quarter-wave matching.
  • Common confusion: Quarter-wave matching is effective but often results in structures at least λ/4 long, which can be large compared to associated electronics—other techniques like single stub matching may yield smaller structures.
  • Beyond matching: Quarter-wave lines also serve as impedance inverters, transforming low impedances to high (useful for RF/DC decoupling in amplifiers).

🔢 Derivation of the quarter-wave property

🔢 Starting from the general input impedance formula

The general expression for input impedance of a lossless transmission line is:
Z_in(l) = Z₀ · [1 + Γ·e^(−j2βl)] / [1 − Γ·e^(−j2βl)]

where:

  • Z₀ is the characteristic impedance
  • Γ is the reflection coefficient
  • β is the phase propagation constant
  • l is the line length

📐 What happens at λ/4

When l = λ/4:

  • 2βl = 2 · (2π/λ) · (λ/4) = π
  • Therefore e^(−jπ) = −1
  • Substituting: Z_in(λ/4) = Z₀ · (1 − Γ) / (1 + Γ)

🔄 Substituting the reflection coefficient

The reflection coefficient is: Γ = (Z_L − Z₀) / (Z_L + Z₀)

Substituting this into the input impedance expression and simplifying (by multiplying numerator and denominator by Z_L + Z₀):

  • Numerator becomes: (Z_L + Z₀) − (Z_L − Z₀) = 2Z₀
  • Denominator becomes: (Z_L + Z₀) + (Z_L − Z₀) = 2Z_L
  • Result: Z_in(λ/4) = Z₀² / Z_L

Key insight: The input impedance is inversely proportional to the load impedance.

🔁 Why it's called an impedance inverter

A transmission line of length λ/4 is sometimes referred to as a quarter-wave inverter or simply as an impedance inverter.

  • Small impedances transform into large impedances
  • Large impedances transform into small impedances
  • This inversion property is the foundation for many applications

🎯 Impedance matching with quarter-wave lines

🎯 The matching formula

Solving Z_in(λ/4) = Z₀²/Z_L for Z₀:

  • Z₀ = √[Z_in(λ/4) · Z_L]

This means:

  • To match load Z_L to source impedance Z_in(λ/4)
  • Choose characteristic impedance equal to the geometric mean of the two impedances
  • Set the line length to λ/4
ParameterRoleHow to determine
Z₀Characteristic impedance√(source impedance · load impedance)
lLine lengthλ/4 (one-quarter wavelength in the line)

📝 Example: 300 Ω to 50 Ω match at 10 GHz

Given: Match 300 Ω to 50 Ω at 10 GHz using microstrip line where propagation wavelength is 60% of free space.

Solution steps:

  1. Free-space wavelength λ₀ = c/f ≈ 3 cm at 10 GHz
  2. Wavelength in line: λ = 0.6λ₀ ≈ 1.8 cm
  3. Line length: l = λ/4 ≈ 4.5 mm
  4. Characteristic impedance: Z₀ = √(300 Ω · 50 Ω) ≈ 122.5 Ω

This Z₀ value would then determine the physical width of the microstrip line.

⚠️ Important limitation

For real-valued Z₀: The product of source and load impedances must be real-valued.

Not suitable when: Z_L has a significant imaginary component and matching to a real-valued source is desired.

🔧 Two-stage matching for complex loads

🔧 The problem with complex loads

When the load impedance has a significant imaginary component (e.g., Z_L = R + jX where X is large), the simple quarter-wave formula fails because it would require a complex characteristic impedance (which is not physically realizable).

🔧 The two-stage solution

Stage 1: Transform the complex load to a real-valued impedance

  • Use a transmission line of length l₁
  • Characteristic impedance Z₀₁ is not critical (choose for convenience)
  • Find l₁ using the general input impedance equation (numerical search or Smith chart)
  • l₁ will always be less than λ/4

Why l₁ < λ/4: Z_in(l₁) is periodic with period λ/2, and there are two sign changes in the imaginary component as l₁ increases from 0 to λ/2, so the first zero crossing occurs before λ/4.

Stage 2: Apply quarter-wave matching

  • Now that the impedance is real-valued, use the standard quarter-wave technique
  • Length l₂ = λ/4
  • Z₀₂ = √(real impedance from stage 1 · source impedance)

📝 Example: Matching a patch antenna (35 + j35 Ω) to 50 Ω

Given: Patch antenna with Z_A = 35 + j35 Ω; use Z₀₁ = 50 Ω for the first section.

Stage 1 solution:

  • Reflection coefficient: Γ = (Z_A − Z₀₁)/(Z_A + Z₀₁) ≈ −0.0059 + j0.4142
  • Find smallest β₁l₁ where imaginary part of Z₁(l₁) = 0
  • Result: β₁l₁ = 0.793 rad gives Z₁ ≈ 120.719 − j0.111 Ω (close enough to real)
  • Since β₁ = 2π/λ: l₁ = (β₁l₁)/(2π) · λ ≈ 0.126λ

Stage 2 solution:

  • Length: l₂ = 0.25λ
  • Characteristic impedance: Z₀₂ = √(120.719 Ω · 50 Ω) ≈ 77.7 Ω

Total length: l₁ + l₂ ≈ 0.376λ

⚠️ Don't confuse: Size considerations

  • A patch antenna typically has sides of length ≈ λ/2
  • The matching structure (0.376λ) is nearly as large as the antenna itself
  • At UHF frequencies (300–3000 MHz), this can be problematic
  • Alternative: Single stub matching (Section 3.23) typically results in significantly smaller structures
  • Trade-off: Quarter-wave matching is effective and commonly used, but the λ/4 minimum length is often large compared to associated electronics

🔌 Beyond matching: RF/DC decoupling application

🔌 The problem in transistor amplifiers

Scenario: Transistor amplifiers for RF often receive DC current at the same terminal that delivers the amplified RF signal.

Issue:

  • Power supply typically has low output impedance
  • If connected directly, RF flows predominantly toward the power supply (low impedance path)
  • Desired RF path has higher impedance, so signal is lost

🔌 Traditional solution: series inductor

  • Place an inductor in series with the power supply output
  • Inductor exhibits low impedance at DC (allows DC power through)
  • Inductor exhibits high impedance at RF (blocks RF from entering power supply)

Problem with discrete inductors at high RF frequencies:

  • Practical inductors also have parallel capacitance
  • Capacitance tends to decrease impedance at high frequencies
  • This defeats the purpose

🔌 Quarter-wave solution

Replace the inductor with a transmission line of length λ/4:

At DC:

  • Wavelength at DC is infinite
  • The transmission line is essentially transparent to the power supply
  • DC current flows freely

At RF frequencies:

  • The line transforms the low impedance of the power supply to very high impedance
  • Using Z_in(λ/4) = Z₀²/Z_L: low Z_L → high Z_in
  • RF signal sees high impedance toward power supply, so flows along the desired path (moderate impedance)

Practical advantages:

  • Transmission lines on printed circuit boards are much cheaper than discrete inductors
  • Always in stock (no supply chain issues)
  • No parasitic capacitance problems at high frequencies

🔁 How impedance inversion helps

PathImpedanceResult
Power supply (direct)LowWould attract RF (bad)
Power supply (through λ/4 line)High (inverted)Blocks RF (good)
Desired RF output pathModerateRF flows here (good)

The quarter-wave line effectively "hides" the low-impedance power supply from the RF signal by presenting a high impedance at the connection point.

Budget: 1000000 Used: 292869 Remaining: 707131

35

Power Flow on Transmission Lines

3.20 Power Flow on Transmission Lines

🧭 Overview

🧠 One-sentence thesis

Power flow on transmission lines can be calculated using time-average formulas, and the fraction of power delivered to a load versus reflected depends on the voltage reflection coefficient and impedance mismatch.

📌 Key points (3–5)

  • Time-average power formula: Power associated with a sinusoidal wave on a lossless transmission line is calculated by integrating voltage times current over one period.
  • Incident vs reflected power: Reflected power equals the square of the reflection coefficient magnitude times the incident power.
  • Power delivered to load: The load receives the difference between incident and reflected power, expressed as (1 - |Γ|²) times incident power.
  • Common confusion: Position dependence disappears after integration—power of a traveling wave is constant along the line, not varying with z.
  • Practical implication: Even moderate mismatches (e.g., 50 Ω to 75 Ω) may be acceptable, reflecting only 4% of power while delivering 96%.

⚡ Time-average power calculation

⚡ Definition and setup

Time-average power: P_av(z) = (1/T) times the integral from t₀ to t₀+T of v(z,t) times i(z,t) dt, where T = 2π/f is one period.

  • The excerpt considers a lossless transmission line oriented along the z axis.
  • Because time-average power of a sinusoidal signal does not change with time, the start time t₀ can be set to zero without loss of generality.
  • The calculation focuses on a wave incident from z < 0 on a load impedance Z_L at z = 0.

📐 Voltage and current expressions

For an incident wave:

  • Potential: v⁺(z,t) = |V⁺₀| cos(ωt - βz + φ)
  • Current: i⁺(z,t) = |V⁺₀|/Z₀ cos(ωt - βz + φ)
  • Both have the same phase because they represent a single traveling wave.

🧮 Integration using trigonometric identity

The excerpt employs the identity:

  • cos²θ = 1/2 + (1/2)cos(2θ)
  • When integrated over one period T, the cosine term integrates to zero (it covers two complete periods).
  • Only the constant 1/2 term contributes, yielding T/2.
  • Key result: Position dependence on z is eliminated—the power is the same for all z < 0, as expected for a traveling wave.

🎯 Incident power result

The time-average power associated with the incident wave is:

  • P⁺_av = |V⁺₀|² / (2Z₀)
  • This formula applies at any point z < 0 along the line.
  • The excerpt emphasizes this is "the time-average power associated with a wave traveling in a single direction along a lossless transmission line."

🔄 Reflected and delivered power

🔄 Reflected power

Using the same procedure, the power associated with the reflected wave is:

  • P⁻_av = |ΓV⁺₀|² / (2Z₀) = |Γ|² times |V⁺₀|² / (2Z₀)
  • Simplified: P⁻_av = |Γ|² P⁺_av
  • The excerpt states this "gives the time-average power associated with the wave reflected from an impedance mismatch."
  • Only the magnitude of Γ matters for power; the phase does not affect the squared magnitude.

📦 Power delivered to the load

The excerpt uses the principle of conservation of power:

  • Power incident on the load must equal power reflected plus power absorbed.
  • Formula: P⁺_av = P⁻_av + P_L
  • Rearranging: P_L = (1 - |Γ|²) P⁺_av
  • This is "the time-average power transferred to a load impedance, and is equal to the difference between the powers of the incident and reflected waves."

🔍 Don't confuse: power vs voltage

  • The reflection coefficient Γ relates voltages linearly.
  • But power depends on the square of the reflection coefficient magnitude: |Γ|².
  • Example: if |Γ| = 0.2, reflected voltage is 20% but reflected power is only 4%.

📊 Practical example: 50 Ω to 75 Ω mismatch

📊 The scenario

The excerpt asks: "How important is it to match 50 Ω to 75 Ω?"

  • Two common impedances in radio engineering: 50 Ω and 75 Ω.
  • Often necessary to connect a 50 Ω transmission line to a 75 Ω device (or vice-versa).
  • Question: if no matching is used, what fraction of power is delivered vs reflected, and what is the SWR?

🧮 Calculation details

Going from 50 Ω transmission line to 75 Ω load:

  • Voltage reflection coefficient: Γ = (75 - 50)/(75 + 50) = +0.2
  • Fraction of power reflected: |Γ|² = 0.04 = 4%
  • Fraction of power transmitted: 1 - |Γ|² = 0.96 = 96%
  • SWR = (1 + |Γ|)/(1 - |Γ|) = 1.5

Going from 75 Ω to 50 Ω:

  • Only the sign of Γ changes (becomes -0.2).
  • Fractions of reflected and transmitted power remain 4% and 96%, respectively.
  • SWR is still 1.5.

✅ Practical conclusion

The excerpt states:

  • "This is often acceptable, but may not be good enough in some particular applications."
  • "Suffice it to say that it is not necessarily required to use an impedance matching device to connect 50 Ω to 75 Ω devices."
  • The small reflection loss (4%) means moderate mismatches can sometimes be tolerated without additional matching circuitry.
Parameter50 Ω → 75 Ω75 Ω → 50 Ω
Γ+0.2-0.2
Power reflected4%4%
Power delivered96%96%
SWR1.51.5
36

3.21 Impedance Matching: General Considerations

3.21 Impedance Matching: General Considerations

🧭 Overview

🧠 One-sentence thesis

Impedance matching transforms a given load impedance into a desired input impedance because practical systems cannot conveniently operate all devices at the same impedance, and transmission-line-based solutions become especially useful at high frequencies where discrete components behave non-ideally.

📌 Key points (3–5)

  • What impedance matching does: transforms a particular impedance Z_L into a modified impedance Z_in.
  • Why matching is needed: different devices (cables, antennas, amplifiers) naturally operate at different impedances, and it is impractical to manufacture cables for every possible termination impedance.
  • Discrete-component limitations at high frequencies: resistors, capacitors, and inductors behave as combinations of ideal components (e.g., resistors act like resistors plus inductors), making analysis difficult.
  • Common confusion: impedance matching does not always mean maximizing power transfer—some designs intentionally mismatch to meet other goals (e.g., transistor amplifier design).
  • Transmission-line solution: replacing discrete components with transmission-line structures (e.g., microstrip lines) is convenient at UHF and higher frequencies where wavelengths are small.

🎯 Why impedance matching is necessary

🎯 Practical constraints in real systems

"Impedance matching" refers to the problem of transforming a particular impedance Z_L into a modified impedance Z_in.

The excerpt lists several reasons why all devices in a system cannot operate at the same impedance:

  • Coaxial cables: It is not convenient or practical to manufacture cables with characteristic impedance equal to every possible terminating impedance.
  • Antennas: Different antenna types operate at different impedances, and antenna impedance varies significantly with frequency.
  • Amplifiers: Different amplifier types operate most effectively at different output impedances.
    • Current-source amplifiers work best with low output impedance.
    • Voltage-source amplifiers work best with high output impedance.

🔄 Intentional mismatching

  • Some design techniques (e.g., transistor amplifier design) rely on intentionally mismatching impedances.
  • This means matching to an impedance different from the one that maximizes power transfer or minimizes reflection.
  • Various design goals are met by applying particular impedances to the input and output ports of the transistor.
  • Don't confuse: impedance matching is not always about maximizing power transfer; it can serve other design objectives.

⚡ Problems with discrete components at high frequencies

⚡ Non-ideal behavior of real components

The excerpt explains that discrete components (resistors, capacitors, inductors) do not behave ideally at high frequencies:

ComponentActual behavior
ResistorBehaves as ideal resistor in series with ideal inductor
CapacitorBehaves as ideal capacitor in series with ideal resistor
InductorBehaves as ideal inductor in parallel with ideal capacitor, and in series with ideal resistor
  • These parasitic effects make the use of discrete components increasingly difficult with increasing frequency.

🛠️ Two possible solutions

  1. More precise modeling: Model each component more accurately and account for non-ideal behavior in analysis and design.
  2. Replace with transmission-line devices: Replace troublesome discrete components—or all discrete components—with transmission-line structures.

📡 Transmission-line-based matching

📡 Why transmission lines are convenient

  • Transmission-line structures are particularly convenient in circuits implemented on printed circuit boards at UHF frequencies and higher.
  • Reason 1: Transmission lines (e.g., microstrip lines) are easy to implement on PCBs.
  • Reason 2: At UHF and higher, wavelengths are relatively small, so transmission-line structures are relatively compact.
  • Applications using transmission lines as impedance-matching components can also be found at lower frequencies.

📡 Context: acceptable mismatch

The excerpt begins with a numerical example (not the main focus):

  • Going from a 50 Ω transmission line to a 75 Ω termination results in:
    • 4% reflected power
    • 96% transmitted power
    • Standing wave ratio (SWR) = 1.5
  • This level of mismatch is often acceptable, but may not be good enough in some applications.
  • Implication: It is not necessarily required to use an impedance matching device to connect 50 Ω to 75 Ω devices, but matching may still be desirable depending on the application.

🔧 Single-reactance matching strategy

🔧 Overview of the approach

The excerpt introduces a specific matching technique that uses a section of transmission line combined with a discrete reactance (capacitor or inductor).

Strategy:

  1. Use the transmission line to transform the real part of the load impedance (or admittance) to the desired value.
  2. Use the reactance to modify the imaginary part to the desired value.

🔧 Difference from quarter-wave technique

  • Don't confuse: This approach differs from the quarter-wave technique (Section 3.19).
  • In the quarter-wave approach, the first transmission line is used to zero the imaginary part.
  • In single-reactance matching, the transmission line adjusts the real part first.

🔧 Series reactance version

The excerpt describes one version in detail (Figure 3.28):

Step 1: Match the real part

  • The transmission line transforms load impedance Z_L into a new impedance Z_1 such that the real part of Z_1 equals the real part of Z_in.
  • This is done by solving an equation (3.146) for the transmission line length l, using numerical search or a Smith chart.
  • The characteristic impedance Z_0 and phase propagation constant β of the transmission line are independent variables and can be selected for convenience.
  • Normally, the smallest value of l that satisfies the equation is desired; this will be ≤ λ/4 (one-quarter wavelength) because the real part of Z_1 is periodic in l with period λ/4.

Step 2: Match the imaginary part

  • After matching the real component, the imaginary component of Z_1 may be transformed to the desired value (imaginary part of Z_in) by attaching a reactance X_s in series with the transmission line input.
  • The result is: Z_in = Z_1 + j X_s.
  • Choose X_s = imaginary part of (Z_in − Z_1).
  • The sign of X_s determines the component type:
    • X_s < 0 → capacitor
    • X_s > 0 → inductor
  • The value of the component is determined from X_s and the design frequency.

🔧 Example scenario

The excerpt includes an example (3.9):

  • Goal: Match a source impedance of 50 Ω to a load impedance of 33.9 + j17.6 Ω at 1.5 GHz.
  • Given: Transmission line with characteristic impedance 50 Ω and phase velocity 0.6c.
  • (The full solution is not provided in the excerpt.)
37

Single-Reactance Matching

3.22 Single-Reactance Matching

🧭 Overview

🧠 One-sentence thesis

Single-reactance matching uses a transmission line to transform the real part of impedance (or admittance) to the desired value, then adds a single discrete or stub reactance to cancel the imaginary part, offering a practical alternative to discrete components at high frequencies.

📌 Key points (3–5)

  • Core strategy: transmission line adjusts the real part; a single reactance (capacitor or inductor) cancels the imaginary part.
  • Two versions: series reactance (works with impedance) vs. parallel reactance (works with admittance); parallel is more common because most transmission lines share a ground reference.
  • Discrete vs. stub reactance: discrete components (capacitors/inductors) can be replaced by open- or short-circuited transmission line stubs to avoid non-ideal behavior or component availability issues at high frequencies.
  • Common confusion: series vs. parallel—series requires both terminals floating, but most transmission lines have one conductor grounded, making parallel attachment more practical.
  • Design trade-offs: parallel reactance often requires shorter transmission line length but smaller (harder-to-achieve) component values; stub matching avoids discrete components entirely but requires careful choice of stub termination.

🔧 Series reactance approach

🔧 How it works

The transmission line transforms the load impedance Z_L into an intermediate impedance Z_1 such that the real part of Z_1 equals the real part of the desired input impedance Z_in.

  • The real part is matched by solving:
    Re{Z_1} = Re{Z_0 × (1 + Γ exp(−j2βl)) / (1 − Γ exp(−j2βl))}
    for the transmission line length l, where Γ is the reflection coefficient and β is the phase propagation constant.
  • The smallest l satisfying this equation is chosen; it will be ≤ λ/4 because the real part is periodic with period λ/4.

⚡ Adding the series reactance

After the transmission line, a reactance X_s is placed in series to cancel the imaginary part:

  • X_s = Im{Z_in − Z_1}
  • If X_s < 0, use a capacitor; if X_s > 0, use an inductor.
  • The component value is determined from X_s and the design frequency.

Example: Matching 50 Ω source to 33.9 + j17.6 Ω load at 1.5 GHz with a 50 Ω transmission line (phase velocity 0.6c) requires a 7.8 mm line and a 3.7 pF series capacitor.

🔀 Parallel reactance approach

🔀 Why use admittance

When the reactance is attached in parallel, it is easier to work with admittance (Y = 1/Z) because parallel admittances simply add:

  • Y_in = Y_1 + jB_p
  • Y_1 is the admittance after the transmission line; B_p is the parallel susceptance (imaginary part of admittance).

🔀 Matching procedure

The transmission line transforms the load admittance Y_L into Y_1 such that Re{Y_1} = Re{Y_in}.

  • Solve:
    Re{Y_1} = Re{Y_0 × (1 − Γ exp(−j2βl)) / (1 + Γ exp(−j2βl))}
    for the smallest l, where Y_0 = 1/Z_0 is the characteristic admittance.
  • Then attach a parallel susceptance B_p = Im{Y_in − Y_1}.
  • If B_p > 0, use a capacitor; if B_p < 0, use an inductor.

Example: Same problem as above, but with parallel reactance: requires only a 2.4 mm line and a 1.2 pF parallel capacitor—much shorter line but smaller capacitor value.

🔀 Series vs. parallel: practical distinction

AspectSeries reactanceParallel reactance
Impedance/admittanceWorks with impedance ZWorks with admittance Y
Ground referenceRequires both terminals floatingBoth stub and main line share ground
PracticalityLess convenient for grounded transmission linesMore convenient; most common
Trade-offLonger line, larger componentShorter line, smaller component

Don't confuse: The choice is not about performance but about convenience—parallel is preferred because most transmission lines (coax, microstrip) have one conductor grounded.

🛠️ Single-stub matching

🛠️ Replacing discrete reactance with a stub

A stub is a transmission line section that is either open-circuited or short-circuited at one end, acting as a pure reactance.

Stub: a transmission line terminated in an open or short circuit, used to replace a discrete capacitor or inductor.

  • Stubs avoid problems with discrete components at high frequencies: non-ideal behavior, non-standard values, cost, and availability.
  • The stub is attached in parallel (series is impractical for the same grounding reasons as discrete reactances).

🛠️ Stub susceptance formulas

The susceptance provided by a stub depends on its termination and length l_2:

  • Short-circuited stub: B_p = −Y_02 cot(β_2 l_2)
  • Open-circuited stub: B_p = +Y_02 tan(β_2 l_2)

where Y_02 = 1/Z_02 is the characteristic admittance of the stub, and β_2 is its phase constant.

🛠️ Choosing open vs. short termination

  • Default rule: choose the termination that yields the shortest stub.
  • Other considerations: if DC is present on the line, avoid short-circuit termination (it would short the DC).

Example: Same matching problem, using stubs with 50 Ω characteristic impedance throughout:

  • Primary line length: 0.020λ
  • Open-circuited stub: 0.084λ
  • Short-circuited stub: 0.334λ → choose open-circuited (much shorter).

🛠️ Microstrip implementation

In microstrip (common at UHF and above):

  • The top trace is one conductor; the ground plane underneath is the other.
  • Open-circuit stub: the end of the stub is not connected to the ground plane.
  • Short-circuit stub: the end is connected to the ground plane via a via (a plated-through hole).

Don't confuse: Open vs. short termination is a design choice based on stub length and DC considerations, not a fundamental difference in matching capability.

📐 Design workflow summary

📐 Step-by-step for parallel reactance/stub matching

  1. Calculate reflection coefficient Γ from the load impedance and transmission line characteristic impedance.
  2. Solve for transmission line length l_1 such that Re{Y_1} = Re{Y_in}, using numerical search or Smith chart; choose the smallest l_1.
  3. Compute the required susceptance B_p = Im{Y_in − Y_1} to cancel the imaginary part.
  4. For discrete reactance: determine if capacitor (B_p > 0) or inductor (B_p < 0); calculate component value from B_p and frequency.
  5. For stub: choose open or short termination; solve for stub length l_2 from the susceptance formula; pick the shorter option.

📐 Key independent variables

  • Characteristic impedance Z_0 (or Y_0) of the main transmission line: chosen for convenience, often 50 Ω.
  • Phase propagation constant β: determined by the transmission line's phase velocity and frequency.
  • Stub characteristic impedance Z_02: also chosen for convenience, often matched to the main line.

Don't confuse: The transmission line length is not arbitrary—it must satisfy the real-part matching equation; only the characteristic impedance is a free design choice.

38

Single-Stub Matching

3.23 Single-Stub Matching

🧭 Overview

🧠 One-sentence thesis

Single-stub matching replaces discrete reactances with open- or short-circuited transmission line stubs to achieve impedance matching, avoiding the cost and practical limitations of discrete components.

📌 Key points (3–5)

  • Why use stubs: discrete reactances may not be standard values, may behave non-ideally at the desired frequency, or may add cost and logistical issues.
  • What a stub is: a transmission line that has been open- or short-circuited to replace a discrete reactance.
  • Parallel vs series attachment: parallel attachment is preferred because most transmission lines use one conductor as ground, making parallel stubs easier to implement than series configurations.
  • Common confusion: open- vs short-circuited stubs—choose the termination that yields the shortest stub, unless DC is present (in which case avoid short circuits).
  • Practical implementation: widely used in microstrip lines at UHF frequencies (300–3000 MHz) and above.

🔧 Why replace discrete reactances with stubs

🔧 Practical limitations of discrete components

The excerpt identifies three main reasons to avoid discrete reactances:

  • Non-standard values: the required reactance may not be a standard component value.
  • Non-ideal behavior: discrete inductors and capacitors may not behave ideally at the desired frequency (Section 3.21 discusses this further).
  • Cost and logistics: additional components add cost and complexity.

📡 What a stub does

A stub: a transmission line which has been open- or short-circuited.

  • Section 3.16 (referenced in the excerpt) explains how a stub can replace a discrete reactance.
  • The stub provides the same reactive effect as a discrete inductor or capacitor, but using only transmission line geometry.
  • Example: Figure 3.30 shows a microstrip implementation where the stub is open-circuited.

🔌 Parallel vs series stub attachment

🔌 Why parallel attachment is preferred

The excerpt states that single-stub matching "is usually implemented using the parallel reactance approach."

Reason: most transmission lines use one conductor as a local ground reference.

  • Microstrip: the ground plane of the printed circuit board is tied to ground.
  • Coaxial cable: the outer conductor (shield) is usually tied to ground.

The problem with series attachment: a discrete reactance (capacitor or inductor) does not require either terminal to be tied to ground, but a series stub would interrupt the main line in a way that is inconvenient given the grounded-conductor structure.

The advantage of parallel attachment: both the stub and the main transmission line have one terminal at ground, so they can share the ground reference naturally.

📐 Susceptance formulas for stubs

The excerpt provides formulas for the susceptance (B_p) of stubs, expressed in terms of admittance:

Stub terminationSusceptance formulaNotes
Short-circuitedB_p = −Y_02 cot(β_2 l_2)Y_02 is the characteristic admittance; β_2 is the phase constant; l_2 is the stub length
Open-circuitedB_p = +Y_02 tan(β_2 l_2)Same variables as above
  • These are derived from Equations 3.117 and 3.119 (input impedance of open- and short-circuited stubs from Section 3.16).
  • Admittances are used because parallel reactance matching is most easily done in admittance terms.
  • The characteristic impedance Z_02 = 1/Y_02 is an independent variable chosen for convenience.

🔀 Choosing open- vs short-circuited termination

🔀 Default rule: shortest stub wins

Given no other basis for selection, the termination that yields the shortest stub is chosen.

  • Both open- and short-circuited stubs can provide the required susceptance, but at different lengths.
  • The excerpt's Example 3.11 illustrates this: the open-circuited stub is approximately 0.084 wavelengths, while the short-circuited stub would be approximately 0.334 wavelengths—much longer.
  • Therefore, the open-circuited stub is chosen.

⚡ Special consideration: DC presence

"Other basis for selection": whether DC might be present on the line.

  • If DC is present with the signal of interest, a short-circuit termination would create a DC short circuit.
  • This would be "certainly a bad idea" without some kind of remediation.
  • In such cases, use an open-circuited stub instead.

Don't confuse: the choice is not about signal quality alone; it also depends on the DC/bias conditions of the circuit.

🛠️ Design procedure and example

🛠️ Single-stub matching procedure

The procedure is essentially the same as the single parallel reactance method (Section 3.22), except:

  1. Use a stub instead of a discrete inductor or capacitor.
  2. Work in admittances (since parallel reactance matching is easier in admittance form).

Steps (from Example 3.11):

  1. Calculate the reflection coefficient Γ from the load impedance Z_L and characteristic impedance Z_0.
  2. Find the length l_1 of the primary line (the main transmission line connecting the two ports) such that the real part of the input admittance Y_1 matches the source admittance.
  3. Calculate the imaginary part of Y_1 after attaching the primary line.
  4. Find the shortest stub length l_2 that provides a susceptance equal and opposite to the imaginary part of Y_1, canceling it out.
  5. Choose open- or short-circuited termination based on which gives the shorter stub (and DC considerations).

📝 Example 3.11 summary

Problem: Match a source impedance of 50 Ω to a load impedance of 33.9 + j17.6 Ω using 50 Ω transmission lines.

Solution:

  • Reflection coefficient Γ ≈ −0.142 + j0.239.
  • Primary line length l_1 ≈ 0.020 wavelengths.
  • After attaching the primary line, input admittance Y_1 ≈ 0.0200 − j0.0116 mho.
  • Need a stub with susceptance +j0.0116 mho to cancel the imaginary part.
  • Open-circuited stub: l_2 ≈ 0.084 wavelengths.
  • Short-circuited stub: l_2 ≈ 0.334 wavelengths (much longer).
  • Result: use an open-circuited stub with length ≈ 0.084 wavelengths, attached in parallel at the source end of the primary line.

🏭 Practical implementation in microstrip

🏭 Microstrip construction

Figure 3.30 shows a practical single-stub match using microstrip transmission line.

Structure:

  • The top (visible) traces comprise one conductor.
  • The ground plane (underneath, not visible) comprises the other conductor.
  • The stub is open-circuited: the end of the stub is not connected to the ground plane.

How to create a short-circuit termination:

  • Connect the end of the stub to the ground plane using a via (a plated-through hole that electrically connects the top and bottom layers).

📶 Frequency range

Single-stub matching is very common for impedance matching using microstrip lines at frequencies in the UHF band (300–3000 MHz) and above.

  • At these frequencies, transmission line stubs are practical in size (wavelengths are short enough that stub lengths are physically manageable).
  • Discrete components may have significant parasitic effects at these frequencies, making stubs more attractive.
39

Vector Arithmetic

4.1 Vector Arithmetic

🧭 Overview

🧠 One-sentence thesis

Vectors are mathematical objects combining magnitude and direction that enable precise description of physical quantities like velocity and force, with operations (addition, dot product, cross product) that reveal relationships between directions and magnitudes in three-dimensional space.

📌 Key points (3–5)

  • What vectors represent: mathematical objects with both scalar magnitude (and possibly phase) and direction, used to describe physical quantities like velocity, force, and electromagnetic fields.
  • Position-fixed vs position-free distinction: position vectors are anchored to specific points in space (e.g., the origin), while position-free vectors (e.g., velocity) describe properties independent of location and are equal if they share magnitude and direction.
  • Cartesian representation: any vector can be expressed as a sum of components along three perpendicular basis vectors (x-hat, y-hat, z-hat), making calculations straightforward.
  • Common confusion: the dot product measures directional similarity (yielding a scalar), while the cross product produces a new vector perpendicular to both operands—these serve fundamentally different purposes.
  • Why arithmetic matters: vector operations (addition, subtraction, dot product, cross product) enable calculation of relative positions, angles between directions, and perpendicular directions essential for physics and engineering.

📐 What vectors are and how to write them

📏 Definition and notation

A vector is a mathematical object that has both a scalar part (i.e., a magnitude and possibly a phase), as well as a direction.

  • Written as A with magnitude A = |A| and direction a-hat (a unit vector).
  • The relationship: A = A a-hat (magnitude times direction).
  • If complex-valued, the magnitude A is also complex-valued.
  • Example: velocity is speed (scalar, m/s) plus direction of movement; force is magnitude (N) plus direction of application.

🧭 Basis vectors in Cartesian coordinates

A basis vector is a position-free unit vector that is perpendicular to all other basis vectors for that coordinate system.

  • The three Cartesian basis vectors: x-hat, y-hat, z-hat.
  • Each points in the direction where its coordinate increases most rapidly.
  • All are mutually perpendicular and have magnitude equal to one.
  • Any vector A can be written: A = x-hat A_x + y-hat A_y + z-hat A_z.

📊 Magnitude and unit vectors

  • Magnitude calculated from components: |A| = square root of (A_x squared + A_y squared + A_z squared).
  • The associated unit vector: a-hat = A / |A|.
  • This gives the direction of A with magnitude normalized to one.

🔄 Position-fixed vs position-free vectors

📍 Position vectors (position-fixed)

  • Describe a location in space as distance and direction from the origin (or another reference point).
  • "Position-fixed" means defined with respect to a specific point.
  • Two position vectors are equal only if they have the same magnitude, direction, and reference point.
  • Example: r₁ and r₂ in the excerpt identify specific locations relative to the origin.

🌊 Position-free vectors

  • Not tied to any particular location in space.
  • Describe properties that can exist anywhere (velocity, force direction, field vectors).
  • Two position-free vectors are equal if they have the same magnitude and direction, regardless of where they are located.
  • Example: two particles 1 m apart both traveling at 2 m/s in the same direction share the same velocity vector v₁ = v₂, even though they occupy different positions.

Don't confuse: position vectors (which locate points) with position-free vectors (which describe properties independent of location).

➕ Addition, subtraction, and scalar multiplication

➕ Vector addition

  • Add component-wise: C = A + B means C_x = A_x + B_x, C_y = A_y + B_y, C_z = A_z + B_z.
  • Commutative: A + B = B + A (order doesn't matter).
  • Example: two forces applied to the same point combine into a single resultant force.

➖ Vector subtraction

  • Subtract component-wise: D = A - B means D_x = A_x - B_x, etc.
  • Not commutative: A - BB - A.
  • Key application: relative position r₁₂ = r₂ - r₁ gives the vector pointing from position 1 to position 2.
  • The magnitude |r₁₂| is the distance between the two points.
  • The unit vector r₁₂ / |r₁₂| indicates the direction from point 1 to point 2.

✖️ Scalar multiplication

  • Multiplying vector A by scalar α gives αA: same direction, magnitude scaled by α.
  • Example: doubling a force F yields 2F (twice the magnitude, same direction).

🔢 The dot product (scalar product)

🎯 What the dot product measures

The dot product (or scalar product) measures the similarity in direction between two vectors.

  • Written A · B, pronounced "A dot B."
  • Formula: A · B = A B cos ψ, where ψ is the angle between the vectors (measured tail-to-tail).
  • Result is a scalar, not a vector.

📐 Special cases and properties

ConditionAngle ψResultMeaning
Perpendicularπ/2 (90°)0No directional similarity
Same direction0ABMaximum positive similarity
Opposite directionπ (180°)-ABMaximum negative similarity
  • Commutative: A · B = B · A.
  • Distributive: A · (B + C) = A · B + A · C.
  • Self dot product: A · A = |A|² = A².

🧮 Cartesian calculation

  • Easy component formula: A · B = A_x B_x + A_y B_y + A_z B_z.
  • Eliminates the need to find angle ψ directly.
  • To find the angle: calculate the dot product using components, then solve cos ψ = (A · B) / (A B) for ψ.
  • Basis vector dot products: x-hat · x-hat = y-hat · y-hat = z-hat · z-hat = 1; all other basis pairs give zero.

⚡ The cross product

🔄 What the cross product produces

The cross product is a form of vector multiplication that results in a vector that is perpendicular to both operands.

  • Written A × B, pronounced "A cross B."
  • Formula: A × B = n-hat A B sin ψ_AB, where n-hat is determined by the right-hand rule.
  • Result is a vector, not a scalar.

🖐️ Right-hand rule for direction

  • Curl the fingers of your right hand from A toward B through angle ψ_AB.
  • Your extended thumb points in the direction of n-hat.
  • Not commutative; instead anticommutative: A × B = -(B × A).
  • Distributive: A × (B + C) = A × B + A × C.

📋 Special cases

  • Vector crossed with itself: A × A = 0 (zero vector).
  • Perpendicular vectors (ψ = π/2): A × B = n-hat A B (maximum magnitude).
  • Basis vector cross products:
    • x-hat × x-hat = y-hat × y-hat = z-hat × z-hat = 0.
    • Counter-clockwise order (x→y→z→x): x-hat × y-hat = z-hat, y-hat × z-hat = x-hat, z-hat × x-hat = y-hat.
    • Clockwise order gives the negative.

🧮 Cartesian calculation

  • Component formula: A × B = x-hat (A_y B_z - A_z B_y) + y-hat (A_z B_x - A_x B_z) + z-hat (A_x B_y - A_y B_x).
  • Can be remembered as a matrix determinant with basis vectors in the first row, A components in the second, B components in the third.
  • Avoids manually determining n-hat using the right-hand rule.

Don't confuse: dot product (scalar result, measures directional alignment) with cross product (vector result, produces perpendicular direction).

40

Cartesian Coordinates

4.2 Cartesian Coordinates

🧭 Overview

🧠 One-sentence thesis

The Cartesian coordinate system is most useful when the geometry of the problem—whether a path, surface, or volume—aligns naturally with constant or varying x, y, z directions, minimizing the number of coordinates that must vary during integration.

📌 Key points (3–5)

  • Integration over length, area, and volume: Cartesian coordinates provide formulas for integrating vector fields along curves, over surfaces, and through volumes using differential elements dl, ds, and dv.
  • Differential elements: length is dl = x̂ dx + ŷ dy + ẑ dz; area is ds (a vector normal to the surface with magnitude equal to differential area); volume is dv = dx dy dz.
  • When Cartesian is appropriate: choose Cartesian when the geometry (straight lines, rectangles, cubes) can be described with the fewest varying coordinates.
  • Common confusion: the direction of the surface normal ds—there are two perpendicular directions; the choice determines the sign of the result and matters for flux calculations.
  • Why coordinate choice matters: using the wrong system (e.g., Cartesian for a circle or cylinder) forces all three basis directions to vary, dramatically complicating the mathematics.

📏 Integration over length

📏 What the differential length element is

Differential length element: dl = x̂ dx + ŷ dy + ẑ dz

  • This describes a tiny segment of a curve C in space.
  • It breaks the segment into components along each Cartesian axis.
  • The integral of a vector field A along the curve is the sum of A · dl over all segments.

🧮 How to integrate along a curve

  • The integral is written as: integral over C of A · dl.
  • The dot product A · dl gives the contribution from each differential segment.
  • Example: if A = x̂ A₀ (constant) and C is a straight line from x = x₁ to x = x₂ along constant y and z, then dl = x̂ dx, A · dl = A₀ dx, and the integral becomes A₀ (x₂ − x₁).
  • When A₀ = 1, this integral simply gives the length of C.

⚠️ When Cartesian is not ideal for length

  • The excerpt gives a counter-example: if C is a circle in the z = 0 plane, both x and y must vary.
  • In this case, cylindrical coordinates (which use only one varying coordinate, φ) would be simpler.
  • Don't confuse: the formalism may seem unnecessary for simple straight-line examples, but it becomes essential for paths that vary in multiple directions or have complicated integrands.

📐 Integration over area

📐 What the differential surface element is

Differential surface element: ds—a vector with magnitude equal to the differential area ds, normal (perpendicular) to each point on the surface.

  • The integral of a vector field A over a surface S is: integral over S of A · ds.
  • The direction of ds is perpendicular to the surface at each point.
  • There are actually two such perpendicular directions (pointing opposite ways).

🧮 How to integrate over a surface

  • Example: if A = ẑ and S is the surface bounded by x₁ ≤ x ≤ x₂, y₁ ≤ y ≤ y₂, then ds = ẑ dx dy.
  • Here, dx dy is the differential area in the z = 0 plane, and ẑ is normal to that plane.
  • A · ds = dx dy, so the integral becomes (x₂ − x₁)(y₂ − y₁), which is the area of the rectangle.
  • Cartesian is appropriate here because the surface is described by constant z with variable x and y—only two coordinates vary.

⚠️ The ambiguity of surface normal direction

  • Why not choose −ẑ instead of +ẑ? Both are normal to the surface.
  • Choosing the opposite direction gives a negative area.
  • "Negative area" means the expected (positive) area, but with respect to the opposite-facing normal.
  • When the sign matters: if A represents a flux density, the direction of ds defines the direction of positive flux.
    • Example: electric flux density D (Section 2.4) has positive flux flowing away from a positively-charged source.
  • Don't confuse: in simple area calculations the sign may not matter, but in flux problems the choice of normal direction is physically meaningful.

🚫 When Cartesian is not ideal for area

  • If the surface is a cylinder or sphere, all three basis directions would be variable, and the surface normal itself would vary.
  • This makes the problem dramatically more complicated in Cartesian coordinates.

📦 Integration over volume

📦 What the differential volume element is

Differential volume element: dv = dx dy dz

  • This is the contribution from a tiny box-shaped volume element.
  • The integral of a scalar or vector quantity A(r) over a volume V is: integral over V of A(r) dv.
  • The procedure is the same whether A is scalar or vector; the excerpt uses a scalar for simplicity.

🧮 How to integrate over a volume

  • Example: if A(r) = 1 and V is a cube bounded by x₁ ≤ x ≤ x₂, y₁ ≤ y ≤ y₂, z₁ ≤ z ≤ z₂, the integral becomes (x₂ − x₁)(y₂ − y₁)(z₂ − z₁).
  • This is simply a calculation of the volume of the cube.
  • Cartesian is appropriate because a cube is easy to describe in Cartesian coordinates and relatively difficult in any other system.

🔄 Choosing the right coordinate system

🔄 The principle: minimize varying coordinates

  • The Cartesian system is best when the geometry aligns with x, y, z directions.
  • Straight line along one axis: only one coordinate varies (e.g., x from x₁ to x₂ with constant y, z).
  • Rectangle in a plane: two coordinates vary (e.g., x and y with constant z).
  • Cube: all three coordinates vary, but the boundaries are simple constants.

🚫 When to avoid Cartesian

GeometryProblem in CartesianBetter system
Circle in z = 0 planeBoth x and y varyCylindrical (only φ varies)
Cylinder or sphere surfaceAll three directions variable, normal variesCylindrical or spherical
  • The excerpt emphasizes: using the wrong system forces unnecessary complexity and "dramatically more complicated" mathematics.
  • Don't confuse: the formalism may seem like overkill for simple examples, but it becomes essential for more complicated integrands and paths/surfaces/volumes that vary in multiple directions.
41

Cylindrical Coordinates

4.3 Cylindrical Coordinates

🧭 Overview

🧠 One-sentence thesis

Cylindrical coordinates simplify problems with cylindrical symmetry by replacing two Cartesian coordinates with a radius and angle, dramatically reducing mathematical complexity in integration and vector analysis.

📌 Key points (3–5)

  • When to use cylindrical coordinates: problems with cylindrical symmetry (e.g., circles, cylinders concentric with the z-axis) become much simpler than in Cartesian coordinates.
  • The three coordinates: ρ (distance from the z-axis), φ (angle in the plane of constant z, starting from +x toward +y), and z (same as Cartesian z).
  • Key difference from Cartesian: basis vectors ρ̂ and φ̂ change direction depending on position, unlike Cartesian basis vectors which are constant everywhere.
  • Common confusion: the differential length contribution from φ is ρdφ (not just dφ) because φ is an angle, not a distance; the arc length depends on the radius.
  • Why it matters: integration over curves, surfaces, and volumes becomes dramatically easier when the geometry matches the coordinate system.

📐 Coordinate system definition

📐 The three cylindrical coordinates

The cylindrical system replaces x and y from Cartesian coordinates with two new parameters:

  • ρ (rho): the distance measured from the closest point on the z-axis
  • φ (phi): the angle measured in a plane of constant z, beginning at the +x axis (φ = 0) with φ increasing toward the +y direction
  • z: identical to the Cartesian z coordinate

Some textbooks use "r" instead of ρ for the radial coordinate.

🧭 Basis vectors

The basis vectors are ρ̂, φ̂, and ẑ.

  • Like basis vectors: dot product equals one
  • Unlike basis vectors: dot product equals zero
  • Cross products follow a cyclic pattern:
    • ρ̂ × φ̂ = ẑ
    • φ̂ × ẑ = ρ̂
    • ẑ × ρ̂ = φ̂

A diagram summarizes these relationships (Figure 4.12 in the excerpt).

🔄 Converting between coordinate systems

🔄 Cylindrical to Cartesian

To convert from cylindrical to Cartesian coordinates:

  • x = ρ cos φ
  • y = ρ sin φ
  • z is identical in both systems

🔄 Cartesian to cylindrical

To convert from Cartesian to cylindrical coordinates:

  • ρ = square root of (x squared plus y squared)
  • φ = arctan(y, x), where arctan is the four-quadrant inverse tangent function

Important: The four-quadrant inverse tangent considers the signs of both x and y individually, not just their ratio. In the first quadrant (x > 0, y > 0) it is arctan(y/x), but other quadrants may require adjustment. This function is available in MATLAB and Octave as atan2(y,x).

🔄 Converting basis vectors

Basis vectors can be converted using dot products.

Cylindrical to Cartesian basis vectors:

  • x̂ = ρ̂ cos φ − φ̂ sin φ
  • ŷ = ρ̂ sin φ + φ̂ cos φ
  • ẑ requires no conversion

Cartesian to cylindrical basis vectors:

  • ρ̂ = x̂ cos φ + ŷ sin φ
  • φ̂ = −x̂ sin φ + ŷ cos φ

A table (Table 4.1) summarizes dot products between basis vectors in the two systems.

⚠️ Position-dependent basis vectors

⚠️ Why cylindrical is awkward for positions

The cylindrical system is usually less useful than Cartesian for identifying absolute and relative positions because the basis directions depend on position.

  • Example: ρ̂ is directed radially outward from the z-axis, so ρ̂ = x̂ for locations along the x-axis but ρ̂ = ŷ for locations along the y-axis
  • Similarly, the direction φ̂ varies as a function of position

Common strategy: Set up a problem in cylindrical coordinates to exploit cylindrical symmetry, but convert to Cartesian coordinates at some point to overcome this awkwardness.

📏 Integration over length

📏 Differential length element

A differential-length segment of a curve in cylindrical coordinates is:

dl = ρ̂ dρ + φ̂ ρ dφ + ẑ dz

Critical detail: The contribution of the φ coordinate to differential length is ρ dφ, not simply dφ, because φ is an angle, not a distance.

📏 Why the φ term is ρ dφ

The circumference of a circle of radius ρ is 2πρ. If only a fraction of the circumference is traversed:

  • The associated arc length is the circumference scaled by φ/(2π), where φ is the angle formed by the traversed circumference
  • Therefore, the distance is 2πρ · φ/(2π) = ρφ
  • The differential distance is ρ dφ

📏 Example: circumference of a circle

To calculate the integral of A(r) = φ̂ over a circle of radius ρ₀ in the z = 0 plane:

  • Along the curve C, ρ = ρ₀ and z = 0 are both constant
  • So dl = φ̂ ρ₀ dφ
  • A · dl = ρ₀ dφ
  • The integral from 0 to 2π of ρ₀ dφ = 2πρ₀

This is a calculation of circumference. The cylindrical system is appropriate here because the problem can be expressed with the minimum number of varying coordinates. In Cartesian, both x and y would vary over C in a relatively complex way.

📐 Integration over area

📐 Differential surface element for planar surfaces

For a circular surface S in the z = 0 plane with radius ρ₀:

ds = ẑ (dρ)(ρ dφ) = ẑ ρ dρ dφ

  • The quantities in parentheses are the radial and angular dimensions
  • The direction of ds indicates the direction of positive flux

📐 Example: area of a circle

To calculate the integral of A = ẑ over the circular surface:

  • A · ds = ρ dρ dφ
  • The integral from ρ = 0 to ρ₀ and φ = 0 to 2π of ρ dρ dφ
  • This separates into (integral of ρ dρ from 0 to ρ₀) times (integral of dφ from 0 to 2π)
  • Result: (one-half ρ₀ squared) times (2π) = π ρ₀ squared

This is the area of the circle, as expected. The corresponding calculation in Cartesian is quite difficult in comparison.

📐 Differential surface element for curved surfaces

For a cylindrical surface S concentric with the z-axis, radius ρ₀, extending from z = z₁ to z = z₂:

ds = ρ̂ (ρ₀ dφ)(dz) = ρ̂ ρ₀ dφ dz

📐 Example: curved surface area of a cylinder

To calculate the integral of A = ρ̂ over the cylindrical surface:

  • The integral from φ = 0 to 2π and z = z₁ to z₂ of ρ₀ dφ dz
  • Result: ρ₀ times (integral of dφ) times (integral of dz) = 2π ρ₀ (z₂ − z₁)

This is the area of S, as expected. Again, the Cartesian calculation would be quite difficult in comparison.

📦 Integration over volume

📦 Differential volume element

The differential volume element in cylindrical coordinates is:

dv = dρ (ρ dφ) dz = ρ dρ dφ dz

📦 Example: volume of a cylinder

If A(r) = 1 and the volume V is a cylinder bounded by ρ ≤ ρ₀ and z₁ ≤ z ≤ z₂:

  • The integral over V of A(r) dv becomes the integral from ρ = 0 to ρ₀, φ = 0 to 2π, and z = z₁ to z₂ of ρ dρ dφ dz
  • This separates into three integrals: (integral of ρ dρ) times (integral of dφ) times (integral of dz)
  • Result: π ρ₀ squared times (z₂ − z₁)

This is area times length, which is volume.

📦 When the formalism is necessary

The procedure above is more complicated than necessary if we are only computing volume. However, if the integrand is not constant-valued, then we are no longer simply computing volume—in this case, the formalism is appropriate and possibly necessary.

42

Spherical Coordinates

4.4 Spherical Coordinates

🧭 Overview

🧠 One-sentence thesis

Spherical coordinates dramatically simplify problems with spherical symmetry by describing positions with radius r, polar angle θ, and azimuthal angle φ, making integration and analysis much easier than Cartesian coordinates for such geometries.

📌 Key points (3–5)

  • When to use spherical coordinates: preferred when the geometry exhibits spherical symmetry (e.g., a sphere's surface needs only one parameter r, versus three in Cartesian).
  • The three coordinates: r (distance from origin), θ (angle from +z axis toward z=0 plane), φ (angle in constant-z plane, same as cylindrical φ).
  • Basis vectors are position-dependent: r-hat, θ-hat, and φ-hat change direction depending on location, making absolute positioning awkward—often requires conversion to/from Cartesian.
  • Common confusion: the differential lengths are not simply dr, dθ, dφ; they are dr, r dθ, and r sin θ dφ because θ and φ are angles, not distances.
  • Integration formulas scale with geometry: differential length, area, and volume elements include geometric factors (r, r², sin θ) that reflect spherical geometry.

📐 Coordinate system definition

📐 The three spherical coordinates

Spherical coordinate system: defined by r (distance from origin), θ (angle from +z axis toward z=0 plane), and φ (angle in constant-z plane, identical to cylindrical φ).

  • r: radial distance from the origin (some textbooks use R instead).
  • θ: polar angle measured from the positive z-axis downward.
  • φ: azimuthal angle, measured in planes of constant z, exactly as in cylindrical coordinates.
  • Example: A sphere centered at the origin is described by a single constant r = a, whereas Cartesian requires all three coordinates x, y, z to describe the same surface.

🧭 Basis vectors and their properties

The spherical system uses three basis vectors: r-hat, θ-hat, and φ-hat.

Dot products:

  • Like basis vectors: r-hat · r-hat = 1, θ-hat · θ-hat = 1, φ-hat · φ-hat = 1.
  • Unlike basis vectors: all equal zero.

Cross products:

  • r-hat × θ-hat = φ-hat
  • θ-hat × φ-hat = r-hat
  • φ-hat × r-hat = θ-hat

A diagram (Figure 4.17) summarizes these relationships in a cyclic pattern.

⚠️ Position-dependent basis directions

  • Unlike Cartesian basis vectors (which are constant everywhere), spherical basis vectors change direction with position.
  • Example: r-hat points along x-hat on the x-axis, along y-hat on the y-axis, and along z-hat on the z-axis.
  • Similarly, θ-hat and φ-hat vary as functions of position.
  • Don't confuse: this position-dependence makes spherical coordinates less useful for identifying absolute and relative positions compared to Cartesian.
  • Common workaround: start a problem in spherical coordinates, then convert to Cartesian later in the analysis.

🔄 Coordinate conversions

🔄 Spherical to Cartesian

The conversions from spherical (r, θ, φ) to Cartesian (x, y, z):

  • x = r cos φ sin θ
  • y = r sin φ sin θ
  • z = r cos θ

🔄 Cartesian to spherical

The conversions from Cartesian (x, y, z) to spherical (r, θ, φ):

  • r = square root of (x² + y² + z²)
  • θ = arccos(z / r)
  • φ = arctan(y, x), where arctan is the four-quadrant inverse tangent function (available in MATLAB/Octave as atan2(y,x))

🔄 Basis vector conversions

Table 4.2 provides dot products between spherical and Cartesian basis vectors:

·r-hatθ-hatφ-hat
x-hatsin θ cos φcos θ cos φ−sin φ
y-hatsin θ sin φcos θ sin φcos φ
z-hatcos θ−sin θ0

These dot products enable conversion between basis vectors using formulas like:

  • x-hat = r-hat (r-hat · x-hat) + θ-hat (θ-hat · x-hat) + φ-hat (φ-hat · x-hat)
  • r-hat = x-hat (x-hat · r-hat) + y-hat (y-hat · r-hat) + z-hat (z-hat · r-hat)

Example from the excerpt: A vector field G = x-hat xz/y can be converted to spherical by substituting Cartesian expressions with spherical ones using Table 4.2 and the coordinate conversions, yielding G = (r-hat sin θ cot φ + θ-hat cos θ cot φ − φ-hat) · r cos θ cos φ.

📏 Integration over length

📏 Differential length element

Differential length in spherical coordinates: dl = r-hat dr + θ-hat r dθ + φ-hat r sin θ dφ

  • Why not just dr, dθ, dφ? Because θ and φ are angles, not distances.
  • The distance associated with angle θ is r dθ in the θ direction.
  • The distance associated with angle φ is r dφ in the z=0 plane, reduced by factor sin θ for z ≠ 0.
  • Don't confuse: the coefficients (1, r, r sin θ) reflect the geometry of the spherical system.

📏 Line integral example: pole-to-pole arc

The excerpt calculates the integral of vector field A(r) = θ-hat along a curve C from north pole to south pole on a sphere of radius a.

  • Setup: r = a (constant), φ = constant (any value), so dl = θ-hat a dθ.
  • A · dl = a dθ.
  • Integral: from θ=0 to θ=π of a dθ = πa.
  • Result: half the circumference of the sphere, as expected.
  • Why spherical is appropriate: only one coordinate (θ) varies along C; in Cartesian, both z and x or y (or all three) would vary in a complex way.

📐 Integration over area

📐 Differential surface element

Differential surface in spherical coordinates: ds = r-hat (r dθ)(r sin θ dφ) = r-hat r² sin θ dθ dφ

  • The direction is normal to the surface, in the direction of positive flux.
  • The quantities in parentheses are the distances associated with varying θ and φ, respectively.
  • General surface integral: integral over S of A · ds.

📐 Surface integral example: area of a sphere

The excerpt calculates the integral of A = r-hat over the surface S of a sphere of radius a centered at the origin.

  • A · ds = a² sin θ dθ dφ.
  • Integral: from θ=0 to π and φ=0 to 2π of a² sin θ dθ dφ.
  • Factored: a² · (integral of sin θ dθ from 0 to π) · (integral of dφ from 0 to 2π).
  • Result: a² · 2 · 2π = 4πa², the area of the sphere.
  • Why spherical simplifies: the corresponding calculation in Cartesian or cylindrical is "quite difficult in comparison."

📦 Integration over volume

📦 Differential volume element

Differential volume in spherical coordinates: dv = dr (r dθ)(r sin θ dφ) = r² dr sin θ dθ dφ

  • The three factors correspond to the differential lengths in the r, θ, and φ directions.
  • General volume integral: integral over V of A(r) dv.

📦 Volume integral example: volume of a sphere

The excerpt calculates the integral of A(r) = 1 over volume V, a sphere of radius a centered at the origin.

  • Integral: from r=0 to a, θ=0 to π, φ=0 to 2π of r² dr sin θ dθ dφ.
  • Factored: (integral of r² dr from 0 to a) · (integral of sin θ dθ from 0 to π) · (integral of dφ from 0 to 2π).
  • Result: (one-third a³) · 2 · 2π = four-thirds πa³, the volume of a sphere.
  • Comparison to cylindrical example: the excerpt earlier showed a cylindrical volume integral yielding πρ₀²(z₂ − z₁) = area times length; both examples show that when the integrand is constant, the formalism computes geometric volume, but the formalism is "appropriate and possibly necessary" when the integrand is not constant.
43

Gradient

4.5 Gradient

🧭 Overview

🧠 One-sentence thesis

The gradient operator transforms a scalar field into a vector field that points in the direction of the field's most rapid increase, with magnitude equal to the rate of change in that direction.

📌 Key points (3–5)

  • What the gradient does: converts a scalar field into a vector field pointing toward the steepest increase, with magnitude equal to the rate of change.
  • Key application in electromagnetics: the gradient relates electric field intensity E(r) (a vector field) to electric potential V(r) (a scalar field) via E = −∇V.
  • Why the negative sign matters: E points in the direction where V decreases most quickly, hence the minus sign in the formula.
  • Common confusion: the gradient is a general mathematical operator; the electric field example is just one application—many other scalar fields can have gradients.
  • Broader role: the gradient operator is foundational for other differential operators in electromagnetics (divergence, curl, Laplacian).

🔍 Core concept: what the gradient measures

🔍 Definition and meaning

Gradient of a scalar field: a vector that points in the direction in which the field is most rapidly increasing, with the scalar part equal to the rate of change.

  • Input: a scalar field (a function that assigns a number to each point in space).
  • Output: a vector field (a function that assigns a vector to each point in space).
  • The direction of the gradient vector tells you where the scalar field grows fastest.
  • The magnitude of the gradient vector tells you how fast it grows in that direction.

🧮 Cartesian formula

In the Cartesian coordinate system, the gradient of a scalar field f is:

∇f = x̂ (∂f/∂x) + ŷ (∂f/∂y) + ẑ (∂f/∂z)

  • ∂f/∂x, ∂f/∂y, ∂f/∂z are the partial derivatives of f with respect to x, y, and z.
  • Each component measures how f changes along that coordinate axis.
  • The symbol (nabla) denotes the gradient operator.

📐 Simple example: ramp function

Example: Find the gradient of f = ax (a "ramp" with slope a along the x direction).

  • ∂f/∂x = a (the slope along x).
  • ∂f/∂y = 0 and ∂f/∂z = 0 (no change along y or z).
  • Therefore ∇f = x̂ a.
  • Interpretation: ∇f points in the x direction (where f increases) and has magnitude a (the slope).

⚡ Application: electric field and electric potential

⚡ The relationship E = −∇V

The gradient connects the electric field intensity E(r) (a vector field, units V/m) to the electric potential V(r) (a scalar field, units V):

E(r) = −∇V(r)

or simply E = −∇V when position dependence is understood.

🔽 Why the negative sign?

  • The direction of E(r) is the direction in which V(r) decreases most quickly.
  • The magnitude of E(r) is the rate of change of V(r) in that direction.
  • Because E points toward decreasing potential (not increasing), the formula includes a minus sign.

📏 Units consistency

  • V(r) has units of volts (V).
  • E(r) has units of volts per meter (V/m).
  • The gradient operator involves spatial derivatives (per meter), so ∇V naturally has units V/m, matching E.

🔌 Reference to earlier material

The excerpt notes that this relationship is demonstrated in Section 2.2 ("Electric Field Intensity"), particularly in a battery-charged capacitor example, where:

  • E(r) points in the direction where V(r) decreases most quickly.
  • The scalar part of E(r) equals the rate of change of V(r) in that direction.

🧰 Broader context and related operators

🧰 Gradient as a general tool

  • The electric field example is just one application of the gradient.
  • In electromagnetics, you often need the gradient ∇f of various scalar fields f(r).
  • Don't confuse: the gradient is not limited to electric potential; it applies to any scalar field.

🔗 Foundation for other operators

The excerpt emphasizes that the gradient operator is foundational for other important differential operators in electromagnetics:

OperatorSection referenceRole
DivergenceSection 4.6Measures flux at a point
CurlSection 4.8Measures rotation of a vector field
LaplacianSection 4.10Second-order operator

All of these can be interpreted or defined in terms of the gradient operator .

🌐 Other coordinate systems

  • The Cartesian formula is given explicitly in the excerpt.
  • The gradient operator also exists in cylindrical and spherical coordinate systems.
  • The excerpt refers to Appendix B.2 for those formulas (not reproduced here).
44

Divergence

4.6 Divergence

🧭 Overview

🧠 One-sentence thesis

Divergence measures the flux per unit volume through an infinitesimally-small closed surface surrounding a point, providing a way to calculate local source or sink behavior of a vector field at that point.

📌 Key points (3–5)

  • What divergence measures: flux per unit volume through an infinitesimally-small closed surface around a point, converting a surface concept (flux) into a point-based property.
  • How to calculate it: divergence of a vector field A is the dot product ∇ · A, which in Cartesian coordinates equals the sum of partial derivatives of each component.
  • Physical interpretation: divergence tells you the net "outflow" at a point—positive divergence means the point acts as a source, zero divergence means uniform flow, negative divergence means a sink.
  • Common confusion: flux vs. divergence—flux is integrated over a finite surface (scalar result), while divergence is a local property at a point (also scalar, but describes density of sources/sinks).
  • Why it matters: divergence connects field behavior to sources (e.g., Gauss' Law: div D = ρ_v relates electric flux density to charge density).

🌊 From flux to divergence

🌊 What is flux?

Flux is the scalar quantity obtained by integrating a vector field, interpreted in this case as a flux density, over a specified surface.

  • Mathematically: the integral of a vector field A(r) over a surface S gives flux F.
  • Units: if A has units of (something)/m², then F has units of (something).
  • Example: a vector field with unit magnitude normal to all points on S has flux equal to the area of S.

🔬 Two electromagnetic examples

Flux densityUnitsFlux obtained by integratingUnitsPhysical meaning
Electric flux density DC/m²Enclosed charge Q_enclCGauss' Law: integral of D over closed surface = total charge inside
Magnetic flux density BWb/m²Magnetic flux ΦWbTime rate of change of Φ is proportional to electric potential (Faraday's Law)
  • D and B are called "flux densities" because integrating them over area yields a flux quantity.
  • Don't confuse: flux density is the field itself (per unit area); flux is the integrated total over a surface.

📍 Moving from surface to point

  • Flux applies to a finite surface, but often we want to know behavior at a single point.
  • Example: instead of total charge (C) enclosed by a surface, we want charge density (C/m³) at a point.
  • Strategy: start with flux integral over a closed surface S enclosing volume V, divide both sides by V, then let V shrink to zero.
  • Taking the limit as V → 0 on the left side gives divergence; on the right side (for D) gives volume charge density ρ_v.

🧮 Calculating divergence

🧮 Definition via the ∇ operator

Divergence is the flux per unit volume through an infinitesimally-small closed surface surrounding a point.

  • Notation: div A = ∇ · A (read "del dot A").
  • The symbol ∇ is the same operator used in the gradient (Section 4.5) and appears in other differential operators (curl, Laplacian).
  • In Cartesian coordinates:
    • ∇ = x̂ ∂/∂x + ŷ ∂/∂y + ẑ ∂/∂z
    • If A = x̂ A_x + ŷ A_y + ẑ A_z, then div A = ∂A_x/∂x + ∂A_y/∂y + ∂A_z/∂z

🔍 Why this formula makes sense

  • Dimensionally correct: taking the derivative of a quantity with units C/m² with respect to distance (m) yields C/m³.
  • Intuitive: flux from a point should relate to the sum of the rates of change of the flux density in each basis direction.
  • Example: if the x-component of A is increasing in the x-direction, there is net outward flux in that direction.

🌐 Other coordinate systems

  • Cylindrical: ∇ = ρ̂ (1/ρ) ∂/∂ρ (ρ·) + φ̂ (1/ρ) ∂/∂φ (ρ·) + ẑ ∂/∂z
  • Spherical: ∇ = r̂ (1/r²) ∂/∂r (r²·) + θ̂ (1/(r sin θ)) ∂/∂θ (sin θ ·) + φ̂ (1/(r sin θ)) ∂/∂φ
  • Alternatively, explicit expressions for divergence in these systems are given in Appendix B.2.
  • Choose the coordinate system that matches the symmetry of the problem to simplify calculations.

⚖️ Linearity of divergence

  • Divergence is a linear operator: for any constant scalars a and b and vector fields A and B:
    • ∇ · (aA + bB) = a ∇ · A + b ∇ · B
  • This property is useful for breaking down complex fields into simpler components.

📐 Worked examples

📐 Uniform field (Example 4.5)

  • Field: A = x̂ A_x + ŷ A_y + ẑ A_z where A_x, A_y, A_z are all constants.
  • Divergence: ∇ · A = 0 because each component is constant with respect to position (all partial derivatives are zero).
  • Interpretation: if the flux is uniform, the flux into an infinitesimally-small closed surface equals the flux out, resulting in net flux of zero.
  • Don't confuse: a field can be large in magnitude but still have zero divergence if it is uniform.

📈 Linearly-increasing field (Example 4.6)

  • Field: A = x̂ A_0 x where A_0 is a constant.
  • Divergence: ∇ · A = A_0.
  • Interpretation: if A is a flux density, the net flux per unit volume is simply the rate at which the flux density is increasing with distance.
  • Example: as you move in the x-direction, the field grows linearly, so there is a constant positive divergence (source behavior).

📉 Radially-decreasing field (Example 4.7)

  • Field: A = r̂ A_0/r where A_0 is a constant (directed radially outward, decreasing linearly with distance).
  • Coordinate system: spherical is easiest because partial derivatives in θ and φ directions are zero.
  • Divergence: ∇ · A = (1/r²) ∂/∂r (r² · A_0/r) = A_0/r².
  • Interpretation: if A is a flux density, the flux per unit volume decreases with the square of distance from the origin.
  • Don't confuse: the field itself decreases as 1/r, but the divergence (flux per unit volume) decreases as 1/r².

🔗 Connection to Gauss' Law

🔗 Divergence of electric flux density

  • Starting from Gauss' Law: the integral of D over a closed surface S equals the enclosed charge Q_encl.
  • Divide both sides by the enclosed volume V and take the limit as V → 0.
  • Left side becomes div D; right side becomes volume charge density ρ_v.
  • Result: div D = ρ_v (Equation 4.98).
  • Physical meaning: the divergence of the electric flux density at a point equals the charge density at that point—positive charge is a source of electric flux, negative charge is a sink.

🧲 Divergence of magnetic flux density

  • The excerpt mentions magnetic flux density B and magnetic flux Φ.
  • Faraday's Law relates the time rate of change of Φ to electric potential.
  • The excerpt does not explicitly state the divergence of B, but the framework is the same: divergence would relate B to magnetic charge density (which is zero in classical electromagnetics).
45

Divergence Theorem

4.7 Divergence Theorem

🧭 Overview

🧠 One-sentence thesis

The Divergence Theorem converts a volume integral of divergence into a surface integral of flux, enabling problems defined throughout a volume to be solved using only information on the bounding surface (and vice versa).

📌 Key points (3–5)

  • What the theorem states: the integral of the divergence of a vector field over a volume equals the flux of that field through the surface bounding that volume.
  • Why it works: flux flowing between internal infinitesimal cubes cancels out; only flux through the outer bounding surface remains.
  • Key application: convert volume-based problems into surface-based problems and vice versa.
  • Common confusion: divergence is flux per unit volume at a point; integrating it over a volume recovers total flux through the boundary, not flux at individual interior points.

🔍 What divergence measures

🔍 Divergence as flux per unit volume

Divergence of a vector field A: ∇ · A = f, where f(r) is the flux per unit volume through an infinitesimally-small closed surface surrounding the point at r.

  • Divergence is not the flux itself; it is the density of flux sources or sinks at a point.
  • Example: if A represents electric flux density D or magnetic flux density B, then ∇ · A tells you how much flux is being generated or absorbed per unit volume at each location.

🧮 Linearity of divergence

  • Divergence is a linear operator: ∇ · (aA + bB) = a(∇ · A) + b(∇ · B) for any constant scalars a and b and vector fields A and B.
  • This means you can split divergence calculations across sums and factor out constants.

🧱 How the theorem is derived

🧱 From flux density to total flux

  • Since f is flux per unit volume, total flux through any larger volume V is obtained by integrating over V:
    • flux through V = integral over V of f dv
  • This step converts a point-by-point density into a cumulative total.

🧊 Cancellation inside the volume

  • Imagine V as a three-dimensional grid of infinitesimally-small cubes with side lengths dx, dy, and dz.
  • Flux out of any face of one cube equals flux into the adjacent cube through that shared face.
  • When summing over all cubes, internal fluxes cancel out.
  • Only fluxes corresponding to faces on the bounding surface S remain, because the integration stops there.
  • Mathematically: integral over V of f dv = closed surface integral over S of A · ds

🔗 Combining the steps

  • Start with ∇ · A = f (Equation 4.106).
  • Integrate both sides over V: integral over V of (∇ · A) dv = integral over V of f dv.
  • Apply the cancellation result (Equation 4.108) to the right-hand side: integral over V of (∇ · A) dv = closed surface integral over S of A · ds.
  • This is the Divergence Theorem (Equation 4.110).

📐 Statement and utility of the theorem

📐 The Divergence Theorem

The integral of the divergence of a vector field over a volume is equal to the flux of that field through the surface bounding that volume.

  • In symbols: integral over V of (∇ · A) dv = closed surface integral over S of A · ds
  • Left side: volume integral (information throughout the interior).
  • Right side: surface integral (information only on the boundary).

🔄 Principal utility

  • Convert volume problems to surface problems: if you know the divergence throughout a volume, you can find the flux through the boundary without evaluating the interior.
  • Convert surface problems to volume problems: if you know the flux through a bounding surface, you can infer the total divergence inside.
  • This is especially useful in electromagnetic analysis, where field quantities may be easier to measure or compute on one domain versus the other.
  • Example: if you know the charge density (which relates to divergence of D) inside a region, you can compute the electric flux through the surface without integrating over every interior point.

⚠️ Don't confuse

  • The theorem does not say that divergence at a point equals flux through a surface; it relates the integral of divergence over a volume to the integral of flux over the boundary.
  • The cancellation argument applies only to the interior; boundary faces do not have adjacent cubes outside the volume, so their contributions survive.
46

Curl

4.8 Curl

🧭 Overview

🧠 One-sentence thesis

Curl quantifies the circulation of a vector field at a point—a concept central to Maxwell's equations—and is most practically computed using the del operator as ∇ × A.

📌 Key points (3–5)

  • What curl measures: the circulation (line integral around a closed path) of a vector field per unit area at a point, in the plane that maximizes this value.
  • Why it matters in electromagnetics: circulation of electric fields relates to changing magnetic fields (Maxwell-Faraday equation), and circulation of magnetic fields relates to current and changing electric fields (Ampere's Law).
  • How to compute it: although defined via a limit of circulation per unit area, curl is practically calculated as the cross product ∇ × A.
  • Direction convention: the direction of curl follows the right-hand rule applied to the path of integration.
  • Common confusion: curl is not the circulation itself (which depends on path size), but the circulation per unit area as the path shrinks to zero area.

🔄 What is circulation?

🔄 Definition and intuition

"Circulation" is the integral of a vector field over a closed path.

  • For a vector field A(r) and closed path C, circulation is the line integral ∮_C A · dl.
  • It measures how much the field "flows around" the closed path.
  • Key insight: circulation captures rotational behavior of the field.

🧪 Uniform vs rotating fields

  • Uniform field: circulation is zero for any valid closed path.
    • Example: A field A = x̂ A₀ (constant in the x-direction) has zero circulation because contributions at each point on the path cancel out when summed over the closed loop.
  • Rotating field: circulation can be non-zero.
    • Example: A field A = φ̂ A₀ over a circular path of constant ρ and z yields non-zero circulation because contributions at each point are equal and accumulate around the path.

🔌 Example: magnetic field around a current

Scenario: Current I flows along the z-axis in the +z direction, producing a magnetic field intensity H = φ̂ H₀/ρ.

  • Circulation of H along a circular path of radius a in a plane of constant z:
    • ∮_C H · dl = 2πH₀
  • Two remarkable features:
    1. Independent of the radius of the integration path.
    2. Has units of amperes (A), suggesting a connection to current.
  • Result: the circulation of H equals the enclosed source current I along any path enclosing the current—a consequence of Ampere's Law.

🌀 From circulation to curl

🌀 The limiting process

Curl answers: "What is the circulation at a point in space?"

  • As the closed path C shrinks to its smallest size, the circulation itself goes to zero (nothing to integrate over).
  • Instead, ask: what is the circulation per unit area in the limit?
  • Constrain C to lie in a plane, and define S as the open surface bounded by C.
  • The scalar part of curl is:
    • lim (Δs → 0) [∮_C A · dl] / Δs
    • where Δs is the area of S.
  • Important: choose the plane that maximizes this result.

🧭 Direction via the right-hand rule

  • Because S and C lie in a plane, a direction can be assigned: the normal to the plane.
  • Two normals exist at each point; the right-hand rule selects the correct one:
    • Align your right thumb along C in the direction of integration.
    • Your fingers point through the plane in the direction of .
  • This convention is arbitrary but ensures consistency (see Stokes' Theorem for justification).

📐 Formal definition

curl A ≡ lim (Δs → 0) [∮_C A · dl] / Δs

  • Magnitude: circulation per unit area at a point, with the path shrinking to enclose zero area in the plane that maximizes the result.
  • Direction: determined by the right-hand rule applied to the integration path.

Don't confuse: Curl is not the total circulation (which depends on path size), but the density of circulation at a point.

⚙️ Practical computation of curl

⚙️ The del operator representation

Although the definition via circulation is conceptually important, it is difficult to apply directly.

curl A ≡ ∇ × A

  • The same ∇ operator used for gradient, divergence, and Laplacian also defines curl via the cross product.
  • This representation is far more practical for calculations.

🧮 Cartesian coordinates

In Cartesian coordinates:

  • ∇ = x̂ ∂/∂x + ŷ ∂/∂y + ẑ ∂/∂z
  • A = x̂ Aₓ + ŷ Aᵧ + ẑ A_z

Curl is computed as a determinant:

Basisŷ
Operator∂/∂x∂/∂y∂/∂z
ComponentsAₓAᵧA_z

Evaluating the determinant:

  • ∇ × A = x̂ (∂A_z/∂y − ∂Aᵧ/∂z) + ŷ (∂Aₓ/∂z − ∂A_z/∂x) + ẑ (∂Aᵧ/∂x − ∂Aₓ/∂y)

🔧 Other coordinate systems

  • Expressions for curl in cylindrical and spherical coordinates are provided in Appendix B.2 (not included in this excerpt).

➕ Linearity property

Curl is a linear operator:

  • ∇ × (aA + bB) = a(∇ × A) + b(∇ × B)
  • where a and b are constant scalars and A and B are vector fields.

🔗 Connection to Maxwell's equations

⚡ Maxwell-Faraday equation

  • The circulation of an electric field is proportional to the rate of change of the magnetic field.
  • This is the Maxwell-Faraday equation (Section 8.8).
  • Special case: Kirchhoff's Voltage Law for electrostatics (Section 5.11).

🧲 Ampere's Law

  • The circulation of a magnetic field is proportional to the source current and the rate of change of the electric field.
  • This is Ampere's Law (Sections 7.9 and 8.9).
  • Example: the magnetic field circulation around a line current equals the enclosed current.

Why curl matters: These two fundamental laws of electromagnetism are expressed in terms of circulation, making curl essential for electromagnetic analysis.

📘 Stokes' Theorem (brief mention)

📘 Relating surface and line integrals

Stokes' Theorem connects curl to circulation:

  • ∫_S (∇ × A) · ds = ∮_C A · dl
  • where S is an open surface bounded by the closed path C.
  • The surface normal ds = ds and the direction of integration along C are related by the right-hand rule (same convention as for curl).

Utility: Transforms problems defined over a surface into problems over a bounding curve, or vice versa—a purely mathematical tool for electromagnetic analysis.

47

Stokes' Theorem

4.9 Stokes’ Theorem

🧭 Overview

🧠 One-sentence thesis

Stokes' Theorem provides a mathematical tool to convert surface integrals of curl into line integrals around the boundary, enabling flexible problem-solving in electromagnetic analysis.

📌 Key points (3–5)

  • What the theorem states: an integral of curl over an open surface equals a line integral around the curve bounding that surface.
  • Orientation rule: the surface normal direction and integration direction along the boundary are related by the right-hand rule.
  • Nature of the theorem: purely mathematical, not a physical principle of electromagnetics itself.
  • Why it matters: serves as a transformation tool to switch between surface integrals and closed-path integrals depending on which is easier to solve.
  • Common confusion: this is a mathematical technique for analysis, not a statement about electromagnetic phenomena.

🔄 The theorem statement

🔄 Mathematical relationship

Stokes' Theorem: the integral over surface S of (curl of A) dot ds equals the closed line integral over curve C of A dot dl.

  • Written symbolically: integral over S of (del cross A) dot ds = closed integral over C of A dot dl.
  • S is the open surface.
  • C is the closed path that bounds surface S.
  • The theorem connects two different types of integrals of the same vector field A.

📐 The right-hand rule constraint

  • The direction of the surface normal (ds = unit normal times ds) must match the direction of integration along C via the right-hand rule.
  • How it works: align your right thumb along C in the direction of integration; your fingers curl through the surface in the direction the normal must point.
  • This ensures consistency between the two sides of the equation.
  • Don't confuse: the normal direction is not arbitrary—it is determined by the chosen direction of integration around C.

🧰 Role in electromagnetic theory

🧰 A mathematical tool, not a physical law

  • The excerpt emphasizes: "Stokes' Theorem is a purely mathematical result and not a principle of electromagnetics per se."
  • It does not describe how fields behave; it describes how integrals relate.
  • Relevance: it is "primarily as a tool in the associated mathematical analysis."

🔀 Transformation between problem forms

  • Typical use: transform a problem expressed as a surface integral into a closed-path integral, or vice-versa.
  • Why this helps: one form may be easier to evaluate than the other depending on the geometry or known quantities.
  • Example scenario: if you know the vector field A along a boundary curve but not across the entire surface, you can use the line integral instead of the surface integral.

🔗 Connection to curl

🔗 Curl as the integrand

  • The left side of Stokes' Theorem involves the curl of vector field A.
  • Recall from the excerpt's earlier context: curl of A is defined as del cross A, a measure of rotation in the field.
  • The theorem thus relates the "total rotation" over a surface to the circulation around its edge.

📚 Further study

  • The excerpt notes that more information on the theorem and its derivation is available in additional reading (Wikipedia article on Stokes' Theorem).
  • The theorem itself is presented as a definition and application tool, not derived within this section.
48

The Laplacian Operator

4.10 The Laplacian Operator

🧭 Overview

🧠 One-sentence thesis

The Laplacian operator, defined as the divergence of the gradient, serves as a second spatial derivative that links electric potential to charge density and describes wave propagation in electromagnetic fields.

📌 Key points (3–5)

  • What the Laplacian is: the divergence of the gradient of a field, essentially a second derivative with respect to three spatial dimensions.
  • Key application in electrostatics: relates electric potential V to charge density through Poisson's Equation.
  • Why this relationship makes sense: V connects to charge through two derivative steps—gradient links V to electric field E, divergence links E to charge density.
  • Works for both scalars and vectors: can be applied to scalar fields (like potential) and vector fields (like electric field in wave equations).
  • Common confusion: the Laplacian is not a single derivative but a combination of two operations (gradient then divergence), making it a second-order operator.

🔧 Definition and basic form

🔧 What the Laplacian operator is

The Laplacian of a field f(r) is the divergence of the gradient of that field: ∇²f ≡ ∇·(∇f).

  • It combines two vector calculus operations: first take the gradient, then take the divergence of the result.
  • The excerpt emphasizes this is "essentially a definition of the second derivative with respect to the three spatial dimensions."
  • In Cartesian coordinates, this becomes the sum of three second partial derivatives: ∇²f = (∂²f/∂x²) + (∂²f/∂y²) + (∂²f/∂z²).
  • This can be verified by applying the definitions of gradient and divergence in Cartesian coordinates to the defining equation.

📐 Coordinate systems

  • The excerpt provides the explicit Cartesian form.
  • For cylindrical and spherical coordinate systems, the Laplacian takes different forms (given in Appendix B.2 of the source material).
  • The operator adapts to the coordinate system but the underlying concept remains the same.

⚡ Application to electric potential

⚡ Poisson's Equation

The Laplacian relates electric potential V (units of V) to electric charge density ρ_v (units of C/m³) through Poisson's Equation: ∇²V = −ρ_v/ε, where ε is the permittivity of the medium.

  • This is a fundamental relationship in electrostatics.
  • The negative sign and the permittivity factor are part of the physical relationship.

🔗 Why this relationship makes physical sense

The excerpt explains the logic behind Poisson's Equation through a two-step derivative chain:

  1. First derivative (gradient): Electric field intensity E (units of V/m) is proportional to the derivative of V with respect to distance via the gradient.
  2. Second derivative (divergence): Charge density ρ_v is proportional to the derivative of E with respect to distance via the divergence.
  • Combining these two steps: V → (gradient) → E → (divergence) → ρ_v.
  • The Laplacian (divergence of gradient) therefore directly connects V to ρ_v.
  • The excerpt states "this should not be surprising" given these two derivative relationships.

🌊 Application to vector fields

🌊 Laplacian of vector fields

  • The Laplacian can be applied to vector fields, not just scalar fields.
  • In Cartesian coordinates, if A = x̂A_x + ŷA_y + ẑA_z, then: ∇²A = x̂∇²A_x + ŷ∇²A_y + ẑ∇²A_z.
  • Each component of the vector field is treated independently with the scalar Laplacian.

📡 Wave equation application

An important application is the wave equation for electric field E in a lossless and source-free region: ∇²E + β²E = 0, where β is the phase propagation constant.

  • This describes electromagnetic wave propagation.
  • The Laplacian of the vector field E appears alongside a term involving the field itself.
  • This application is detailed in Section 9.2 of the source material.

🔄 Alternative expression

The excerpt provides an alternative form for the Laplacian of a vector field:

∇²A = ∇(∇·A) − ∇×(∇×A)

  • This expresses the Laplacian in terms of gradient, divergence, and curl.
  • The excerpt notes "it is sometimes useful to know" this identity.
  • This form can be helpful for certain derivations or coordinate systems.

📊 Summary comparison

AspectScalar fieldVector field
Definition∇²f = ∇·(∇f)∇²A (component-wise in Cartesian)
Cartesian formSum of three second partialsApply scalar Laplacian to each component
Key application (from excerpt)Poisson's Equation (potential to charge)Wave equation (field propagation)
Alternative formNot given∇²A = ∇(∇·A) − ∇×(∇×A)

Don't confuse: The Laplacian is not just "any second derivative"—it is specifically the sum of second derivatives across all three spatial dimensions, capturing how a field varies in all directions simultaneously.

49

Coulomb's Law

5.1 Coulomb’s Law

🧭 Overview

🧠 One-sentence thesis

Coulomb's Law quantifies the force between charged particles and provides the foundation for calculating electric fields from point charges and continuous charge distributions.

📌 Key points (3–5)

  • The force law: the force between two charged particles is proportional to the product of their charges and inversely proportional to the square of the distance between them.
  • Sign determines direction: same-sign charges repel, opposite-sign charges attract.
  • From force to field: Coulomb's Law combined with the definition of electric field intensity yields a direct formula for the field around a point charge.
  • Superposition principle: the total electric field from multiple charges is the vector sum of the fields from each individual charge.
  • Continuous distributions: charge can be described as distributed along lines, over surfaces, or within volumes using density functions.

⚡ The fundamental force between charges

⚡ Coulomb's Law formula

Coulomb's Law: F = (unit vector R) × (q₁ q₂) / (4 π ε R²)

  • F is the force experienced by particle 2
  • q₁ and q₂ are the charges on the two particles
  • R is the distance between them
  • unit vector R points from particle 1 toward particle 2
  • ε is the permittivity of the medium

🧲 How charges interact

  • If the product q₁ q₂ is positive (same sign charges) → particles repel
  • If the product q₁ q₂ is negative (opposite sign charges) → particles attract
  • The force on particle 1 is equal and opposite to the force on particle 2

📐 Where the formula comes from

  • Physical observations show force magnitude is proportional to q₁ q₂
  • Force magnitude is inversely proportional to R²
  • The constant F₀ has units of inverse permittivity: (F/m)⁻¹
  • Observations confirm force is inversely proportional to permittivity ε, with an additional factor of 1/(4π)

🌐 Electric field from point charges

🌐 Defining the electric field

Electric field intensity E₁ is defined by: F = q₂ E₁

  • This is the definition of electric field intensity (from Section 2.2 of the source)
  • It describes the force per unit charge
  • Combining this with Coulomb's Law gives the field from a single charge

📍 Field formula for a point charge

E₁ = (unit vector R) × q₁ / (4 π ε R²)

Where unit vector R is the vector from particle 1 to the evaluation point.

Four key properties:

  1. Directed away from positive charge (toward negative charge)
  2. Proportional to the magnitude of the charge
  3. Inversely proportional to permittivity of the medium
  4. Inversely proportional to distance squared

🔢 Example: electron at the origin

  • For a charge q at the origin, using spherical coordinates: E(r) = (radial unit vector) × q / (4 π ε r²)
  • Numerical case: single electron 1 μm away in free space
    • Charge: q ≈ −1.60 × 10⁻¹⁹ C (note the minus sign!)
    • Result: E(r) = −(radial unit vector) × (1.44 kV/m)
  • This is large compared to typical engineering applications
  • Individual electron effects usually cancel because electrons are accompanied by equal positive charge (protons)
  • Exception: shot noise in practical electronics

🔗 Multiple charges and superposition

🔗 Superposition principle

The electric field from a set of charged particles equals the sum of the fields from individual particles.

Mathematically: E(r) = sum over n from 1 to N of E(r; rₙ)

  • N is the number of particles
  • This applies under usual assumptions about medium permittivity
  • Each particle contributes independently to the total field

🧮 Combined formula

E(r) = [1/(4 π ε)] × sum over n from 1 to N of [(r − rₙ) / |r − rₙ|³] × qₙ

  • r is the evaluation point
  • rₙ is the position of the nth particle
  • qₙ is the charge of the nth particle
  • The vector (r − rₙ) points from the charge to the evaluation point

📦 Charge distributions

📏 Line charge distribution

Line charge density ρₗ: the limit as segment length approaches zero of (charge / length) = dq/dl

  • Units: C/m
  • Describes charge distributed along a curve C
  • Total charge along the curve: Q = integral over C of ρₗ(l) dl
  • Key insight: line charge density integrated over length yields total charge

🎨 Surface charge distribution

Surface charge density ρₛ: the limit as patch area approaches zero of (charge / area) = dq/ds

  • Units: C/m²
  • Describes charge distributed over a surface
  • Total charge on surface S: Q = integral over S of ρₛ ds
  • Key insight: surface charge density integrated over area yields total charge

🧊 Volume charge distribution

Volume charge density ρᵥ: the limit as cell volume approaches zero of (charge / volume) = dq/dv

  • Units: C/m³
  • Describes charge distributed throughout a volume
  • Total charge in volume V: Q = integral over V of ρᵥ dv
  • Key insight: volume charge density integrated over volume yields total charge

🔄 From discrete to continuous

  • The smallest isolated charge is one electron (≈ −1.60 × 10⁻¹⁹ C)
  • This is very small; we rarely deal with electrons individually
  • More convenient to describe charge as continuous over a region
  • Three distribution types allow modeling realistic charge configurations
50

Electric Field Due to Point Charges

5.2 Electric Field Due to Point Charges

🧭 Overview

🧠 One-sentence thesis

The electric field from multiple charged particles is the vector sum of the fields from each individual particle, and this principle extends to continuous charge distributions by replacing summation with integration.

📌 Key points (3–5)

  • Single particle field: the electric field from one charged particle depends on the charge magnitude, distance, and direction from the particle to the observation point.
  • Superposition principle: the total electric field from multiple particles equals the sum of the individual fields.
  • Continuous distributions: charge can be described as distributed along a line, over a surface, or within a volume, each with its own density definition.
  • From discrete to continuous: summation over discrete particles becomes integration over continuous charge distributions.
  • Common confusion: charge density (C/m, C/m², or C/m³) vs. total charge (C)—density must be integrated over length/area/volume to yield total charge.

⚡ Electric field from point charges

⚡ Single particle at the origin

Electric field intensity from a single particle with charge q₁ at the origin: E(r) = r̂ · q₁ / (4πε r²)

  • The field points radially outward (for positive charge) along the unit vector r̂.
  • The strength falls off with the square of the distance r.
  • ε is the permittivity of the medium.

📍 Single particle at arbitrary position

When the particle is located at position r₁ instead of the origin, the expression becomes:

E(r; r₁) = (r − r₁) / |r − r₁|³ · q₁ / (4πε)

  • The numerator (r − r₁) is the vector from the source charge to the observation point.
  • The denominator |r − r₁|³ combines the distance-squared factor with the normalization.
  • Example: if a charge sits at position r₁ and you measure the field at position r, the field points along the direction from r₁ toward r (for positive charge).

🔁 Multiple particles via superposition

The excerpt states that under usual assumptions about permittivity, the superposition principle applies:

The electric field resulting from a set of charged particles is equal to the sum of the fields associated with the individual particles.

Mathematically: E(r) = sum from n=1 to N of E(r; rₙ)

Or explicitly: E(r) = (1 / 4πε) · sum from n=1 to N of [(r − rₙ) / |r − rₙ|³ · qₙ]

  • N is the number of particles.
  • Each particle contributes its own field vector; the total field is the vector sum.
  • Example: two charges at different positions produce fields that add vectorially at any observation point.

📏 Charge distributions

📏 Why continuous distributions?

The excerpt notes that the smallest isolated charge is approximately −1.60 × 10⁻¹⁹ C (one electron), which is very small. Since we rarely deal with electrons one at a time, it is more convenient to describe charge as continuous over a region of space.

Three types of distributions are defined: line, surface, and volume.

🧵 Line charge density

Line charge density ρₗ: the limit as segment length Δl approaches zero of Δq / Δl, equal to dq / dl.

  • Units: C/m (coulombs per meter).
  • Charge is distributed along a curve C.
  • Total charge along the curve: Q = integral over C of ρₗ(l) dl
  • Don't confuse: ρₗ is charge per unit length; integrating it over length gives total charge Q.

🗺️ Surface charge density

Surface charge density ρₛ: the limit as patch area Δs approaches zero of Δq / Δs, equal to dq / ds.

  • Units: C/m² (coulombs per square meter).
  • Charge is distributed over a surface.
  • Total charge over surface S: Q = integral over S of ρₛ ds
  • Example: a thin charged sheet has charge spread over its area; ρₛ describes how much charge per unit area.

🧊 Volume charge density

Volume charge density ρᵥ: the limit as cell volume Δv approaches zero of Δq / Δv, equal to dq / dv.

  • Units: C/m³ (coulombs per cubic meter).
  • Charge is distributed throughout a volume.
  • Total charge within volume V: Q = integral over V of ρᵥ dv
  • Example: a cloud of charged particles fills a region; ρᵥ describes charge per unit volume.
Distribution typeDensity symbolUnitsTotal charge formula
LineρₗC/mQ = ∫_C ρₗ dl
SurfaceρₛC/m²Q = ∫_S ρₛ ds
VolumeρᵥC/m³Q = ∫_V ρᵥ dv

🔄 From discrete particles to continuous distributions

🔄 The transition from sum to integral

The excerpt extends the discrete-particle formula to continuous distributions by replacing summation with integration.

For N discrete particles: E(r) = (1 / 4πε) · sum from n=1 to N of [(r − rₙ) / |r − rₙ|³ · qₙ]

For a continuous distribution along a curve C:

  • Divide the curve into short segments of length Δl.
  • Charge in the nth segment at position rₙ: qₙ = ρₗ(rₙ) Δl
  • Substitute into the sum: E(r) = (1 / 4πε) · sum from n=1 to N of [(r − rₙ) / |r − rₙ|³ · ρₗ(rₙ) Δl]
  • Take the limit as Δl → 0 to obtain the integral.

🧵 Electric field from line charge

The result for a continuous line distribution:

E(r) = (1 / 4πε) · integral over C of [(r − r′) / |r − r′|³ · ρₗ(r′) dl]

  • r′ represents the varying position along the curve C during integration.
  • The integral sums the contributions from each infinitesimal segment.
  • Example: a charged wire bent into a shape; the field at any point is the integral of contributions from each tiny piece of the wire.

🔍 Example: ring of charge

The excerpt includes an example of a ring of radius a in the z = 0 plane, centered on the origin, with uniform line charge density ρₗ. The goal is to find the electric field along the z axis.

  • Source position on the ring (in cylindrical coordinates): r′ = ρ̂ a
  • Field point on the z axis: r = ẑ z
  • Vector from source to field point: r − r′ = −ρ̂ a + ẑ z
  • Distance: |r − r′| = √(a² + z²)
  • The integral is set up using the formula above; the excerpt does not complete the calculation but shows the setup.
  • Don't confuse: the ring is in the z = 0 plane, but the field is evaluated along the z axis (off the plane).

📝 Note on methods

The excerpt mentions that for a straight line distribution, it is easier to use Gauss' Law (covered in Section 5.6) than to directly evaluate the integral. The ring example is given as a case where the integral approach is more appropriate.

51

Charge Distributions

5.3 Charge Distributions

🧭 Overview

🧠 One-sentence thesis

Charge distributions—line, surface, and volume—allow us to calculate total charge and electric fields by integrating charge density over the relevant geometric domain.

📌 Key points (3–5)

  • Three types of distributions: line (charge per length), surface (charge per area), and volume (charge per volume), each with its own density definition and units.
  • How to find total charge: integrate the appropriate charge density over the curve, surface, or volume.
  • Electric field from continuous distributions: extend the discrete-particle formula by replacing sums with integrals over the charge distribution.
  • Common confusion: the density function (ρ_l, ρ_s, or ρ_v) is defined at a point using a limit as the small element (Δl, Δs, or Δv) shrinks to zero; it is not simply "charge divided by total length/area/volume."
  • Why it matters: continuous distributions model realistic charge configurations (wires, sheets, volumes) and enable electric field calculations through integration.

📏 Three types of charge density

📏 Line charge density

Line charge density ρ_l at any point along a curve is defined as the limit of Δq/Δl as Δl approaches zero, equal to dq/dl.

  • Units: coulombs per meter (C/m).
  • Meaning: how much charge exists per unit length at a specific point on the curve.
  • The density can vary along the curve, so ρ_l is a function of position along the curve, parameterized by l: ρ_l(l).
  • Total charge along curve C: integrate ρ_l(l) over the curve: Q = integral over C of ρ_l(l) dl.
  • In other words, line charge density integrated over length yields total charge.

📐 Surface charge density

Surface charge density ρ_s at any point on a surface is defined as the limit of Δq/Δs as Δs approaches zero, equal to dq/ds.

  • Units: coulombs per square meter (C/m²).
  • Meaning: how much charge exists per unit area at a specific point on the surface.
  • The density can vary over the surface, so ρ_s is a function of position on the surface.
  • Total charge over surface S: integrate ρ_s over the surface: Q = integral over S of ρ_s ds.
  • In other words, surface charge density integrated over a surface yields total charge.

📦 Volume charge density

Volume charge density ρ_v at any point in a volume is defined as the limit of Δq/Δv as Δv approaches zero, equal to dq/dv.

  • Units: coulombs per cubic meter (C/m³).
  • Meaning: how much charge exists per unit volume at a specific point in the volume.
  • The density is a function of position within the volume.
  • Total charge within volume V: integrate ρ_v over the volume: Q = integral over V of ρ_v dv.
  • In other words, volume charge density integrated over a volume yields total charge.

🔍 Comparison table

Distribution typeDensity symbolUnitsSmall elementTotal charge formula
Lineρ_lC/mΔl (length)Q = ∫_C ρ_l dl
Surfaceρ_sC/m²Δs (area)Q = ∫_S ρ_s ds
Volumeρ_vC/m³Δv (volume)Q = ∫_V ρ_v dv

⚡ Electric field from continuous distributions

⚡ Starting point: discrete particles

  • The electric field from N discrete charged particles is given by a sum: E(r) = (1 / 4πε) times the sum from n=1 to N of [(r - r_n) / |r - r_n|³] times q_n.
  • Here q_n is the charge of the nth particle and r_n is its position.
  • However, many real situations involve continuous distributions of charge rather than countable particles.
  • The excerpt extends this formula to continuous distributions by replacing sums with integrals.

🔄 From sum to integral: the key idea

  • Divide the continuous distribution into small elements (segments, patches, or cells).
  • Express the charge in each element using the appropriate density: q_n = ρ(r_n) times the small element size (Δl, Δs, or Δv).
  • Substitute this into the discrete formula, then take the limit as the element size approaches zero.
  • The sum becomes an integral over the distribution.

🌀 Line distribution: electric field along a curve

🌀 Formula for line charge

  • Consider a continuous distribution of charge along a curve C.
  • Divide the curve into short segments of length Δl.
  • The charge in the nth segment at position r_n is: q_n = ρ_l(r_n) Δl.
  • Substitute into the discrete formula: E(r) = (1 / 4πε) times the sum of [(r - r_n) / |r - r_n|³] times ρ_l(r_n) Δl.
  • Taking the limit as Δl approaches zero yields: E(r) = (1 / 4πε) times the integral over C of [(r - r') / |r - r'|³] times ρ_l(r') dl.
  • Here r' represents the varying position along C during integration.

💍 Example: ring of uniform charge

  • Setup: a ring of radius a in the z=0 plane, centered on the origin, with uniform line charge density ρ_l (C/m). Find the electric field along the z axis.
  • Source position: in cylindrical coordinates, r' = ρ-hat times a (ρ-hat is the radial unit vector).
  • Field point: along the z axis, r = z-hat times z.
  • Difference vector: r - r' = -ρ-hat times a + z-hat times z.
  • Distance: |r - r'| = square root of (a² + z²).
  • Integral: E(z) = (1 / 4πε) times the integral from 0 to 2π of [(-ρ-hat a + z-hat z) / (a² + z²)^(3/2)] times ρ_l times (a dφ).
  • Simplification: pull out constants and separate into ρ-hat and z-hat components.
  • The integral of ρ-hat over 0 to 2π equals zero (by symmetry: ρ-hat(φ + π) = -ρ-hat(φ), so contributions cancel).
  • The integral of dφ from 0 to 2π equals 2π.
  • Result: E(z) = z-hat times (ρ_l a / 2ε) times [z / (a² + z²)^(3/2)].
  • Check: confirm dimensional correctness; when z is much greater than a, the result should approximate that of a point charge with the same total charge as the ring.

🔑 Symmetry argument

  • The excerpt highlights a symmetry argument: the integral of ρ-hat over all φ is zero because ρ-hat(φ + π) = -ρ-hat(φ).
  • For any given φ, the integrand is equal and opposite to the integrand π radians later, so contributions cancel.
  • This is one example of using symmetry to simplify integrals in electrostatics.

🏞️ Surface distribution: electric field from a surface

🏞️ Formula for surface charge

  • Consider a continuous distribution of charge over a surface S.
  • Divide the surface into small patches of area Δs.
  • The charge in the nth patch at position r_n is: q_n = ρ_s(r_n) Δs.
  • Substitute into the discrete formula: E(r) = (1 / 4πε) times the sum of [(r - r_n) / |r - r_n|³] times ρ_s(r_n) Δs.
  • Taking the limit as Δs approaches zero yields: E(r) = (1 / 4πε) times the integral over S of [(r - r') / |r - r'|³] times ρ_s(r') ds.
  • Here r' represents the varying position over S during integration.

💿 Example: disk of uniform charge

  • Setup: the excerpt mentions a circular disk of radius a in the z=0 plane, centered on the origin, as an example for surface charge distribution.
  • The excerpt does not provide the full solution, but the approach would follow the same pattern: set up the integral using the surface charge density formula, express r and r' in appropriate coordinates, and integrate over the disk.
  • (The excerpt cuts off before completing this example.)

🧮 Volume distribution: electric field from a volume

🧮 Implied formula for volume charge

  • Although the excerpt does not explicitly derive the electric field formula for volume charge distributions, the pattern is clear from the line and surface cases.
  • Divide the volume into small cells of volume Δv.
  • The charge in the nth cell at position r_n is: q_n = ρ_v(r_n) Δv.
  • Substitute into the discrete formula and take the limit as Δv approaches zero.
  • The result would be: E(r) = (1 / 4πε) times the integral over V of [(r - r') / |r - r'|³] times ρ_v(r') dv.
  • Here r' represents the varying position within V during integration.
52

Electric Field Due to a Continuous Distribution of Charge

5.4 Electric Field Due to a Continuous Distribution of Charge

🧭 Overview

🧠 One-sentence thesis

Continuous charge distributions require extending the discrete-charge electric field equation into integral form, where the field is computed by integrating over curves, surfaces, or volumes depending on how the charge is distributed.

📌 Key points (3–5)

  • Why continuous distributions matter: Real-world charge is often spread continuously along curves, over surfaces, or throughout volumes, not just in countable discrete particles.
  • Core method: Replace discrete sums with integrals by dividing the charge distribution into infinitesimal segments/patches/cells, then taking the limit as size approaches zero.
  • Three cases: Charge along a curve (line charge density ρₗ, C/m), over a surface (surface charge density ρₛ, C/m²), or in a volume (volume charge density ρᵥ, C/m³).
  • Symmetry arguments: When integrating, contributions from opposite directions often cancel (e.g., the radial component around a ring), simplifying calculations.
  • Common confusion: Don't confuse the three charge densities—each has different units and applies to different geometric distributions (curve vs surface vs volume).

🔄 From discrete to continuous charge

🔄 Why we need integrals

  • The discrete-charge formula (Equation 5.18) works for countable charged particles.
  • Real problems often involve charge spread continuously along wires, over plates, or throughout materials.
  • The excerpt extends the discrete formula by replacing sums with integrals.

🧮 The general approach

  1. Divide the continuous distribution into small pieces (length Δl, area Δs, or volume Δv).
  2. Express the charge in each piece using the appropriate charge density.
  3. Substitute into the discrete formula to get a sum.
  4. Take the limit as piece size → 0, converting the sum into an integral.

📏 Charge along a curve (line charge)

📏 Line charge density

Line charge density ρₗ: charge per unit length along a curve C, measured in C/m.

  • For a small segment of length Δl at position rₙ, the charge is qₙ = ρₗ(rₙ) Δl.
  • Substituting into the discrete formula and taking Δl → 0 yields the integral form.

🔁 The line-charge integral

The electric field at position r due to charge along curve C is:

  • E(r) = (1 / 4πε) times the integral over C of [(r − r′) / |r − r′|³] ρₗ(r′) dl
  • Here r′ is the varying position along C during integration.
  • dl is the differential length element along the curve.

💍 Example: Ring of uniform charge

Setup: A ring of radius a in the z = 0 plane, centered at the origin, with uniform line charge density ρₗ. Find the electric field along the z-axis.

Key steps:

  • Source position in cylindrical coordinates: r′ = ρ̂a (the radial unit vector times radius a).
  • Field point on z-axis: r = ẑz.
  • Distance: |r − r′| = √(a² + z²).
  • The integral over angle φ from 0 to 2π has two components: radial (ρ̂) and vertical (ẑ).

Symmetry argument:

  • The radial component integral equals zero because ρ̂(φ + π) = −ρ̂(φ).
  • For every radial direction, there is an opposite direction π radians later that cancels it.
  • Only the vertical (ẑ) component survives.

Result:

  • E(z) = ẑ (ρₗ a / 2ε) z / [a² + z²]^(3/2)
  • The field points along the z-axis, as expected from symmetry.
  • When z ≫ a (far from the ring), the result approximates a point charge with the same total charge.

🎨 Charge over a surface (surface charge)

🎨 Surface charge density

Surface charge density ρₛ: charge per unit area over a surface S, measured in C/m².

  • For a small patch of area Δs at position rₙ, the charge is qₙ = ρₛ(rₙ) Δs.
  • Taking Δs → 0 converts the sum into a surface integral.

🔁 The surface-charge integral

The electric field at position r due to charge over surface S is:

  • E(r) = (1 / 4πε) times the integral over S of [(r − r′) / |r − r′|³] ρₛ(r′) ds
  • r′ varies over the surface S.
  • ds is the differential area element.

💿 Example: Disk of uniform charge

Setup: A circular disk of radius a in the z = 0 plane, centered at the origin, with uniform surface charge density ρₛ. Find the electric field along the z-axis.

Key steps:

  • Source position in cylindrical coordinates: r′ = ρ̂ρ (radial distance ρ from center).
  • Field point on z-axis: r = ẑz.
  • Distance: |r − r′| = √(ρ² + z²).
  • Double integral over ρ (from 0 to a) and φ (from 0 to 2π).

Symmetry argument:

  • The integral over φ of the radial component (ρ̂) is zero.
  • As φ varies from 0 to 2π, ρ̂ rotates through a complete revolution.
  • Each pointing of ρ̂ is canceled by the opposite pointing, so only the vertical (ẑ) component remains.

Result:

  • E(z) = ẑ (ρₛ / 2ε) [sgn(z) − z / √(a² + z²)]
  • sgn(z) is the signum function: +1 for z > 0, −1 for z < 0.
  • The field direction depends on which side of the disk you are on.

🌐 Special case: Infinite sheet of charge

  • Let a → ∞ in the disk formula.
  • Result: E(r) = ẑ (ρₛ / 2ε) sgn(z)
  • The field is constant in magnitude, independent of distance from the sheet.
  • The field points away from the sheet on both sides (direction depends on sign of z).
  • This result applies at any point above or below the sheet, not just on the z-axis, due to symmetry (from any point, the edges are infinitely far away).

Dimensional check:

  • ρₛ has units C/m², ε has units F/m.
  • (C/m²) / (F/m) = (C/m²) / (C/V·m) = V/m, which is correct for electric field.

📦 Charge in a volume (volume charge)

📦 Volume charge density

Volume charge density ρᵥ: charge per unit volume within a region V, measured in C/m³.

  • For a small cell of volume Δv at position rₙ, the charge is qₙ = ρᵥ(rₙ) Δv.
  • Taking Δv → 0 converts the sum into a volume integral.

🔁 The volume-charge integral

The electric field at position r due to charge in volume V is:

  • E(r) = (1 / 4πε) times the integral over V of [(r − r′) / |r − r′|³] ρᵥ(r′) dv
  • r′ varies throughout the volume V.
  • dv is the differential volume element.

🔍 Comparison of the three cases

Distribution typeCharge densityUnitsDifferential elementIntegral domain
Line (curve)ρₗC/mdl (length)Curve C
SurfaceρₛC/m²ds (area)Surface S
VolumeρᵥC/m³dv (volume)Volume V

Don't confuse:

  • Each charge density applies to a different geometric scenario.
  • The units tell you which one to use: C/m for wires/curves, C/m² for sheets/surfaces, C/m³ for bulk materials/volumes.
  • The integral form is the same structure in all three cases, just with different density, differential element, and integration domain.

🎯 Practical tips from the examples

🎯 Symmetry simplifies integrals

  • Look for cancellations due to symmetry before computing integrals.
  • Example: radial components cancel when integrating around a full circle (0 to 2π).
  • This is called a symmetry argument and can save significant calculation effort.

🎯 Dimensional analysis

  • Always check that your final result has the correct units (V/m for electric field).
  • Example: In the infinite sheet case, (C/m²) / (F/m) = V/m ✓

🎯 Limiting cases

  • Verify that your result makes sense in extreme cases.
  • Example: For the ring, when z ≫ a, the field should approximate a point charge with total charge Q = ρₗ · 2πa.
  • Example: For the disk, when a → ∞, you get the infinite sheet result.

🎯 When to use this method vs Gauss's Law

  • The excerpt mentions that for a straight line of charge, Gauss's Law (Section 5.6) is much easier than the integral method.
  • The integral method (Equation 5.21 and similar) is more appropriate for distributions where Gauss's Law is harder to apply, such as rings and finite disks.
53

Gauss' Law: Integral Form

5.5 Gauss’ Law: Integral Form

🧭 Overview

🧠 One-sentence thesis

Gauss' Law provides an alternative method to Coulomb's Law for computing electric fields by relating the flux of the electric field through a closed surface to the enclosed charge, often simplifying calculations when combined with symmetry arguments.

📌 Key points (3–5)

  • What Gauss' Law states: the flux of the electric field through any closed surface equals the total charge enclosed by that surface.
  • Mathematical form: the surface integral of electric flux density D over a closed surface S equals the enclosed charge Q_encl.
  • Key advantage over Coulomb's Law: when combined with symmetry arguments, Gauss' Law can be simpler and more useful for certain charge distributions.
  • How symmetry helps: if the charge distribution has symmetry (spherical, cylindrical, etc.), you can constrain the form of the electric field before solving, making the integral tractable.
  • Common confusion: Gauss' Law applies to any closed surface enclosing the charge, not just surfaces that match the charge distribution's shape—but choosing a symmetric surface makes the calculation easier.

⚡ What Gauss' Law says

📐 The fundamental statement

Gauss' Law: The flux of the electric field through a closed surface is equal to the enclosed charge.

  • Gauss' Law is one of the four fundamental laws of classical electromagnetics (Maxwell's Equations).
  • It connects a surface property (flux through a boundary) to a volume property (charge inside).

🔢 Mathematical expression

The law is written as:

Surface integral of D · ds = Q_encl

Where:

  • D is the electric flux density (equal to permittivity ε times electric field E)
  • S is a closed surface
  • ds is the differential surface normal (outward-facing)
  • Q_encl is the total charge enclosed by the surface

✅ Dimensional check

  • D has units of C/m² (charge per area)
  • Integrating D over a surface: C/m² · m² = C (charge units)
  • This matches the units of Q_encl, confirming dimensional correctness

🔄 How to use Gauss' Law

🎯 The basic strategy

  1. Choose a closed surface that encloses the charge distribution
  2. Use symmetry arguments to constrain the form of D (magnitude and direction)
  3. Evaluate the surface integral with the simplified form
  4. Solve for D, then find E using D = ε E

🔍 Why symmetry matters

  • The charge distribution's symmetry determines which surfaces make the calculation simple
  • If the magnitude of D depends only on one coordinate (e.g., radius r) and not others (e.g., angles θ, φ), you can pull it out of the integral
  • The direction of D can often be determined by symmetry alone (e.g., must point radially for a spherical charge)

Don't confuse: You can use any closed surface for Gauss' Law—the law is always true—but only symmetric surfaces make the integral easy to evaluate.

🧮 The role of symmetry arguments

From the excerpt's example:

  • For a point charge at the origin, the magnitude of D can depend only on r, not on angles θ or φ
  • The field must point either toward or away from the charge (radial direction)
  • This means D = r̂ D(r), where is the radial unit vector
  • These constraints come from the fact that "the charge has no particular orientation"

🌟 Example: Point charge

🎯 Problem setup

  • A particle of charge q is located at the origin
  • Goal: find the electric field using Gauss' Law
  • This is the simplest possible charge distribution

🔧 Solution steps

Step 1: Choose the surface

  • Use a sphere of radius r centered at the origin
  • For any r > 0, the enclosed charge Q_encl = q

Step 2: Apply symmetry

  • The magnitude of D depends only on r (not θ or φ)
  • The direction is radial: D = r̂ D(r)

Step 3: Evaluate the integral

  • Gauss' Law becomes: integral over θ from 0 to π and φ from 0 to 2π of D · (r̂ r² sin θ dθ dφ) = q
  • The dot product r̂ · r̂ = 1
  • Since D(r) and are constants with respect to integration, factor them out: r² D(r) · (integral of sin θ dθ dφ) = q
  • The remaining integral equals

Step 4: Solve for D and E

  • D(r) = q / (4π r²)
  • Including direction: D = r̂ q / (4π r²)
  • Using D = ε E: E = r̂ q / (4π ε r²)

This matches the known result from Coulomb's Law.

💡 Key takeaway

Gauss' Law combined with a symmetry argument may be sufficient to determine the electric field due to a charge distribution. Thus, Gauss' Law may be an easier alternative to Coulomb's Law in some applications.

🔌 Example: Infinite line charge

🎯 Problem setup

  • An infinite line of charge along the z axis
  • Charge density ρ_l (units of C/m)
  • Goal: find the electric field intensity using Gauss' Law

🔧 Solution approach

Step 1: Choose the surface

  • Use a cylinder of radius a concentric with the z axis
  • The cylinder is "maximally symmetric with the charge distribution"
  • Initial concern: the charge extends to infinity in both ±z directions
  • Strategy: start with a cylinder of finite length l, then let l → ∞ if needed

Step 2: Symmetry argument

  • The excerpt begins to develop a symmetry argument by considering how to "assemble" an infinite line from point charges added in pairs
  • (The excerpt cuts off before completing the solution)

🎓 Why this approach works

  • Although the problem can be solved using the direct approach from earlier sections (integrating Coulomb's Law), the Gauss' Law approach "turns out to be relatively simple"
  • The cylindrical surface matches the cylindrical symmetry of the line charge
  • This is a useful "building block" for other problems (e.g., capacitance of coaxial cable)

🆚 Gauss' Law vs. Coulomb's Law

AspectCoulomb's Law approachGauss' Law approach
MethodDirect integration of contributions from charge elementsSurface integral of flux equals enclosed charge
When easierGeneral distributions without symmetryDistributions with high symmetry (spherical, cylindrical, planar)
RequirementsNeed to sum/integrate over all charge elementsNeed to choose appropriate closed surface and apply symmetry arguments
Described inSections 5.1, 5.2, and 5.4Section 5.5 and beyond

Don't confuse: Both methods give the same answer for the electric field; Gauss' Law is not a different physics, just a different mathematical route that exploits symmetry.

54

Electric Field Due to an Infinite Line Charge using Gauss' Law

5.6 Electric Field Due to an Infinite Line Charge using Gauss’ Law

🧭 Overview

🧠 One-sentence thesis

Gauss' Law combined with symmetry arguments provides a simpler alternative to Coulomb's Law for finding the electric field of an infinite line charge, yielding a field that points radially outward and decreases inversely with distance.

📌 Key points (3–5)

  • Why use Gauss' Law: It can be easier than Coulomb's Law when the charge distribution has sufficient symmetry.
  • The symmetry argument: By analyzing how particles combine, we determine that the electric field must point radially outward (in the ρ-hat direction) and cannot depend on φ or z.
  • The Gaussian surface choice: A finite cylinder concentric with the line charge encloses only part of the charge, but remarkably the result is independent of cylinder length.
  • Common confusion: The infinite charge seems impossible to enclose, but the finite-cylinder approach works because the length cancels out in the calculation.
  • The result: The electric field points radially away from the line and has magnitude proportional to charge density divided by distance (ρ_l / 2περ).

🎯 Why Gauss' Law for this problem

🎯 Alternative to direct integration

  • The excerpt states this problem can be solved using the "direct approach" from Coulomb's Law (Section 5.4).
  • However, the Gauss' Law approach "turns out to be relatively simple."
  • The key advantage: symmetry makes the integration trivial.

🧱 Building block for other problems

  • The result serves as a "useful building block" for other applications.
  • Example from the excerpt: determining the capacitance of coaxial cable.

🔍 Setting up the problem

🔍 The charge distribution

  • An infinite line of charge along the z-axis.
  • Charge density: ρ_l (units of coulombs per meter).
  • Extends to infinity in both +z and −z directions.

🛠️ Choosing the Gaussian surface

  • A cylinder of radius a, concentric with the z-axis.
  • Why a cylinder? It is "maximally symmetric with the charge distribution" and yields the simplest analysis.
  • Initial concern: the charge extends to infinity, so how do we enclose all of it?
  • Solution strategy: use a cylinder of finite length l, solve for that fraction of charge, and prepare to let l → ∞ if needed.

🧮 The symmetry argument

🧮 Building the line from point charges

The excerpt uses a clever construction:

  • Start with the field of a point charge q at the origin: D = r-hat · q / (4πr²).
  • "Assemble" an infinite line by adding particles in pairs: one at +z, one at −z, each equidistant from the origin.
  • Continue adding pairs until the charge extends continuously to infinity in both directions.
  • By superposition, the resulting field is the sum of all constituent particle fields.

🔒 What symmetry tells us

From this construction, we deduce three constraints:

ConstraintReason
No φ-hat componentNone of the constituent particle fields have a φ-hat component
No dependence on φThe charge distribution is identical (invariant) under rotation in φ
No z-hat component and no z-dependenceFor any z, the charge distribution above and below is identical

Conclusion: D must point radially outward in the ρ-hat direction only.

D = ρ-hat · D_ρ(ρ)

  • The magnitude D_ρ can depend only on ρ (the radial distance from the z-axis).
  • Don't confuse: ρ here is the cylindrical radial coordinate, not the charge density ρ_l.

🧪 Applying Gauss' Law

🧪 The integral form

Gauss' Law states:

∮_S D · ds = Q_encl

  • S is a closed surface with outward-facing differential surface normal ds.
  • Q_encl is the enclosed charge.
  • For our finite cylinder of length l: Q_encl = ρ_l · l.

🧪 Breaking down the cylinder surface

The cylinder S has three parts: flat top, curved side, flat bottom.

The integral becomes:

  • ρ_l · l = (integral over top) + (integral over side) + (integral over bottom)

Examining the dot products:

  • Top surface: D is ρ-hat, surface normal is +z-hat → dot product is zero (perpendicular).
  • Bottom surface: D is ρ-hat, surface normal is −z-hat → dot product is zero (perpendicular).
  • Side surface: D is ρ-hat, surface normal is +ρ-hat → dot product is 1.

Result: Only the side surface contributes flux.

🧪 Solving for D

After eliminating top and bottom:

  • ρ_l · l = ∫_side [D_ρ(a)] ds

On the side surface, ρ = a (constant), so D_ρ(a) is constant and can be pulled out:

  • ρ_l · l = D_ρ(a) · ∫_side ds

The remaining integral is the area of the side surface: 2πa · l.

Solving:

  • D_ρ(a) = (ρ_l · l) / (2πa · l) = ρ_l / (2πa)

🎉 The remarkable cancellation

The excerpt emphasizes: "Remarkably, we see D_ρ(a) is independent of l."

  • The initial concern about not enclosing all the charge "doesn't matter."
  • The length l cancels out completely.
  • This means we don't need to take the limit l → ∞ after all.

📐 The final result

📐 Electric flux density

Since the result must be the same for any value of ρ (not just ρ = a):

D = ρ-hat · ρ_l / (2πρ)

  • Points radially outward from the line charge.
  • Magnitude decreases inversely with distance from the line.

📐 Electric field intensity

Using the relationship D = ε · E:

E = ρ-hat · ρ_l / (2περ)

  • ε is the permittivity of the surrounding medium.
  • The field is directed radially away from the line charge.
  • Magnitude is inversely proportional to distance ρ.

✅ Dimensional check

The excerpt suggests: "Check to ensure that this solution is dimensionally correct."

  • ρ_l has units C/m (charge per length).
  • ρ has units m (distance).
  • ε has units related to C²/(N·m²).
  • The combination should yield units of electric field (N/C or V/m).

🔄 Comparison: Integral vs Differential Gauss' Law

🔄 When the integral form works

The excerpt (Section 5.7 preview) explains limitations:

  • The integral form (∮_S D · ds = Q_encl) works well when symmetry permits.
  • Examples: point charge (Section 5.5) and infinite line charge (Section 5.6).

🔄 When you need the differential form

The excerpt introduces a key limitation of the integral approach:

  • It "does not account for the presence of structures that may influence the electric field."
  • Example: a charge near a perfectly-conducting surface produces a different field than the same charge in free space.
  • These approaches "do not account for the possibility of any spatial variation in material composition."

🔄 The differential form

The excerpt derives:

∇ · D = ρ_v

  • ∇ · D is the divergence of D (flux per unit volume).
  • ρ_v is the volume charge density at a point.
  • This form applies at individual points in space, not over a whole surface.

Interpretation: "Gauss' Law in differential form says that the electric flux per unit volume originating from [a point equals the charge density there]."

Don't confuse:

  • Integral form: relates total flux through a closed surface to total enclosed charge; requires symmetry for practical use.
  • Differential form: relates flux density at a point to charge density at that point; works even with material boundaries and spatial variations.
55

5.7 Gauss' Law: Differential Form

5.7 Gauss’ Law: Differential Form

🧭 Overview

🧠 One-sentence thesis

Gauss' Law in differential form enables the calculation of electric fields at individual points in space—even in problems with material boundaries and spatial variations—by relating the divergence of electric flux density to volume charge density at each point.

📌 Key points (3–5)

  • What the differential form does: converts Gauss' Law from an integral over a closed surface into a point-by-point differential equation that applies at every location in space.
  • Why it matters: the integral form requires high symmetry and cannot handle material boundaries or spatial variations in material properties; the differential form addresses these broader engineering problems.
  • The core equation: divergence of electric flux density D equals volume charge density ρ_v at that point.
  • Common confusion: the integral form calculates enclosed charge from surrounding flux; the differential form does the inverse—it determines the electric field from a given charge distribution, especially when boundary conditions from materials and structures are present.
  • Fundamental insight: Gauss' Law (in either form) is fundamental; Coulomb's Law is merely a consequence of Gauss' Law for the special case of charge in a uniform, unbounded medium with no boundary conditions.

🔄 From integral to differential

🔄 The limitation of the integral form

The integral form of Gauss' Law is:

Surface integral of D · ds over a closed surface S equals the enclosed charge Q_encl.

  • This form works well when the problem has sufficient symmetry (as shown in earlier sections).
  • Two critical limitations:
    • It cannot handle problems lacking the necessary symmetry.
    • It does not account for structures that influence the electric field (e.g., a charge near a perfectly-conducting surface) or any spatial variation in material composition.
  • Even the Coulomb's Law / direct integration approach (from Section 5.4) shares this limitation: it assumes uniform, unbounded media and no material boundaries.
  • Example: the electric field due to a charge in free space is different from the field due to the same charge located near a perfectly-conducting surface.

🎯 What the differential form provides

  • An alternative form that applies at individual points in space, not over a closed surface.
  • Expressed as a differential equation (not an integral equation).
  • Facilitates solving for electric fields in problems that:
    • Do not exhibit sufficient symmetry.
    • Involve material boundaries.
    • Involve spatial variations in material constitutive parameters.
  • Given the differential equation and boundary conditions imposed by structure and materials, one can solve for the electric field in these more complicated scenarios.

🧮 Deriving the differential form

🧮 Method 1: via the definition of divergence

Start with the integral form and divide both sides by the enclosed volume V:

  • Take the limit as V → 0.
  • The right-hand side becomes the volume charge density ρ_v (units of C/m³) at the point where the volume converges to zero.
  • The left-hand side is, by definition, the divergence of D, written as "∇ · D" (see Section 4.6).
  • Result:

    Gauss' Law in differential form: ∇ · D = ρ_v

🧮 Method 2: via the divergence theorem

  • Apply the Divergence Theorem (Section 4.7): the volume integral of (∇ · D) over V equals the surface integral of D · ds over S.
  • From the integral form, the surface integral equals the enclosed charge Q_encl.
  • Express Q_encl as the volume integral of ρ_v over V.
  • Now both sides are volume integrals: the integral of (∇ · D) equals the integral of ρ_v.
  • This relationship must hold regardless of the specific location or shape of V.
  • The only way this is possible is if the integrands are equal: ∇ · D = ρ_v.

🔍 Why two methods?

The excerpt notes that both methods are presented because each provides a different bit of insight:

  • The first method emphasizes the physical meaning of divergence (flux per unit volume).
  • The second method uses a mathematical theorem to connect the integral and differential forms rigorously.

🧠 Interpreting the differential form

🧠 Physical meaning

Gauss' Law in differential form says that the electric flux per unit volume originating from a point in space is equal to the volume charge density at that point.

  • Recall that divergence is simply the flux (in this case, electric flux) per unit volume.
  • The equation ∇ · D = ρ_v directly links the local "source strength" (charge density) to the local "flux generation" (divergence of D).

📐 Example: determining charge density from electric field

Given: The electric field intensity in free space is E(r) = x-hat · A·x² + y-hat · B·z + z-hat · C·x²·z, where A = 3 V/m³, B = 2 V/m², and C = 1 V/m⁴. Find the charge density at r = x-hat · 2 − y-hat · 2 m.

Solution steps:

  1. Use D = ε·E to get D. Since the problem is in free space, ε = ε₀.
  2. The volume charge density is ρ_v = ∇ · D = ∇ · (ε₀·E) = ε₀·(∇ · E).
  3. Calculate the divergence: ∇ · E = ∂/∂x (A·x²) + ∂/∂y (B·z) + ∂/∂z (C·x²·z) = 2·A·x + 0 + C·x².
  4. At the specified location r: ε₀ · [2·(3 V/m³)·(2 m) + 0 + (1 V/m⁴)·(2 m)²] = ε₀ · (16 V/m) = 142 pC/m³.

Don't confuse: This example goes from field to charge density. The more common engineering problem is the inverse: finding the field from a given charge distribution with boundary conditions.

🛠️ Using the differential form in practice

🛠️ Boundary conditions and solving for electric field

  • To obtain the electric field from the charge distribution in the presence of boundary conditions imposed by materials and structure, one must enforce the relevant boundary conditions.
  • These boundary conditions are presented in Sections 5.17 and 5.18.
  • A simpler approach requiring only the boundary conditions on the electric potential V(r) is often possible; this is presented in Section 5.15.

⚠️ Additional constraint: Kirchhoff's Voltage Law

  • Gauss' Law does not always fully constrain possible solutions for the electric field.
  • For that, one might also need Kirchhoff's Voltage Law (see Section 5.11).

🔓 Special case: no boundary conditions

In the special case where there are no boundary conditions to satisfy (i.e., for charge only, in a uniform and unbounded medium), the differential form can be solved directly:

D(r) = (1 / 4π) · integral over V of [(r − r′) / |r − r′|³] · ρ_v(r′) dv

  • This is one of the results obtained in Section 5.4 (after dividing both sides by ε to get E).
  • Fundamental conclusion: It is reasonable to conclude that Gauss' Law (in either integral or differential form) is fundamental, whereas Coulomb's Law is merely a consequence of Gauss' Law.

🔗 Connections and context

🔗 Relationship to earlier results

  • The excerpt includes a preceding example (finding the electric field of an infinite line of charge using Gauss' Law) that demonstrates the integral form's power when symmetry is present.
  • The result for that example: the electric field is directed radially away from the line charge and decreases in magnitude in inverse proportion to distance from the line charge.
  • The differential form extends this capability to cases where such symmetry is absent.

🔗 Broader scope of problems

The differential form addresses a broader scope of problems by:

  • Allowing spatial variation in material composition.
  • Accounting for the presence of structures that influence the electric field (e.g., perfectly-conducting surfaces).
  • Enabling the use of Gauss' Law even in problems that do not exhibit sufficient symmetry.
56

Force, Energy, and Potential Difference

5.8 Force, Energy, and Potential Difference

🧭 Overview

🧠 One-sentence thesis

The potential difference (voltage) between two points in an electric field depends only on the start and end positions, not on the path taken, because it represents the change in potential energy per unit charge.

📌 Key points (3–5)

  • Force and work: A charged particle in an electric field experiences a force; moving the particle changes the system's potential energy, quantified as work.
  • Potential difference definition: Voltage between two points is the work done per unit charge, measured in volts (J/C).
  • Path independence: The potential difference depends only on the start and end points, not on the path taken between them.
  • Common confusion: The sign of work—negative work means the system "relaxes" (loses potential energy); positive work means an external force compresses the system (increases potential energy).
  • Why it matters: Potential difference allows analysis without explicitly stating the charge involved, simplifying electromagnetic problems.

⚡ Force and work in electric fields

⚡ Force on a charged particle

The force F_e experienced by a particle at location r bearing charge q in an electric field intensity E is: F_e = q E(r).

  • A particle left alone in free space immediately begins to move under this force.
  • The resulting displacement represents a loss of potential energy.

🔧 Work and incremental displacement

The incremental work ΔW done by moving the particle a short distance Δl is: ΔW ≈ −F_e · l̂ Δl, where is the unit vector in the direction of motion.

  • The minus sign indicates that the potential energy of the system (electric field + particle) is being reduced.
  • The system is "relaxing," like a compressed spring being released.
  • Units: force (N) times distance (m) gives N·m, which equals 1 joule (J) of energy.

🔄 External force and positive work

  • If an external force is applied to overcome F_e, the unit vector changes sign.
  • The same magnitude of work is done, but now it is positive.
  • Positive work means the external force opposes and overcomes the electric field force, increasing the system's potential energy.
  • Example: Positive work is like compressing a spring; negative work is like releasing it.

📐 The role of the dot product

  • The dot product in the work equation ensures that only the component of motion parallel to the electric field direction is counted.
  • Motion in any other direction cannot be due to E and does not change the associated potential energy.

🧮 Generalizing work over longer paths

🧮 Accounting for varying electric fields

  • For short distances, ΔW ≈ −q E(r) · l̂ Δl.
  • For larger distances, the electric field E may vary along the path.
  • Sum contributions from points along the path: W ≈ Σ ΔW(r_n), where r_n are positions defining the path.
  • Substituting the incremental work: W ≈ −q Σ E(r_n) · l̂(r_n) Δl.

∫ Taking the limit to an integral

  • As Δl → 0, the sum becomes an integral:

    W = −q ∫_C E(r) · l̂(r) dl, where C is the path followed.

  • Simplified notation: W = −q ∫_C E · dl, where dl = l̂ dl as usual.
  • This formula determines the change in potential energy for a charged particle moving along any path in space, given the electric field.

🔋 Potential difference (voltage)

🔋 Definition of potential difference

The electric potential difference V_21 between the start point (1) and end point (2) of path C is defined as the work done per unit charge: V_21 = W / q.

  • Units: joules per coulomb (J/C), which is volts (V).
  • Substituting the work formula: V_21 = −∫_C E · dl.
  • Advantage: Analysis in terms of potential no longer requires explicitly stating the charge involved.

📏 Uniform electric field example

  • Consider a uniform electric field E(r) = ẑ E_0 (constant magnitude and direction).
  • Path: a line from ẑ z_1 to ẑ z_2.
  • From the potential difference formula: V_21 = −∫_{z_1}^{z_2} (ẑ E_0) · ẑ dz = −E_0 (z_2 − z_1).
  • Result: V_21 is simply the electric field intensity times the distance between the points.
  • This agrees with findings from battery-charged capacitor experiments.

🧭 Path orientation and the dot product

  • If the direct path between two points is parallel to E, the solution is straightforward (as in the example above).
  • If the path is perpendicular to E, then V_21 = 0 (the dot product is zero).
  • In all other cases, V_21 is proportional to the component of the direct path that is parallel to E, as determined by the dot product.

🛤️ Path independence

🛤️ Why potential difference is path-independent

  • The potential difference is: V_21 = −∫_{r_1}^{r_2, along C} E · dl, where r_1 and r_2 are the start and end position vectors.
  • The work done by a particle with charge q is: W_21 = q V_21.
  • This work represents the change in potential energy: W_21 = W_2 − W_1, where W_2 and W_1 are the potential energies at r_2 and r_1.
  • Because W_21 depends only on W_2 and W_1, it does not depend on the path C—only on the start and end positions.

🔁 Independence of path principle

V_21 = −∫_{r_1}^{r_2} E · dl, independent of C.

  • Any path that begins at r_1 and ends at r_2 yields the same value of W_21 and V_21.
  • This concept is called independence of path.
  • Don't confuse: The integral is taken over a path, but the result does not depend on which path is chosen—only on the endpoints.
57

Independence of Path

5.9 Independence of Path

🧭 Overview

🧠 One-sentence thesis

The potential difference (voltage) between two points in an electric field depends only on the start and end positions, not on the path taken between them, which leads directly to Kirchhoff's Voltage Law for electrostatics.

📌 Key points (3–5)

  • Core claim: The line integral of the electric field between two points yields the same voltage regardless of which path is chosen, because voltage depends only on the difference in potential energy at the endpoints.
  • Why path independence holds: Work done by a charged particle equals the change in potential energy (W₂ − W₁), which is determined solely by the positions of the start and end points.
  • Practical advantage: You may choose any convenient path (e.g., one parallel or perpendicular to the field) to simplify the integral calculation, even if it differs from the actual path traveled.
  • Common confusion: The integral notation shows a path C, but the result does not depend on C—only on the endpoints r₁ and r₂.
  • Consequence: When the path is a closed loop (start = end), the integral of the electric field is zero, which is Kirchhoff's Voltage Law for electrostatics.

⚡ Voltage and work in an electric field

⚡ Potential difference formula

The potential difference (voltage) V₂₁ associated with a path C in an electric field intensity E is the negative line integral of E along that path: V₂₁ = − ∫ E · dl from point 1 to point 2.

  • The curve begins at position r₁ and ends at position r₂.
  • The dot product E · dl picks out the component of the field parallel to the path element.
  • Example: If the path is perpendicular to E, then E · dl = 0 and V₂₁ = 0; if parallel, V₂₁ is proportional to the field intensity times distance.

⚙️ Work and potential energy

  • The work done by a particle bearing charge q is W₂₁ = q V₂₁.
  • This work represents the change in potential energy of the system (field + particle): W₂₁ = W₂ − W₁.
  • W₂ and W₁ are the potential energies when the particle is at r₂ and r₁, respectively.
  • Because W₂₁ is a difference of two values tied to positions, it cannot depend on the intermediate points along C.

🛤️ The independence-of-path principle

🛤️ What independence of path means

Independence of path: The integral of the electric field over a path between two points depends only on the locations of the start and end points and is independent of the path taken between those points.

  • Formally: V₂₁ = − ∫ E · dl from r₁ to r₂, independent of C.
  • Any path that begins at r₁ and ends at r₂ yields the same value of W₂₁ and V₂₁.
  • Don't confuse: The integral is written "along C," but the result does not vary with the choice of C.

🧰 Practical use

  • Some paths may be easier to integrate than others.
  • You may compute the integral using a convenient path (e.g., one aligned with the field or broken into perpendicular segments) even if the particle actually traveled a different route.
  • Example: If the actual path is curved, you might instead integrate along a straight line or a right-angle path to simplify the calculation.

🔁 Kirchhoff's Voltage Law for electrostatics

🔁 Closed-loop case

  • Consider a path that begins and ends at the same point: r₂ = r₁.
  • The path of integration is now a closed loop.
  • Since V₂₁ depends only on the positions of the start and end points, and the potential energy at those points is the same, the integral must be zero.

Kirchhoff's Voltage Law for Electrostatics: The integral of the electric field over a closed path is zero: ∮ E · dl = 0.

🔌 Connection to circuit theory

  • In electric circuit theory, the sum of voltages over any closed loop in a circuit is zero—also called Kirchhoff's Voltage Law.
  • This is the same principle applied to circuits:
    • Partition the closed path into branches (each representing one component).
    • The integral of E over each branch is the branch voltage (units: V/m × m = V).
    • The sum of these branch voltages over any closed loop is zero, as dictated by the closed-loop integral.
  • Don't confuse: The circuit version is a special case of the general electrostatic principle; both stem from path independence.
58

Kirchoff's Voltage Law for Electrostatics: Integral Form

5.10 Kirchoff’s Voltage Law for Electrostatics: Integral Form

🧭 Overview

🧠 One-sentence thesis

Kirchoff's Voltage Law for Electrostatics states that the integral of the electric field around any closed path is zero, which generalizes the familiar circuit-theory principle that voltages sum to zero around a loop.

📌 Key points

  • Core principle: When you integrate the electric field around a closed path (starting and ending at the same point), the result is always zero.
  • Why it's true: Electrical potential depends only on start and end positions; if they're the same point, the potential difference is zero.
  • Connection to circuits: This is the same Kirchoff's Voltage Law from circuit theory—each branch integral gives a voltage, and they sum to zero around any loop.
  • Common confusion: This law applies only to electrostatics (static or non-time-varying magnetic fields); if the magnetic field varies with time, the right-hand side is no longer zero and the law must be modified (Maxwell-Faraday Equation).
  • Practical advantage: Because the integral is path-independent, you can choose any convenient path between two points to compute potential.

🔄 The closed-loop principle

🔄 What happens when start equals end

  • The excerpt begins with the general formula for potential difference between two points r₁ and r₂: V₂₁ equals the negative integral of E (electric field) dot dl along any path from r₁ to r₂.
  • Now consider a special case: the path forms a closed loop, so r₂ = r₁ (start and end are the same point).
  • Because potential depends only on position, the potential energy at the start and end is identical.
  • Therefore, the potential difference V₂₁ must be zero.

⚡ The integral form of Kirchoff's Voltage Law

Kirchoff's Voltage Law for Electrostatics: The integral of the electric field over a closed path is zero.

  • Mathematically: the closed-loop integral of E dot dl equals 0.
  • The circle on the integral symbol indicates a closed path.
  • This is the defining statement of the law.

🔌 Connection to electric circuits

🔌 How circuit theory emerges

  • The excerpt notes that readers likely already know: in circuit theory, the sum of voltages around any closed loop is zero.
  • This circuit law is "precisely the same principle" as the electrostatic version.

🧩 From field integral to branch voltages

  • To see the connection, partition the closed path into branches (each representing one circuit component).
  • The integral of E over each branch gives the branch voltage.
    • Units check: (V/m) times (m) yields (V).
  • Summing these branch voltages over the closed loop gives zero, exactly as the integral form dictates.
  • Example: A loop through a battery and resistor—integrate E along each segment; the sum is zero.

⚠️ Limitations and scope

⚠️ Electrostatics only

  • The excerpt emphasizes: Equation 5.87 (the zero integral) is specific to electrostatics.
  • Electrostatics assumption: The electric field is independent of the magnetic field.
  • This holds true if:
    • The magnetic field is zero, or
    • The magnetic field is not time-varying.

🔁 What happens with time-varying magnetic fields

  • If the magnetic field varies with time, the right-hand side of the equation is no longer zero.
  • The law must be modified to account for the magnetic field's effect.
  • The generalized version is called the Maxwell-Faraday Equation (covered in Section 8.8).
  • Don't confuse: Kirchoff's Voltage Law for electrostatics is a special case; the full Maxwell-Faraday Equation applies when magnetic fields change over time.

🛤️ Path independence and practical use

🛤️ Why path choice matters

  • The excerpt recalls (from Section 5.9) that potential difference depends only on the start and end locations, not on the path taken.
  • This path independence is the reason the closed-loop integral is zero.

🧰 Choosing convenient paths

  • A practical consequence: some paths may be easier to compute than others.
  • You can compute the integral using any path between the specified points, not necessarily the actual physical path.
  • Example: If the actual path is curved and complicated, you might choose a straight line or a path aligned with coordinate axes to simplify the calculation.
59

Kirchoff's Voltage Law for Electrostatics: Differential Form

5.11 Kirchoff’s Voltage Law for Electrostatics: Differential Form

🧭 Overview

🧠 One-sentence thesis

The differential form of Kirchoff's Voltage Law for electrostatics states that the curl of the electrostatic field is zero, providing a partial differential equation that can be used to solve for the electric field in complex scenarios.

📌 Key points

  • Integral form: the line integral of the electric field along any closed path equals zero.
  • Differential form: the curl of the electrostatic field is zero (∇ × E = 0).
  • Derivation method: Stokes' Theorem transforms the integral form into the differential form.
  • Common confusion: this equation is specific to electrostatics (magnetic field is zero or not time-varying); if the magnetic field is time-varying, the right-hand side is no longer zero and the Maxwell-Faraday Equation must be used instead.
  • Why it matters: combined with boundary conditions, this partial differential equation can solve for the electric field in arbitrarily complicated scenarios, though it does not fully constrain the field alone.

🔄 From integral to differential form

🔄 The integral form (starting point)

The integral form of Kirchoff's Voltage Law for electrostatics states that an integral of the electric field along a closed path is equal to zero.

  • Written as: the line integral of E · dl around closed curve C equals 0.
  • E is electric field intensity.
  • C is any closed curve.
  • This is the starting point for deriving the differential form.

🔧 Applying Stokes' Theorem

  • Stokes' Theorem connects a surface integral to a line integral around the boundary of that surface.
  • In this case: the surface integral of (∇ × E) · ds over surface S equals the line integral of E · dl around closed curve C.
  • S is any surface bounded by C.
  • ds is the normal to that surface with direction determined by right-hand rule.

⚡ The key insight

  • Since the integral form tells us the line integral (right-hand side) is zero, the surface integral of (∇ × E) · ds must also be zero.
  • This relationship must hold regardless of the specific location or shape of S.
  • The only way this is possible for all possible surfaces is if the integrand itself is zero at every point in space.
  • Therefore: ∇ × E = 0.

📐 The differential form and its meaning

📐 What the differential form says

The differential form of Kirchoff's Voltage Law for electrostatics states that the curl of the electrostatic field is zero.

  • Written as: ∇ × E = 0.
  • This is a partial differential equation.
  • It describes a local property of the electric field at every point in space.

🛠️ How it is used

  • Combined with appropriate boundary conditions (imposed by structure and materials), this equation can be solved for the electric field in arbitrarily-complicated scenarios.
  • It provides insight into the behavior of the electric field.
  • Example: knowing the curl is zero tells us the electrostatic field is conservative (though the excerpt does not state this explicitly).

🔍 Relationship to other equations

  • This is not the only partial differential equation available for solving for the electric field.
  • Gauss' Law also serves this purpose.
  • A system of partial differential equations is emerging.
  • The electric field is not necessarily fully constrained by either equation alone—both are needed together.

⚠️ Limitations and scope

⚠️ Electrostatics only

  • The excerpt emphasizes that this equation is specific to electrostatics.
  • In electrostatics, the electric field is assumed to be independent of the magnetic field.
  • This is true if the magnetic field is either zero or not time-varying.

🔄 When the magnetic field is time-varying

  • If the magnetic field is time-varying, the equation must be modified.
  • The effect of the time-varying magnetic field makes the right-hand side potentially different from zero.
  • The generalized version that correctly accounts for this effect is known as the Maxwell-Faraday Equation.
  • Don't confuse: the electrostatic form (curl of E equals zero) applies only when the magnetic field is absent or constant; the Maxwell-Faraday Equation is the more general form for time-varying fields.
60

Electric Potential Field Due to Point Charges

5.12 Electric Potential Field Due to Point Charges

🧭 Overview

🧠 One-sentence thesis

Defining electric potential with respect to a datum at infinity simplifies calculating potential differences between any two points, just as node voltages simplify circuit analysis.

📌 Key points (3–5)

  • Why a datum matters: Instead of integrating the electric field along every path, we can define potential at each point relative to a common reference (infinity) and then simply subtract to find potential differences.
  • The datum choice: The datum is arbitrarily chosen as a sphere of infinite radius, so potential at any point r is the potential difference measured from infinity to that point.
  • Single point charge result: For a point charge q at the origin, the potential at distance r is V(r) = q divided by (4πε times r).
  • Generalization to multiple charges: For N point charges, the total potential is the sum of individual potentials, each depending on charge magnitude and distance from that charge to the field point.
  • Common confusion: Potential V(r) is not the same as potential difference V₂₁; V(r) is already a difference (from infinity to r), so V₂₁ = V(r₂) − V(r₁).

🔌 Why use a datum (the circuit analogy)

🔌 The circuit example

The excerpt uses a resistor circuit to motivate the datum concept:

  • In the circuit, there are two ways to find potential difference V₂₁ across a resistor:
    1. Calculate directly using current and resistance: V₂₁ = −IR
    2. Use node voltages: V₂₁ = V₂ − V₁
  • The second method is easier because you don't need to know what happens between the nodes (current, resistance, etc.)—only the node voltages relative to a common ground.

🎯 Advantage of node voltages

A common datum is a reference point with respect to which potential differences at all other locations can be defined.

  • Once you have node voltages (potentials at each point relative to the datum), calculating any potential difference becomes simple subtraction.
  • Example: If you know V₁ and V₂ relative to ground, you immediately get V₂₁ without re-integrating the electric field.

🌐 Datum in electrostatics

  • The datum for electrostatic problems is chosen as a sphere of infinite radius.
  • This choice is arbitrary but universal: it encompasses the entire universe, so every point's potential can be measured relative to "infinity."
  • Don't confuse: the datum is not a physical object; it's a mathematical reference at r → ∞.

🧮 Defining potential at a point

🧮 The definition

The electrical potential at a point r, given by V(r) = − integral from r to ∞ of E · dl, is the potential difference measured beginning at a sphere of infinite radius and ending at point r.

  • This is the "node voltage" for point r when the datum is at infinity.
  • The negative sign comes from the direction of integration (from infinity inward to r).

🛤️ Independence of path

  • The excerpt invokes the principle of independence of path (from Section 5.9): the integral does not depend on the specific path taken, only on the start and end points.
  • So we can choose the easiest path—for a point charge at the origin, a radial line (constant θ and φ) is simplest.
  • This simplifies the integral because dl = r̂ dr along a radial path.

⚡ Potential for a single point charge

⚡ Deriving V(r) for q at the origin

Starting from the electric field E = r̂ q / (4πε r²) and the definition V(r) = − integral from r to ∞ of E · dl:

  1. Choose dl = r̂ dr (radial path).
  2. The dot product E · dl = (q / 4πε r²) dr.
  3. Integrate: − (q / 4πε) times integral from r to ∞ of (1/r²) dr.
  4. Evaluate: − (q / 4πε) times [−1/r] from r to ∞ = + (q / 4πε) times (1/r).

Result:

V(r) = q / (4πε r)

  • The potential is positive if q is positive, negative if q is negative.
  • It depends only on the magnitude of charge and the distance r from the charge.
  • The excerpt suggests confirming dimensional correctness: charge divided by (permittivity times distance) has units of voltage.

🔄 Using V(r) to find potential differences

Once you have V(r) at every point, the potential difference from r₁ to r₂ is:

V₂₁ = V(r₂) − V(r₁)

  • This is "a lot easier" than re-integrating the electric field along a path from r₁ to r₂.
  • Example: If you know V(r₁) and V(r₂) from the formula above, you subtract them directly without worrying about the path.

🔢 Generalizing to multiple point charges

🔢 Charge not at the origin

If the point charge q′ is located at position r′ (not the origin), the potential at field point r is:

V(r; r′) = q′ / (4πε |r − r′|)

  • The formula depends only on the charge magnitude and the distance |r − r′| between the source point r′ and the field point r.
  • Notation: q′ and r′ denote the charge and position of the source; r is the observation point.

🔢 Superposition for N charges

For N point charges at positions r₁, r₂, ..., rₙ with charges q₁, q₂, ..., qₙ, the total potential is the sum:

V(r) = (1 / 4πε) times sum from n=1 to N of (qₙ / |r − rₙ|)

  • Each term is the potential due to one charge.
  • Superposition applies because potential is a scalar (no vector addition complications).
  • Example: Two charges at different locations contribute independently; add their potentials algebraically (watch signs).

🔢 What this formula gives

  • Equation 5.103 (the sum formula) gives the electric potential at any specified location r due to a finite number of charged particles.
  • The excerpt notes that continuous charge distributions (not just discrete particles) are addressed in the next section (5.13).

🔍 Key distinctions and reminders

🔍 Potential vs potential difference

ConceptWhat it isHow to get it
V(r)Potential at r (relative to infinity)Integrate E from ∞ to r, or use the formula for point charges
V₂₁Potential difference from r₁ to r₂V(r₂) − V(r₁), or integrate E from r₁ to r₂
  • Don't confuse: V(r) is already a potential difference (from infinity to r), not an absolute quantity.
  • The datum at infinity makes V(r) well-defined everywhere.

🔍 Why the datum is at infinity

  • The choice is arbitrary but convenient: it ensures that potential goes to zero as r → ∞ for finite charge distributions.
  • It allows a universal reference for all electrostatic problems.
  • Don't confuse: "infinity" is not a physical boundary; it's a limiting concept (r very large).

🔍 Path independence recap

  • The excerpt emphasizes that the integral defining V(r) does not depend on the path taken from infinity to r.
  • This is a consequence of the curl of E being zero (mentioned earlier in the excerpt as Kirchhoff's Voltage Law for electrostatics).
  • Practical implication: always choose the simplest path (e.g., radial for spherically symmetric fields).
61

Electric Potential Field due to a Continuous Distribution of Charge

5.13 Electric Potential Field due to a Continuous Distribution of Charge

🧭 Overview

🧠 One-sentence thesis

The electric potential field for continuous charge distributions is calculated by converting the discrete-particle summation formula into integrals over line, surface, or volume charge densities.

📌 Key points (3–5)

  • Starting point: The potential field for N discrete charged particles is a summation formula that can be extended to continuous distributions.
  • Three distribution types: Charge can be distributed continuously along a curve (line density), over a surface (surface density), or within a volume (volume density).
  • Mathematical technique: Each type uses the same approach—divide the distribution into small elements, express charge in terms of density times element size, then take the limit as element size approaches zero to obtain an integral.
  • Common confusion: The methods are mathematically identical to those used for electric field calculations in Section 5.4, but here they produce a scalar potential field rather than a vector field.

🔢 Foundation formula for discrete charges

🔢 Potential from multiple particles

The electrostatic potential field at position r due to N charged particles is: V(r) = (1 over 4πε) times the sum from n=1 to N of (q_n divided by the distance from r to r_n).

  • This formula gives potential at any specified location r.
  • Each particle n contributes based on its charge q_n and position r_n.
  • The distance term is the magnitude of the vector from the particle position to the field point.
  • This discrete formula is the foundation that will be converted to continuous forms.

📍 Why continuous distributions matter

  • Real-world charge configurations are more commonly continuous rather than countable discrete particles.
  • The excerpt notes that dealing with single charged particles is not typical in practice.
  • Three common continuous distribution types are addressed: along curves, over surfaces, and within volumes.

📏 Line charge distribution

📏 From segments to integral

The process for charge distributed along a curve C:

  1. Divide the curve into short segments of length Δl
  2. Express charge in each segment: q_n = ρ_l(r_n) times Δl, where ρ_l is line charge density in units of C/m
  3. Substitute into the discrete formula to get a summation
  4. Take the limit as Δl approaches zero

Result:

V(r) = (1 over 4πε) times the integral over curve C of (ρ_l(l) divided by the distance from r to r') dl

  • The variable r' represents the varying position along the curve during integration.
  • Integration is performed along the length l of the curve.

🎨 Surface charge distribution

🎨 From patches to integral

The process for charge distributed over a surface S:

  1. Divide the surface into small patches of area Δs
  2. Express charge in each patch: q_n = ρ_s(r_n) times Δs, where ρ_s is surface charge density in units of C/m²
  3. Substitute into the discrete formula to get a summation
  4. Take the limit as Δs approaches zero

Result:

V(r) = (1 over 4πε) times the integral over surface S of (ρ_s(r') divided by the distance from r to r') ds

  • The variable r' represents the varying position over the surface during integration.

📦 Volume charge distribution

📦 From cells to integral

The process for charge distributed within a volume V:

  1. Divide the volume into small cells (volume elements) of size Δv
  2. Express charge in each cell: q_n = ρ_v(r_n) times Δv, where ρ_v is volume charge density in units of C/m³
  3. Substitute into the discrete formula to get a summation
  4. Take the limit as Δv approaches zero

Result:

V(r) = (1 over 4πε) times the integral over volume V of (ρ_v(r') divided by the distance from r to r') dv

  • The variable r' represents the varying position throughout the volume during integration.

🔄 Comparison of the three methods

Distribution typeCharge density symbolDensity unitsElement sizeIntegration variable
Line (curve)ρ_lC/mΔldl along curve C
Surfaceρ_sC/m²Δsds over surface S
Volumeρ_vC/m³Δvdv within volume V

🔄 Unified pattern

All three methods follow the same mathematical structure:

  • Start with the discrete summation formula
  • Replace discrete charge q_n with (density times element size)
  • Convert summation to integral as element size approaches zero
  • The excerpt emphasizes this parallel structure is "essentially the same, from a mathematical viewpoint, as those developed in Section 5.4"

⚠️ Don't confuse

  • These integrals produce a scalar potential field V(r), not a vector field
  • The methods are analogous to electric field calculations but yield different physical quantities
  • The excerpt recommends reviewing Section 5.4 before attempting this section due to the mathematical similarity
62

Electric Field as the Gradient of Potential

5.14 Electric Field as the Gradient of Potential

🧭 Overview

🧠 One-sentence thesis

The electric field at any point can be calculated directly from the electric potential by taking the negative gradient, allowing field determination without explicitly considering the source charge distribution.

📌 Key points (3–5)

  • The inverse relationship: given potential V(r), the electric field E(r) is found by E = −∇V (the negative gradient of potential).
  • Physical interpretation: the electric field points in the direction where electric potential decreases most rapidly.
  • Mathematical origin: the relationship comes from the differential form of the potential-difference integral and the definition of the gradient operator.
  • Common confusion: the gradient points toward increasing potential, but the electric field points toward decreasing potential (hence the negative sign).
  • Practical advantage: you can find both magnitude and direction of E from V alone, without needing to know the charge distribution.

🔄 From potential difference to field

🔄 Starting point: the integral relationship

The excerpt begins with the known relationship from Section 5.8:

V₂₁ = −∫_C E(r) · dl

  • V₂₁ is the potential difference measured over path C.
  • E(r) is the electric field intensity at each point r along C.
  • This is an integral form, so the inverse (finding E from V) will be a differential equation.

🔬 The infinitesimal contribution

At a single point r, the infinitesimal contribution to the total integral is:

dV = −E(r) · dl

  • This isolates the contribution of an infinitesimal length element.
  • The goal is to manipulate this into a form that reveals E explicitly.

🧮 Deriving the gradient relationship

🧮 Using Cartesian coordinates

The derivation is simplest in Cartesian coordinates. The infinitesimal displacement is:

  • dl = x̂ dx + ŷ dy + ẑ dz

For any scalar function V(r), the total differential is (pure mathematics, not specific to electromagnetics):

  • dV = (∂V/∂x) dx + (∂V/∂y) dy + (∂V/∂z) dz

🔗 Connecting the two expressions

The excerpt notes that dx = dl · x̂ (and similarly for dy and dz). Substituting:

  • dV = (∂V/∂x)(dl · x̂) + (∂V/∂y)(dl · ŷ) + (∂V/∂z)(dl · ẑ)

Rearranging:

  • dV = ([x̂ ∂/∂x + ŷ ∂/∂y + ẑ ∂/∂z] V) · dl

⚡ Identifying the electric field

Comparing this rearranged form to dV = −E(r) · dl:

  • E(r) = −[x̂ ∂/∂x + ŷ ∂/∂y + ẑ ∂/∂z] V

The quantity in square brackets is the gradient operator ∇ (from Section 4.5), so:

E = −∇V

This is the central relationship: the electric field intensity at a point is the gradient of the electric potential at that point after a change of sign.

🧭 Physical interpretation

🧭 Direction of the field

The gradient of a scalar field points in the direction where that field increases most quickly. Therefore:

The electric field points in the direction in which the electric potential most rapidly decreases.

  • Don't confuse: ∇V points toward increasing V; E points toward decreasing V (the negative sign flips the direction).
  • This aligns with intuition: the electric field points away from positive charge (high potential) and toward negative charge (low potential).

🎯 Magnitude and direction from potential alone

The excerpt emphasizes a key advantage:

  • Both magnitude and direction of E can be determined from the potential field V alone.
  • You do not need to consider the charge distribution that is the physical source.
  • Example: if you know V(r) as a function of position, apply the gradient operator to get E immediately.

🔬 Example: charged particle at the origin

🔬 Setup and potential

The excerpt works through finding the electric field of a particle with charge q at the origin. The scalar potential (from Section 5.12) is:

  • V(r) = q / (4πεr)

This is expressed in spherical coordinates (r is the radial distance).

🔬 Applying the gradient in spherical coordinates

Using E = −∇V and the spherical-coordinate form of the gradient (from Section B.2):

  • E = −[r̂ ∂/∂r + θ̂ (1/r) ∂/∂θ + φ̂ (1/(r sin θ)) ∂/∂φ] (q / (4πεr))

Because V does not vary with φ or θ (it depends only on r), the second and third terms are zero:

  • E = −r̂ ∂/∂r (q / (4πεr))
  • E = −r̂ (q / 4πε) (−1/r²)
  • E = +r̂ q / (4πεr²)

✅ Consistency check

This result matches what was determined directly using Coulomb's Law in Section 5.1, confirming the gradient method works.

📊 Summary comparison

AspectWhat the excerpt shows
FormulaE = −∇V
SignNegative: field points toward decreasing potential
Coordinate systemDerivation simplest in Cartesian; example uses spherical
AdvantageFind E from V without needing charge distribution
Physical meaningField points where potential drops fastest
63

Poisson's and Laplace's Equations

5.15 Poisson’s and Laplace’s Equations

🧭 Overview

🧠 One-sentence thesis

Poisson's and Laplace's Equations provide an alternative method to calculate electric potential that naturally accommodates boundary conditions at material interfaces, making them especially useful when structures or spatially-varying materials are present.

📌 Key points (3–5)

  • Why an alternative method is needed: integration over charge distributions (Section 5.13) is awkward when material interfaces impose boundary conditions that must be satisfied simultaneously.
  • Poisson's Equation: relates the Laplacian of electric potential to volume charge density divided by permittivity; it is an inhomogeneous partial differential equation where the source term represents the field source.
  • Laplace's Equation: simplifies Poisson's Equation for source-free regions (no charge present); the Laplacian of potential equals zero.
  • Common confusion: Poisson's Equation applies everywhere, but Laplace's Equation only applies in regions free of charge—the charge may exist elsewhere but not in the domain of interest.
  • Practical advantage: combining these equations with boundary conditions reduces the problem to a standard boundary value problem, particularly effective when perfect conductors are present (constant potential surfaces).

🔧 Deriving Poisson's Equation

🔧 Starting from Gauss' Law

The derivation begins with the differential form of Gauss' Law:

Divergence of D equals volume charge density ρ_v

  • Using the relationship D = ε E (under standard material property assumptions), this becomes: divergence of E equals ρ_v divided by ε.
  • Next, apply the relationship E = negative gradient of V (from Section 5.14).
  • This yields: divergence of (gradient of V) equals negative ρ_v divided by ε.

🧮 The Laplacian operator

The operator "divergence of gradient" (∇ · ∇) is identically the Laplacian operator ∇².

In Cartesian coordinates, this expands to:

  • Second partial derivative with respect to x, plus second partial derivative with respect to y, plus second partial derivative with respect to z.
  • This identity holds regardless of the coordinate system employed.

📐 Final form of Poisson's Equation

Poisson's Equation: Laplacian of V equals negative ρ_v divided by ε

  • This states that the Laplacian of the electric potential field equals the volume charge density divided by permittivity, with a sign change.
  • It is a partial differential equation, solvable using well-known techniques.
  • It is an inhomogeneous differential equation: the inhomogeneous part (−ρ_v / ε) represents the source of the field.

🎯 Solving with boundary conditions

🎯 The boundary value problem approach

  • In the presence of material structure, identify relevant boundary conditions at interfaces between materials.
  • The task of finding V(r) reduces to the purely mathematical task of solving the associated boundary value problem.
  • This approach is particularly effective when one material is a perfect conductor or can be modeled as such.

⚡ Perfect conductor boundary condition

  • The electric potential at all points on the surface of a perfect conductor must be equal (constant).
  • This results in a particularly simple boundary condition.
  • Example: A perfectly conducting surface provides a known constant-potential boundary, simplifying the solution.

🌐 Laplace's Equation for source-free regions

🌐 When charge lies outside the domain

In many applications:

  • The charge responsible for the electric field lies outside the domain of the problem.
  • The region of interest has non-zero electric field (and potentially non-zero electric potential) but is free of charge.
  • In this case, Poisson's Equation simplifies to Laplace's Equation.

🌐 Laplace's Equation definition

Laplace's Equation: Laplacian of V equals zero (in source-free regions)

  • This states that the Laplacian of the electric potential field is zero in a source-free region.
  • Like Poisson's Equation, it is combined with relevant boundary conditions to solve for V(r).
  • Important limitation: it can only be used in regions that contain no charge.

🔄 Don't confuse: Poisson vs Laplace

EquationApplies whenRight-hand sideUse case
Poisson'sCharge present in region−ρ_v / ε (non-zero)General case with sources
Laplace'sSource-free region0Charge exists elsewhere, not in domain
  • The same problem may use Laplace's Equation in one region and Poisson's Equation in another.
  • Example: Between capacitor plates (no charge in spacer) → Laplace; in a region with distributed charge → Poisson.

🆚 Comparison with integration methods

🆚 Why not just integrate over charge distributions?

The excerpt contrasts two approaches to obtaining V(r):

Integration method (Section 5.13):

  • Integrate over the source charge distribution directly.
  • Awkward in the presence of material interfaces: boundary conditions (e.g., constant potential on conductors) are not automatically satisfied.
  • Constraints must be imposed separately and simultaneously.

Poisson/Laplace method:

  • Formulate as a partial differential equation with boundary conditions.
  • Boundary conditions are built into the problem from the start.
  • Facilitates analysis near structures and spatially-varying material properties.

🛠️ When to use which method

  • Use integration when the charge distribution is simple and no complex boundaries are present.
  • Use Poisson's/Laplace's Equations when material interfaces, conductors, or complex geometries impose boundary conditions that must be satisfied.
64

Potential Field Within a Parallel Plate Capacitor

5.16 Potential Field Within a Parallel Plate Capacitor

🧭 Overview

🧠 One-sentence thesis

Laplace's Equation can be used to rigorously derive that the electric potential inside a parallel-plate capacitor varies linearly with distance between the plates, confirming the informal result from earlier experiments.

📌 Key points (3–5)

  • What the section demonstrates: applying Laplace's Equation to find the potential field V(r) in the source-free region between capacitor plates.
  • Key simplification: because the plate radius a is much larger than the separation d, the field is approximately constant with radial position ρ until near the edges, so derivatives with respect to ρ and φ can be neglected.
  • Main result: the potential varies linearly with z (distance between plates): V(z) ≈ (V_C / d) × z + V_−, where V_C is the potential difference across the plates.
  • Common confusion: this linear solution applies only deep inside the capacitor (ρ ≪ a); near the edges, fringing fields appear and the radial derivative cannot be ignored.
  • Why it matters: the result matches the informal derivation from Section 2.2 and shows that the electric field intensity E ≈ −ẑ V_C / d, a fundamental result for capacitors.

🧩 Problem setup and geometry

🧩 The parallel-plate capacitor structure

  • Two perfectly conducting circular disks separated by distance d.
  • A spacer material with permittivity ε fills the gap; no charge is present in the spacer, so the region is source-free.
  • Plate radius a is much larger than d (a ≫ d), allowing edge effects to be neglected in the main analysis.

🔋 Boundary conditions

  • The potential difference between the plates is V_C (also the terminal voltage).
  • Let V_− be the node voltage at the negative plate (z = 0).
  • Then the positive plate (z = +d) has voltage V_− + V_C.
  • These two conditions will determine the constants in the solution.

🧮 Applying Laplace's Equation

🧮 Starting equation

Laplace's Equation (source-free region): ∇²V = 0

  • In cylindrical coordinates (ρ, φ, z), this expands to:
    • (1/ρ) ∂/∂ρ (ρ ∂V/∂ρ) + (1/ρ²) ∂²V/∂φ² + ∂²V/∂z² = 0

🔍 Exploiting symmetry

  • Radial symmetry: the problem has cylindrical symmetry around the z-axis, so ∂V/∂φ = 0.
  • Neglecting radial variation: because d ≪ a, the fields are approximately constant with ρ until near the plate edges, so ∂V/∂ρ ≈ 0 for ρ ≪ a.
  • After these simplifications, Laplace's Equation reduces to:
    • ∂²V/∂z² ≈ 0 for ρ ≪ a

⚙️ Solving the simplified equation

  • Integrating ∂²V/∂z² = 0 twice yields the general solution:
    • V(z) = c₁ z + c₂
  • The constants c₁ and c₂ must be determined from boundary conditions.

🎯 Determining the solution

🎯 Applying boundary conditions

  • At z = 0 (negative plate): V(0) = V_−
    • Substituting into V(z) = c₁ z + c₂ gives c₂ = V_−.
  • At z = +d (positive plate): V(d) = V_− + V_C
    • Substituting gives c₁ d + V_− = V_− + V_C, so c₁ = V_C / d.

📐 Final result for potential

  • The electric potential between the plates is:
    • V(z) ≈ (V_C / d) z + V_− for ρ ≪ a
  • This result is dimensionally correct (voltage units) and shows that potential changes linearly with distance z.

⚡ Electric field intensity

  • Applying E = −∇V (from Section 5.14) to the solution above:
    • E ≈ −ẑ V_C / d
  • This is the expected uniform field pointing from positive to negative plate, with magnitude V_C / d.
  • Important: this matches the result derived informally in Section 2.2, now confirmed rigorously.

🌊 Fringing fields and limitations

🌊 What are fringing fields?

Fringing field: the field near the edge of the plates or beyond the plates, where the radial derivative ∂V/∂ρ is no longer negligible.

  • The linear solution V(z) ≈ (V_C / d) z + V_− applies only deep inside the capacitor (ρ ≪ a).
  • Near the edges, the assumption ∂V/∂ρ ≈ 0 breaks down.

🚧 Why fringing fields are harder to calculate

  • Must account for ∂V/∂ρ in Laplace's Equation.
  • Boundary conditions become more complex:
    • Must include the outside surfaces of the plates (sides facing away from the dielectric).
    • Must account for the boundary between the spacer material and free space.
  • When accurate fringing-field calculation is needed, numerical solutions of Laplace's Equation are common.

🛠️ Practical implications

  • Fortunately, accurate calculation of fringing fields is usually not required in practical engineering applications.
  • The linear approximation is sufficient for most purposes when a ≫ d.

🔗 Connection to other concepts

🔗 Relation to earlier work

  • This section provides a rigorous derivation of what was figured out informally in Section 2.2 (the battery-charged capacitor experiment).
  • The result confirms the intuitive understanding with a formal mathematical framework.

🔗 Preview of capacitance

  • The section mentions that familiarity with capacitance or capacitors is not required here.
  • For those interested, Section 5.22 provides a preview of capacitance concepts.
  • The potential field derived here is foundational for understanding capacitor behavior.
65

Boundary Conditions on the Electric Field Intensity (E)

5.17 Boundary Conditions on the Electric Field Intensity ( E )

🧭 Overview

🧠 One-sentence thesis

At interfaces between dissimilar media, the tangential component of the electric field E must be continuous, and on perfectly conducting surfaces it must be zero, providing constraints that govern how electric fields behave across boundaries.

📌 Key points (3–5)

  • Discontinuity at interfaces: In homogeneous media electromagnetic quantities vary smoothly, but at interfaces between dissimilar media they can be discontinuous—boundary conditions describe these discontinuities mathematically.
  • PEC surface rule: On a perfectly electrical conductor (PEC) surface, the tangential component of E is zero, so E must be perpendicular to the surface.
  • General boundary condition: For any two media (not necessarily PEC), the tangential component of E is continuous across the interface—what is tangent on one side equals what is tangent on the other.
  • Common confusion: The PEC case is a special instance of the general rule—when one medium is PEC, the tangential E on the other side must also be zero because it is zero inside the PEC.
  • Broader applicability: This boundary condition holds for both electrostatic and time-varying cases, not just static fields.

⚡ Discontinuities and boundary conditions

⚡ Why boundary conditions matter

  • In homogeneous (uniform) media, electromagnetic quantities vary smoothly and continuously.
  • At an interface between dissimilar media, electromagnetic quantities can be discontinuous—they can jump in value.
  • Boundary conditions are mathematical descriptions of these discontinuities.
  • They constrain solutions for electromagnetic quantities, making it possible to solve problems involving multiple materials.

🧱 The interface setup

  • Consider two media meeting at an interface defined by a mathematical surface S.
  • The boundary condition describes how the electric field E behaves as you cross S from one medium to the other.
  • The unit vector (normal) is perpendicular to the surface at each point.

🔌 Perfect electrical conductor (PEC) surfaces

🔌 Equipotential property

If either material is a perfect electrical conductor (PEC), then S is an equipotential surface: the electric potential V is constant everywhere on S.

  • Because E is proportional to the spatial rate of change of potential (recall E = −∇V), a constant potential means no tangential electric field.
  • The component of E tangent to a PEC surface is zero.

🔌 Mathematical expression for PEC

  • Informally: E_tan = 0 on PEC surface.
  • More precisely: E × = 0 on PEC surface.
    • The cross product of any two vectors is perpendicular to both.
    • Any vector perpendicular to is tangent to S.
    • So E × = 0 means there is no tangential component.
  • Result: The electric field at a PEC surface must be oriented entirely perpendicular to the surface.
  • Example: If a metal plate is one of the media, the electric field lines must hit the metal surface at right angles; they cannot run along the surface.

🔄 General boundary condition (non-PEC case)

🔄 Deriving the condition using KVL

  • The general boundary condition applies even when neither medium is a perfect conductor.
  • It is derived from Kirchhoff's Voltage Law (KVL): the line integral of E around any closed path is zero (in the electrostatic case).
  • Set up a rectangular closed path centered on the interface S, with sides perpendicular and parallel to the surface.

🔄 Shrinking the rectangle

  • Let the perpendicular sides have length w and the parallel sides have length l.
  • Shrink w and l together while keeping the ratio w/l much less than 1 (a very flat rectangle).
  • As the rectangle shrinks:
    • The contributions from the two perpendicular sides (B and D) become equal in magnitude but opposite in sign, so they cancel.
    • Only the contributions from the two parallel sides (A and C) remain.
  • Define a unit vector (tangent) along the parallel sides.
  • When the sides become very small, the integral becomes: E₁ · ΔlE₂ · Δl → 0.
  • This can only be true if the tangential components of E₁ and E₂ are equal.

🔄 The continuity rule

The tangential component of E must be continuous across an interface between dissimilar media.

  • Mathematically: E₁ × = E₂ × on S.
  • More commonly written: × (E₁ − E₂) = 0 on S.
  • This means the tangent component on one side equals the tangent component on the other side.
  • Example: If you measure the component of E parallel to the interface in medium 1, it will equal the parallel component in medium 2.

🔄 PEC as a special case

  • The PEC boundary condition is a special instance of the general rule.
  • Inside a PEC, E = 0 (perfect conductor has no internal field).
  • So the tangential component inside the PEC is zero.
  • By continuity, the tangential component just outside the PEC must also be zero.
  • Don't confuse: The general rule says "tangent components are equal"; the PEC rule says "tangent component is zero"—the PEC rule follows from the general rule when one side has zero field.

🌐 Time-varying fields and broader applicability

🌐 Beyond electrostatics

  • The boundary condition × (E₁ − E₂) = 0 applies to both electrostatic and time-varying (general) cases.
  • In the presence of time-varying magnetic fields, the right-hand side of KVL may become non-zero and is proportional to the area of the closed loop.
  • However, the derivation requires the area of the loop to approach zero.
  • As the area shrinks to zero, any possible difference from the static KVL also converges to zero.
  • Therefore, the tangential continuity condition holds generally, not just for static fields.

🌐 Why this matters

  • You can use the same boundary condition to solve problems involving:
    • Static electric fields (e.g., capacitors with different dielectrics).
    • Time-varying fields (e.g., electromagnetic waves crossing material boundaries).
  • The continuity of the tangential component is a fundamental constraint that applies across all electromagnetic scenarios.
66

Boundary Conditions on the Electric Flux Density (D)

5.18 Boundary Conditions on the Electric Flux Density ( D )

🧭 Overview

🧠 One-sentence thesis

The normal component of electric flux density D is discontinuous across a boundary by an amount equal to the surface charge density, while in the absence of surface charge the normal component must be continuous.

📌 Key points (3–5)

  • What the boundary condition says: any discontinuity in the normal component of D across an interface equals the surface charge density ρₛ at that point.
  • Special case—PEC surface: at a perfect conductor surface, the magnitude of D equals the surface charge density and points perpendicular into the non-conducting medium.
  • No surface charge case: when there is no surface charge at the interface, the normal component of D must be continuous across the boundary.
  • Common confusion: the boundary condition for D (normal component) differs from that for E (tangential component)—E tangential is always continuous, but D normal is continuous only when surface charge is absent.
  • Why flux density matters: using D instead of E simplifies boundary conditions because D directly relates to surface charge, whereas E requires permittivity factors.

🔬 Derivation using Gauss' Law

🔬 The cylindrical surface setup

The excerpt derives the boundary condition by applying Gauss' Law to a closed cylindrical surface centered at the interface:

  • The cylinder has flat ends parallel to the interface, radius a, and total length 2h.
  • One flat end is in Region 1, the other in Region 2.
  • The normal vector points into Region 1.

📐 Shrinking the cylinder

The derivation requires letting both h and a approach zero while maintaining h/a ≪ 1:

  • Because h is much smaller than a, the side area becomes negligible compared to the top and bottom areas.
  • As the cylinder shrinks, the variation in D over the top and bottom becomes negligible.
  • The integral form of Gauss' Law simplifies to contributions from only the top and bottom surfaces.

🧮 The resulting equation

Starting from Gauss' Law:

The integral of D over the closed surface equals the enclosed charge Q_encl.

After shrinking the cylinder, the excerpt obtains:

  • D₁ · n̂ ΔA + D₂ · (−n̂) ΔA → Q_encl
  • Rearranging: n̂ · (D₁ − D₂) → Q_encl / ΔA
  • The right side is charge per unit area, which is surface charge density ρₛ.

The final boundary condition:

n̂ · (D₁ − D₂) = ρₛ

where points into Region 1.

⚡ Special case: perfect conductor boundary

⚡ PEC surface behavior

When Region 2 is a perfect electrical conductor (PEC):

  • The electric field inside the PEC is identically zero (because potential is constant throughout).
  • Therefore D₂ = 0.
  • The boundary condition simplifies to D · n̂ = ρₛ on the PEC surface.

At the surface of a perfect conductor, the magnitude of D is equal to the surface charge density ρₛ (units of C/m²) at that point.

🔍 Direction of D at PEC surface

  • From Section 5.17 (referenced in the excerpt), the tangential component of E must be zero at a PEC surface.
  • Therefore E (and hence D) is directed entirely perpendicular to the surface.
  • D points into the non-conducting medium.
  • Example: if a conductor has positive surface charge, D points outward from the conductor surface with magnitude equal to the charge density.

🌐 General case: two dielectric media

🌐 No surface charge present

When neither region is a PEC and there is no surface charge at the interface:

  • The boundary condition becomes n̂ · (D₁ − D₂) = 0.
  • This means n̂ · D₁ = n̂ · D₂.

In the absence of surface charge, the normal component of the electric flux density must be continuous across the boundary.

🔄 Relationship to electric field E

Since D = ε E (where ε is permittivity), the boundary condition can be rewritten in terms of E:

n̂ · (ε₁ E₁ − ε₂ E₂) = ρₛ

where ε₁ and ε₂ are the permittivities in Regions 1 and 2.

📊 Why prefer D over E

The excerpt notes:

  • The boundary condition for D directly involves surface charge without needing permittivity factors.
  • This illustrates one reason to prefer the "flux" interpretation (D) over the "field intensity" interpretation (E).
  • Example: when comparing fields across a boundary, D normal changes only when surface charge is present, making the physical picture clearer.

🔀 Comparison with E boundary condition

🔀 Tangential vs normal components

FieldComponentBoundary conditionWhen it applies
ETangentialAlways continuous across interfaceBoth electrostatic and time-varying cases
DNormalDiscontinuous by amount ρₛElectrostatic case (derived here)
DNormalContinuous when ρₛ = 0No surface charge present

⚠️ Don't confuse

  • E tangential continuity (from Section 5.17) is a separate condition from D normal discontinuity.
  • The tangential component of E is continuous even when surface charge is present.
  • The normal component of D is discontinuous only when surface charge is present.
  • These two conditions together fully specify the boundary behavior of electromagnetic fields at an interface.
67

Charge and Electric Field for a Perfectly Conducting Region

5.19 Charge and Electric Field for a Perfectly Conducting Region

🧭 Overview

🧠 One-sentence thesis

In a perfect electrical conductor (PEC), all charge resides on the surface, the internal electric field is zero, and the external field is perpendicular to the surface with magnitude equal to the surface charge density.

📌 Key points (3–5)

  • Electric field inside a PEC is zero: both electric field intensity E and electric flux density D are zero throughout the PEC because potential is constant.
  • All charge is on the surface: there can be no static charge within a PEC; charge lies entirely on the surface.
  • Field orientation at the boundary: the electric field outside a PEC is perpendicular to the surface (tangential component is zero).
  • Discontinuity at the boundary: the normal component of D equals the surface charge density ρ_s; D₁ and D₂ have the same magnitude |ρ_s| on opposite sides of a PEC slab.
  • Common confusion: electric flux density D and electric field intensity E have the same magnitude at the surface only if the permittivities on both sides are equal; otherwise E₁ ≠ E₂ even though |D₁| = |D₂|.

⚡ Why the electric field is zero inside a PEC

⚡ Constant potential throughout

The electrical potential throughout a PEC region must be constant.

  • The electric field is proportional to the spatial rate of change of electrical potential: E = −∇V.
  • If potential V is constant everywhere inside the PEC, then the gradient (rate of change) is zero.
  • Therefore, both E and D must be zero throughout the PEC region.

🚫 No internal charge

  • Any non-zero charge distribution inside the PEC would create a non-zero electric field.
  • A non-zero field would produce a potential difference between locations inside the slab.
  • This contradicts the requirement that potential is constant.
  • Conclusion: there can be no static charge within a PEC.

🔲 Charge distribution and boundary behavior

🔲 Surface charge only

Charge associated with a PEC lies entirely on the surface.

  • Since no charge can exist inside the PEC, all charge must reside on the surface.
  • The surface charge density is denoted ρ_s (units: C/m²).

📐 Perpendicular field orientation

  • The electric field outside the PEC is oriented directly away from (perpendicular to) the PEC surface.
  • The component of the electric field tangent to (parallel to) the PEC surface is zero.
  • This is a boundary condition: both E₁ (above) and E₂ (below) must have zero tangential components.
  • Example: in a flat slab scenario, if E₁ is above the slab and E₂ is below, both fields point perpendicular to their respective surfaces.

🔗 Magnitude of D at the surface

  • The magnitude of the electric flux density D at the surface equals the surface charge density ρ_s.
  • This relationship comes from the boundary condition: ˆn · (D₁ − D₂) = ρ_s, where ˆn points into region 1.
  • When region 2 is a PEC, D₂ = 0, so the boundary condition simplifies.

🧲 Discontinuity and field differences across the boundary

🧲 Discontinuity in the normal component

Any discontinuity in the normal component of the electric flux density across the boundary between two material regions is equal to the surface charge.

  • The boundary condition is: ˆn · (D₁ − D₂) = ρ_s.
  • If there is no surface charge (ρ_s = 0), then the normal component of D must be continuous: ˆn · (D₁ − D₂) = 0.
  • The surface charge supports the discontinuity in the normal component of the electric fields.

⚖️ D vs E: same magnitude, different intensities

  • For a PEC slab with dielectric regions on both sides, D₁ and D₂ have the same magnitude |ρ_s| because the surface charge densities on both sides have equal magnitude.
  • However, the electric field intensity E is related to D by the permittivity: E₁ = D₁/ε₁ and E₂ = D₂/ε₂.
  • Don't confuse: even though |D₁| = |D₂|, the electric field intensities E₁ and E₂ are unequal unless the permittivities ε₁ and ε₂ are equal.
QuantityRegion 1Region 2Relationship
Electric flux densityD₁D₂|D₁| = |D₂| = |ρ_s|
Electric field intensityE₁ = D₁/ε₁E₂ = D₂/ε₂E₁ ≠ E₂ unless ε₁ = ε₂

🌀 Field lines in general cases

  • In more general scenarios (e.g., a point charge near PEC regions of various shapes), electric field lines bend in the dielectric material.
  • The bending ensures that the tangential component of the electric field is zero on PEC surfaces.
  • The charge distribution arranges itself on the PEC surfaces to maintain zero electric field and constant potential within the PEC.
  • Example: Figure 5.12 shows field lines bending around PEC regions (shaded) to satisfy boundary conditions.

📏 Boundary condition on E

📏 Relating E to surface charge

  • Since D = εE, the boundary condition on D implies a boundary condition on E:
    • ˆn · (ε₁E₁ − ε₂E₂) = ρ_s
    • where ε₁ and ε₂ are the permittivities in regions 1 and 2, respectively.
  • This equation illustrates why the "flux" interpretation (D) is sometimes preferred over the "field intensity" interpretation (E): D directly relates to surface charge without needing to account for permittivity differences.
68

Dielectric Media

5.20 Dielectric Media

🧭 Overview

🧠 One-sentence thesis

Dielectrics are low-conductivity materials whose molecules remain intact under electric fields, making them effective insulators with properties quantified primarily by relative permittivity.

📌 Key points

  • What dielectrics are: materials with low conductivity because their molecules stay intact when exposed to electric fields, rather than shedding electrons like conductors.
  • Why they matter: dielectrics serve as spacers in printed circuit boards, coaxial cables, and capacitors—they are "good insulators" and "poor conductors."
  • Key property: relative permittivity (ε_r) ranges from close to 1 up to roughly 50; most common low-moisture materials have ε_r less than 6.
  • Common confusion: dielectrics are "non-magnetic"—their permeability is approximately equal to free-space permeability (μ ≈ μ₀), not their permittivity.
  • Practical limit: all dielectrics fail at sufficiently strong electric fields through dielectric breakdown, a sudden increase in conductivity.

🧱 What dielectrics are and how they differ from conductors

🔬 Molecular behavior under electric fields

Dielectric: a particular category of materials that exhibit low conductivity because their constituent molecules remain intact when exposed to an electric field, as opposed to shedding electrons as is the case in good conductors.

  • The defining feature is molecular integrity: molecules do not break apart or release electrons.
  • Conductors, by contrast, shed electrons when exposed to electric fields.
  • Result: dielectrics do not effectively pass current.

🛡️ Insulating vs conducting

  • Dielectrics are considered both:
    • "Good insulators": they block current flow.
    • "Poor conductors": they do not facilitate electron movement.
  • This dual characterization reflects the same underlying property from two perspectives.

🔧 Applications and common materials

🔧 Where dielectrics are used

The excerpt highlights three major applications:

  • Printed circuit boards (PCBs): dielectrics act as spacer material.
  • Coaxial cables: dielectrics separate conductors.
  • Capacitors: dielectrics occupy the space between charged plates.

🧪 Examples of dielectric materials

Common dielectrics include:

  • Air
  • Glass
  • Teflon
  • Fiberglass epoxy (used in "FR4" printed circuit boards)

The excerpt notes that values of constitutive parameters for these and other dielectrics are listed in Section A.1.

📐 Electromagnetic properties

📏 Relative permittivity (ε_r)

The electromagnetic properties of dielectric materials are quantified primarily by relative permittivity ε_r.

  • Range: from very close to 1 upward to roughly 50.
  • Typical values: less than 6 for most commonly-encountered materials with low moisture content.
  • Relative permittivity is introduced in Section 2.3 of the source material.

🧲 Permeability and non-magnetic behavior

  • The permeability of dielectric materials is approximately equal to the free-space value: μ ≈ μ₀.
  • Because of this, dielectrics are said to be "non-magnetic."
  • Don't confuse: "non-magnetic" refers to permeability being close to free-space permeability, not to permittivity or conductivity.

⚡ Dielectric breakdown

⚡ Linearity and its limits

  • An ideal dielectric has permittivity independent of the applied electric field magnitude; such a material is called "linear."
  • All practical dielectrics fail this linearity with sufficiently strong electric fields.
  • The failure is typically abrupt.

💥 What happens at breakdown

Dielectric breakdown: the sudden change in behavior observed in the presence of an electric field greater than the dielectric strength threshold value.

  • Observed as: a sudden, dramatic increase in conductivity.
  • Mechanism: electrons are successfully dislodged from their host molecules.
  • Threshold: the electric field intensity at which this occurs is known as dielectric strength.

📊 Dielectric strength values

MaterialDielectric strength
AirAbout 3 MV/m
Mica (used in capacitors)About 200 MV/m

⚡ Arcing and consequences

  • Breakdown is typically accompanied by "arcing": a sudden flow of current associated with the breakdown.
  • Example: Lightning occurs when charge is exchanged between sky and ground as air (a dielectric) exhibits breakdown.
  • Damage: Dielectric breakdown in solids typically damages the material permanently.
69

Dielectric Breakdown

5.21 Dielectric Breakdown

🧭 Overview

🧠 One-sentence thesis

Dielectric breakdown occurs when a sufficiently strong electric field causes a dielectric material to suddenly lose its insulating properties and become conductive, with the threshold field strength varying widely by material.

📌 Key points (3–5)

  • What breakdown is: a sudden, dramatic increase in conductivity when electric field exceeds a threshold, caused by electrons being dislodged from molecules.
  • Dielectric strength: the threshold electric field intensity at which breakdown occurs; ranges from about 3 MV/m (air) to about 200 MV/m (mica).
  • Common confusion: ideal vs. practical dielectrics—ideal dielectrics have permittivity independent of field strength ("linear"), but all real dielectrics fail at high enough fields.
  • Observable phenomenon: breakdown is typically accompanied by arcing (sudden current flow); lightning is a familiar example.
  • Material damage: breakdown in solids typically damages the material permanently.

⚡ The breakdown phenomenon

⚡ What happens during breakdown

Dielectric breakdown: the sudden change in behavior observed when an electric field exceeds the dielectric strength threshold.

  • In normal operation, a dielectric is "linear"—its permittivity does not depend on the applied electric field magnitude.
  • When the field becomes sufficiently strong, the material fails abruptly.
  • The failure manifests as a sudden, dramatic increase in conductivity.
  • The physical mechanism: electrons are successfully dislodged from their host molecules, allowing current to flow.

🔥 Arcing and observable effects

  • Breakdown is typically accompanied by arcing: a sudden flow of current associated with the breakdown event.
  • Example: Lightning occurs when charge is exchanged between sky and ground after air (a dielectric) exhibits breakdown.
  • In solid dielectrics, breakdown typically damages the material permanently.

📏 Dielectric strength

📏 Definition and range

Dielectric strength: the threshold value of the electric field intensity at which dielectric breakdown occurs.

  • This is a material property that quantifies how much electric field a dielectric can withstand before failing.
  • The excerpt provides a range across common materials:
MaterialDielectric Strength
Air~3 MV/m
Mica~200 MV/m
  • Mica is noted as "a dielectric commonly used in capacitors," highlighting its high breakdown threshold.

🔍 Practical vs. ideal behavior

  • Ideal dielectric: permittivity independent of applied field magnitude; remains linear at all field strengths.
  • Practical dielectrics: all real materials fail this ideal behavior when the electric field becomes sufficiently strong.
  • Don't confuse: the material is still a dielectric below the threshold; breakdown is the transition point where insulating behavior suddenly stops.

🧱 Material context

🧱 General dielectric properties

The excerpt provides context about dielectrics before discussing breakdown:

  • Relative permittivity (ε_r): ranges from very close to 1 up to roughly 50; most common low-moisture materials have ε_r less than about 6.
  • Permeability: approximately equal to free-space value (μ ≈ μ₀), so dielectrics are "non-magnetic."
  • These properties are listed in Section A.1 of the source material.

🌩️ Why breakdown matters

  • Understanding dielectric strength is critical for designing electrical systems that operate safely below the breakdown threshold.
  • Natural phenomena like lightning demonstrate breakdown in air dielectrics.
  • In engineered systems (e.g., capacitors using mica), choosing materials with appropriate dielectric strength prevents catastrophic failure.
70

Capacitance

5.22 Capacitance

🧭 Overview

🧠 One-sentence thesis

Capacitance quantifies a structure's ability to store energy in an electric field, determined by its geometry and the permittivity of the material between charge regions, and is defined as the ratio of charge to potential difference.

📌 Key points (3–5)

  • What capacitance measures: the ability of a structure to store energy in an electric field, not dependent on the amount of charge itself but on the structure's response to charge.
  • What determines capacitance: geometry of the charge distribution and permittivity of the intervening medium; charge and potential are stimulus/response, not determinants.
  • Practical definition: capacitance is the ratio of total positive charge on one conductor to the potential difference between two conductors (C = Q⁺ / V).
  • Common confusion: capacitance does not depend on charge—charge is either the stimulus or the response; capacitance is a property of the structure itself.
  • Connection to circuits: in circuit theory, capacitors relate terminal voltage and current through the integral/differential relationship between charge accumulation and voltage.

⚡ Energy storage and the physical basis

⚡ Potential energy in separated charges

  • When positive and negative charge regions exist near each other, Coulomb forces try to pull them together.
  • This tendency represents stored potential energy in the system.
  • The potential energy is proportional to:
    • Quantity of positive charge squared
    • Inversely proportional to separation squared
    • Inversely proportional to permittivity of the separating material

🔋 Energy stored in the electric field

  • Electric field intensity E is defined through Coulomb force: F = qE.
  • Instead of viewing energy as "in the Coulomb force," we can equivalently view it as stored in the electric field created by the charge distribution.
  • This stored energy depends on:
    • Geometry of the charge distribution
    • Permittivity of the intervening media
  • Don't confuse: the energy is not a property of the charge alone—it depends on how the charge is arranged and what material separates the regions.

📐 Defining capacitance

📐 What capacitance is

Capacitance is the ability of a structure to store energy in an electric field.

The capacitance of a structure depends on its geometry and the permittivity of the medium separating regions of positive and negative charge.

  • Capacitance does not depend on charge.
  • Charge is viewed as either a stimulus (you apply charge and measure voltage) or a response (you apply voltage and measure charge).

🔢 The capacitance formula

In practice, capacitance is defined as the ratio of charge present on one conductor of a two-conductor system to the potential difference between the conductors.

C = Q⁺ / V

Where:

  • Q⁺: total positive charge (units: coulombs, C)
  • V: potential difference associated with this charge, defined to be positive (units: volts, V)
  • C: capacitance (units: farads, F)

Interpretation: A structure has greater capacitance if it stores more charge—and therefore more energy—in response to a given potential difference.

🔌 Two-conductor system setup

  • A battery imposes potential difference V between two regions of perfectly conducting material (PEC).
  • Q⁺ is the total charge on the surface of the PEC region attached to the positive terminal.
  • An equal amount of negative charge (−Q⁺) appears on the surface of the PEC region attached to the negative terminal.
  • This charge distribution creates an electric field.
  • If the two PEC regions are fixed in place, Q⁺ increases linearly with increasing V, at a rate determined by capacitance C.

🧰 Capacitors in circuit theory

🧰 What a capacitor is

A capacitor is a device that is designed to exhibit a specified capacitance.

  • In circuit theory, devices are characterized by terminal voltage V_T and terminal current I_T.
  • Current does not normally flow through a capacitor (if it does, dielectric breakdown is likely occurring).

⚙️ Terminal current interpretation

  • "Terminal current" for a capacitor means the flow of charge arriving at or departing from one conductor via the circuit.
  • An equal flow of charge departs or arrives at the other conductor.
  • This gives the appearance of current flow through the capacitor when viewed from outside.

🔗 Voltage-charge relationship

Starting from the definition C = Q⁺ / V, we express terminal voltage:

V_T = Q⁺ / C

Current is charge per unit time, so the charge on either conductor is the integral of I_T over time:

Q⁺(t) = integral from t₀ to t of I_T(τ) dτ + Q⁺(t₀)

Where t₀ is an arbitrarily selected start time. In words: amps integrated over time equals charge.

🔗 Voltage-current relationships

Substituting the integral expression for charge into the voltage equation:

V_T(t) = (1/C) × [integral from t₀ to t of I_T(τ) dτ] + (1/C) × Q⁺(t₀)

The second term is simply V_T(t₀), the initial voltage. This is the expected relationship from elementary circuit theory.

Differential form (solving for current):

I_T(t) = C × d/dt V_T(t)

In words: the terminal current equals capacitance times the rate of change of terminal voltage.

🧪 The thin parallel plate capacitor

🧪 Structure description

The excerpt introduces the thin parallel plate capacitor:

  • Consists of two flat plates, each with area A
  • Separated by distance d
  • Coordinate system: origin at center of lower plate, +z axis directed toward upper plate, so upper plate lies in the z = +d plane

🔍 Approach to finding capacitance

The method outlined:

  1. Assume a particular charge on one plate
  2. Use the boundary condition on electric flux density D to relate charge density to the internal electric field
  3. Integrate the electric field between the plates to obtain potential difference
  4. Capacitance is the ratio of assumed charge to resulting potential difference

🔍 The challenge: finding the electric field

  • If the plate area were infinite, the electric field would be simple: perpendicular to the plates, constant everywhere between them.
  • For finite plates, the field is more complex (the excerpt does not complete this analysis).

Don't confuse: infinite vs. finite plate assumptions—infinite plates yield uniform fields; real (finite) plates have edge effects that complicate the field distribution.

71

The Thin Parallel Plate Capacitor

5.23 The Thin Parallel Plate Capacitor

🧭 Overview

🧠 One-sentence thesis

The capacitance of a thin parallel plate capacitor is approximately ε·A/d, derived by assuming uniform charge distribution and neglecting the fringing field at the edges when the plate dimensions are much larger than the separation distance.

📌 Key points (3–5)

  • The "thin" condition: when the smallest plate dimension is much greater than the separation distance d, the fringing field becomes negligible and can be ignored.
  • Capacitance formula: C ≈ ε·A/d, where ε is permittivity, A is plate area, and d is separation distance.
  • Key simplifications: the thin condition allows us to assume uniform charge distribution, makes plate shape irrelevant (only area matters), and makes plate thickness irrelevant.
  • Common confusion: fringing fields always exist at finite plate edges, but in thin capacitors most energy is stored in the uniform central field, so fringing effects can be neglected as an approximation.
  • Parameter effects: capacitance increases with permittivity and plate area, but decreases with separation distance.

🧩 The thin capacitor approximation

🧩 What makes a capacitor "thin"

A parallel plate capacitor is "thin" when the smallest identifiable dimension of the plate is much greater than d (the separation distance).

  • This is a geometric condition, not about physical thickness of the conductor material.
  • When this condition is met, the fringing field stores an insignificant fraction of the total energy compared to the uniform central field.
  • The approximation works by simply neglecting the fringing field entirely.

🌊 Understanding fringing fields

  • Infinite plates: if plate area were infinite, the electric field would be perfectly uniform—perpendicular to the plates and constant everywhere between them (from symmetry alone).
  • Finite plates: when plate area is finite, a non-uniform "fringing field" emerges near the edges.
  • Why fringing occurs: boundary conditions on the outside (outward-facing) surfaces of the plates affect the field significantly near the edges.
  • Central vs edge regions: in the central region, the field resembles the infinite-plate case; near edges, it becomes non-uniform.
  • Don't confuse: fringing fields are real and always present, but the thin condition makes them negligible for calculating total capacitance.

🎯 Three key simplifications from the thin condition

When the thin condition is imposed:

  1. Uniform charge distribution: surface charge density can be assumed approximately uniform over each plate, greatly simplifying analysis.
  2. Shape irrelevance: plates can be circular, square, triangular, etc.—only the total plate area A matters for capacitance.
  3. Thickness irrelevance: the physical thickness of each conductor plate becomes unimportant.

🔬 Derivation of capacitance

🔬 Step-by-step derivation strategy

The excerpt uses this approach:

  • Assume a particular charge Q⁺ on one plate.
  • Use boundary conditions on electric flux density D to relate charge density to the internal electric field.
  • Integrate over the electric field between plates to obtain potential difference V.
  • Compute capacitance as the ratio C = Q⁺/V.

⚡ Key steps in the calculation

The excerpt walks through eight steps:

  1. Assume charge: total positive charge Q⁺ on the upper plate.
  2. Uniform density: invoking the thin condition, surface charge density on the bottom side of the upper plate is ρₛ,₊ = Q⁺/A (in coulombs per square meter).
  3. Boundary condition (upper plate): from the boundary condition, D on the bottom surface of the upper plate is −ẑ·ρₛ,₊.
  4. Lower plate charge: total charge on lower plate Q₋ must equal −Q⁺; surface charge density ρₛ,₋ must be −ρₛ,₊.
  5. Boundary condition (lower plate): D on the top surface of the lower plate is +ẑ·ρₛ,₋, which equals −ẑ·ρₛ,₊, so D on the facing sides is equal.
  6. Field between plates: invoking the thin condition again, D between the plates has approximately the same structure as infinite plates, so D ≈ −ẑ·ρₛ,₊ everywhere between the plates.
  7. Potential difference: integrate the electric field from z=0 to z=d to find V = ρₛ,₊·d/ε.
  8. Final capacitance: C = Q⁺/V = (ρₛ,₊·A)/(ρₛ,₊·d/ε) = ε·A/d.

📐 The final formula

C ≈ ε·A/d

  • This is an approximation because the fringing field is neglected.
  • Dimensional check: (farads per meter) times (square meters) divided by (meters) yields farads—dimensionally correct.
  • Parameter effects:
    • Capacitance increases in proportion to permittivity ε.
    • Capacitance increases in proportion to plate area A.
    • Capacitance decreases in proportion to distance d between the plates.

📊 Practical example: printed circuit board capacitance

📊 The scenario

The excerpt provides an example of ground and power planes in a printed circuit board:

  • These planes are separated by a dielectric material.
  • The resulting structure exhibits capacitance.
  • This capacitance acts as an equivalent discrete capacitor in parallel with the power supply.
  • The value may be negligible, significant and beneficial, or significant and harmful—so it's useful to calculate.

🔢 Sample calculation

Given parameters:

  • Dielectric thickness: approximately 1.6 mm
  • Relative permittivity: approximately 4.5
  • Common area between ground and power planes: 25 square centimeters (2.5 × 10⁻³ square meters)

Using the formula C ≈ ε·A/d:

  • ε ≈ 4.5·ε₀ (where ε₀ is the permittivity of free space)
  • Result: equivalent capacitor value is approximately 62.3 picofarads

Example: An organization designing a circuit board can use this formula to predict whether the parasitic capacitance between power and ground planes will affect circuit performance.

72

Capacitance of a Coaxial Structure

5.24 Capacitance of a Coaxial Structure

🧭 Overview

🧠 One-sentence thesis

The capacitance of a coaxial structure can be derived by assuming a charge on the inner conductor, integrating the electric field to find the voltage, and then computing capacitance as the ratio of charge to potential difference, yielding a value that depends only on geometry and material properties.

📌 Key points

  • Derivation strategy: assume a total charge Q+ on the inner conductor, find the associated electric field, integrate to get voltage, then compute capacitance as Q+/V.
  • Key simplification: assuming infinite length (or treating a section of a longer structure) eliminates fringing fields and makes the internal field constant with respect to z.
  • Electric field insight: the field inside a coaxial structure with uniform charge density on the inner conductor is identical to the field of a line charge in free space with the same charge density.
  • Final result depends only on: material permittivity and geometry (inner radius a, outer radius b, length l); it does not depend on charge or voltage, confirming linear behavior.
  • Common confusion: the capacitance per unit length C′ (needed for transmission line models) is obtained by dividing total capacitance by length l.

🔧 Setup and modeling approach

🔧 Physical structure

The coaxial structure consists of:

  • Two concentric perfectly-conducting cylinders with radii a (inner) and b (outer)
  • An ideal dielectric with permittivity ε_s separating them
  • The z-axis placed along the common axis of the cylinders
  • The cylinders described as constant-coordinate surfaces ρ = a and ρ = b

📏 Infinite length assumption

  • Assuming the structure has infinite length (l → ∞) eliminates fringing fields
  • The internal field becomes constant with respect to z
  • For a finite-length capacitor, the central region field closely matches the infinite-length case if fringing field energy is negligible
  • Alternatively: treat length l as one short section of a much longer structure to obtain capacitance per length (exactly what transmission line models need)

⚡ Finding the electric field

⚡ Charge distribution

The charge on the inner conductor is uniformly distributed with density:

ρ_l = Q+ / l (units of C/m)

where Q+ is the total charge on the positively-charged conductor and l is the length.

⚡ Electric field intensity

From Gauss' Law applied to an infinite line charge (Section 5.6), the electric field is:

E = (unit vector ρ) × (ρ_l / (2π ε_s ρ))

This is the radial electric field at distance ρ from the axis.

🔍 Why the line charge result applies

  • Gauss' Law states: surface integral of D · ds = Q_enclosed
  • For a cylindrical surface S with radius ρ < b, the enclosed charge is the same as for a line charge
  • The outer conductor does not change the radial symmetry of the problem
  • Key principle: The electric field inside a coaxial structure with uniform charge density on the inner conductor is identical to the electric field of a line charge in free space having the same charge density

Example: If you place a cylindrical Gaussian surface between the conductors, only the inner conductor's charge is enclosed, and symmetry forces the field to be radial and depend only on ρ.

🔌 Computing voltage and capacitance

🔌 Voltage calculation

Using the definition V = − integral of E · dl along any path C from the negative (outer) conductor to the positive (inner) conductor:

  • Choose the simplest path: a radial line of constant φ and z
  • Integrate from ρ = b (outer) to ρ = a (inner)
  • The result is: V = (ρ_l / (2π ε_s)) × ln(b/a)

The natural logarithm ln(b/a) arises from integrating dρ/ρ.

🔌 Capacitance formula

Using the definition C = Q+ / V:

  • Substitute Q+ = ρ_l × l and the voltage expression
  • The ρ_l terms cancel out in numerator and denominator
  • Final result: C = (2π ε_s l) / ln(b/a)

✅ Properties of the result

PropertyWhat it means
Dimensionally correctUnits are Farads (F)
Depends on materialsThrough permittivity ε_s
Depends on geometryThrough length l, inner radius a, outer radius b
Does NOT depend on charge or voltageConfirms linear (not non-linear) behavior

Don't confuse: The absence of charge/voltage dependence is expected for a linear capacitor; if the formula contained Q or V, it would imply non-linear behavior.

📐 Per-unit-length parameter

📐 Transmission line application

For lumped-element transmission line models (Sections 3.4 and 3.10), divide the total capacitance by length l:

C′ = (2π ε_s) / ln(b/a)

This is the capacitance per unit length, with units of F/m (often expressed as pF/m).

📐 Example: RG-59 coaxial cable

Given:

  • Inner conductor radius a = 0.292 mm
  • Outer conductor radius b = 1.855 mm
  • Polyethylene dielectric with relative permittivity 2.25

Using the per-unit-length formula:

  • ε_s = 2.25 × ε_0 (where ε_0 is the permittivity of free space)
  • C′ ≈ 67.7 pF/m

Example: For a 10-meter length of RG-59 cable, the total capacitance would be approximately 677 pF.

🔄 Context and broader implications

🔄 Circuit board capacitance context

The excerpt begins by noting that structures like circuit boards (with ground and power planes) exhibit capacitance:

  • This capacitance acts as an equivalent discrete capacitor in parallel with the power supply
  • It may be negligible, beneficial, or harmful
  • For a typical board (1.6 mm dielectric thickness, relative permittivity 4.5, 25 cm² area), the equivalent capacitor is about 62.3 pF

🔄 Derivation strategy consistency

The approach used here is the same as for the parallel plate capacitor (Section 5.23):

  1. Assume a charge distribution
  2. Find the electric field
  3. Integrate to get voltage
  4. Compute capacitance as charge/voltage

This consistent methodology can be applied to various capacitive structures.

73

Electrostatic Energy

5.25 Electrostatic Energy

🧭 Overview

🧠 One-sentence thesis

Energy stored in the electric field of a capacitor or capacitive structure can be calculated from capacitance and voltage, and this stored energy directly affects power consumption in electronic systems.

📌 Key points (3–5)

  • Energy storage formula: The energy stored in a capacitor's electric field equals one-half times capacitance times voltage squared.
  • Where energy comes from: Energy is injected by doing work to move charge across a potential difference, building up from zero charge to the final charged state.
  • Energy density perspective: Energy can also be calculated as an integral of energy density over volume, which works even when electric field and permittivity vary with position.
  • Common confusion: Energy increases with permittivity, which makes sense because capacitance is proportional to permittivity—higher capacitance stores more energy at the same voltage.
  • Practical impact: Cyclic charging and discharging of capacitances consumes power in electronic systems; this explains phenomena like why multicore processors don't proportionally increase power consumption.

⚡ How energy gets stored in capacitors

⚡ The charging process

  • A capacitive structure consists of two perfect conductors separated by an ideal dielectric (could be an explicit capacitor or unintended capacitance like a printed circuit board).
  • When potential difference V is applied, positive charge Q+ appears on the higher-potential conductor and negative charge Q− = −Q+ appears on the lower-potential conductor.
  • Energy is stored in the electric field associated with these surface charges.

🔧 Building up from zero charge

The excerpt derives energy by considering the work done to charge the system:

  • Electric potential V is defined as work done (energy injected) per unit charge.
  • For a small charge increment dq, the work contribution is: dW_e = V dq.
  • To go from zero charge to fully charged state Q+, integrate: total work = integral from 0 to Q+ of (q/C) dq = (1/2)(Q+²/C).
  • Since there are no other processes, the energy stored in the electric field equals this work.

The energy stored in the electric field of a capacitor (or a capacitive structure) is given by W_e = (1/2) C V².

📐 Alternative expressions

The energy formula can be written in different forms:

  • In terms of charge and capacitance: W_e = (1/2)(Q+²/C)
  • In terms of capacitance and voltage: W_e = (1/2) C V²
  • These are equivalent because C = Q+/V.

🔋 Energy density and volume integrals

🔋 Parallel plate capacitor example

For a thin parallel plate capacitor:

  • Capacitance C ≈ (ε A)/d, where A is plate area, d is separation, ε is permittivity.
  • Electric field magnitude E relates to voltage: V = E d.
  • Substituting into the energy formula: W_e = (1/2)(ε A/d)(E d)² = (1/2) ε E² (A d).

📦 Energy density concept

  • The product A d is the volume of the capacitor.
  • Since the electric field in a thin parallel plate capacitor is approximately uniform, energy density is also approximately uniform.
  • Energy density w_e = W_e/(A d) = (1/2) ε E², with units of joules per cubic meter (J/m³).

🌐 General volume integral

The energy stored by the electric field present within a volume is given by W_e = (1/2) integral over volume V of ε E² dv.

  • This method works even if E and ε vary with position.
  • Although derived using the parallel-plate example, this result applies generally to any electrostatic field configuration.
  • Energy increases with permittivity of the medium, consistent with capacitance being proportional to permittivity.

💻 Power consumption in electronic systems

💻 Why capacitance matters for power

The excerpt explains several practical implications:

ApplicationRelevance
Capacitors as batteriesNeed to know how much energy can be stored
Unintended capacitanceUses fraction of power supply energy to charge structures
Digital systemsCapacitances are periodically charged and discharged at regular rate
Power consumptionSince power is energy per unit time, cyclic charging/discharging consumes power

⚙️ Multicore processor power analysis

Example 5.11 demonstrates why multicore computing is "power-neutral":

Single-core processor:

  • Energy per clock cycle = (1/2) C₀ V₀²
  • Power consumption P₀ = (1/2) C₀ V₀² f₀, where f₀ is clock frequency and C₀ is sum capacitance.

N-core processor:

  • Sum capacitance increases by factor N (capacitors in parallel add).
  • Clock frequency decreases by factor N (same computation distributed among N cores).
  • Power consumption P_N = (1/2)(N C₀) V₀² (f₀/N) = P₀.

Key insight:

  • The increase in power from hardware replication is nominally offset by the decrease in power from reducing clock rate.
  • Total energy at any given time is N times higher, but capacitances need recharging 1/N times as often.
  • Power consumption is proportional to f₀ only and independent of N.
  • Don't confuse: The usual reason for multicore design is to increase total computation (increase f₀ N product), but it's helpful that power scales with frequency alone.
74

Convection and Conduction Currents

6.1 Convection and Conduction Currents

🧭 Overview

🧠 One-sentence thesis

Two distinct physical mechanisms—convection (mechanical forces) and conduction (electric field response)—produce current, and only conduction current obeys Ohm's Law, making the distinction critical for electromagnetic analysis.

📌 Key points (3–5)

  • Two mechanisms: convection current arises from mechanical forces moving charged particles; conduction current arises from the electric field guiding them.
  • Convection example: a cloud of free electrons blown by wind through the atmosphere.
  • Conduction mechanism: the electric field can dislodge weakly-bound electrons, which travel some distance before reassociating with other atoms—individual electrons do not necessarily travel the full current path.
  • Common confusion: not all current is the same—Ohm's Law applies only to conduction current, not convection current.
  • Why it matters: distinguishing the two is essential for correct electromagnetic analysis and applying the right laws.

⚡ Two types of current

⚡ Convection current

Convection current consists of charged particles moving in response to mechanical forces, as opposed to being guided by the electric field.

  • The driving force is mechanical, not electrical.
  • Particles are carried along by the motion of the surrounding material.
  • Example: a cloud bearing free electrons moves through the atmosphere because wind (a mechanical force) pushes it, not because an electric field pulls the electrons.

⚡ Conduction current

Conduction current consists of charged particles moving in response to the electric field and not merely being carried by motion of the surrounding material.

  • The driving force is the electric field.
  • The field can dislodge weakly-bound electrons from atoms; these electrons then travel some distance before reassociating with other atoms.
  • Important detail: individual electrons do not necessarily travel the full distance over which the current is perceived to exist—they hop from atom to atom.
  • Don't confuse: conduction is not just "electrons flowing in a wire"; it is specifically the electric field causing the motion, not mechanical transport.

🔍 Why the distinction matters

🔍 Ohm's Law applies only to conduction

  • Ohm's Law specifies the relationship between electric field intensity and current.
  • The excerpt emphasizes: Ohm's Law applies only to conduction current.
  • This means:
    • For convection current, you cannot use Ohm's Law to relate field and current.
    • For conduction current, Ohm's Law is the correct tool.
  • Example: if you analyze a cloud of electrons blown by wind (convection), Ohm's Law does not describe the current; if you analyze electrons drifting in a conductor under an applied field (conduction), Ohm's Law does apply.

🔍 Electromagnetic analysis requires correct identification

  • The excerpt states the distinction is "important in electromagnetic analysis."
  • Misidentifying the type of current leads to applying the wrong physical laws.
  • Practical implication: before using Ohm's Law, confirm that the current is conduction, not convection.

🧲 Conduction mechanism details

🧲 Electron dislodging and reassociation

  • In conduction, the electric field does more than just push free electrons—it can also dislodge weakly-bound electrons from atoms.
  • Once dislodged, these electrons travel some distance, then reassociate with other atoms.
  • This means the same electron does not travel the entire length of the conductor; instead, the current is a relay of electrons hopping from atom to atom.
  • Don't confuse: "current flows" does not mean "the same electrons flow all the way"; it means charge is transferred step-by-step through the material.

🧲 Perceived vs actual electron travel

  • The current is perceived to exist over a certain distance (e.g., the length of a wire).
  • But individual electrons may only travel a short segment before reassociating.
  • The net effect is continuous current, even though no single electron completes the full journey.
75

Current Distributions

6.2 Current Distributions

🧭 Overview

🧠 One-sentence thesis

Current distributions extend the elementary circuit-theory view of current as a scalar along wires to a vector quantity that can spread over surfaces and volumes, requiring mathematical descriptions as line current, surface current density, or volume current density depending on the physical situation.

📌 Key points (3–5)

  • Why vector current matters: many real problems (RF interconnects, grounding, lightning) involve current that spreads beyond thin wires, so scalar circuit theory is insufficient.
  • Three distribution types: line current (constrained path), surface current density (distributed over a surface), and volume current density (distributed within a volume).
  • Volume density to total current: integrating volume current density over a surface yields the total current through that surface (a flux calculation).
  • Common confusion: in elementary circuits, current is scalar with sign indicating direction; in electromagnetic analysis, current becomes a vector field that varies with position.
  • Surface choice is arbitrary: when calculating total current from volume density, any cross-sectional surface with edges at the conductor perimeter gives the same result.

🔌 From circuits to fields

🔌 Elementary circuit view

In elementary electric circuit theory, current is the rate at which electric charge passes a particular point in a circuit (e.g., 1 A = 1 C per second).

  • Current is treated as a scalar quantity.
  • Only two possible directions because charge flows along defined channels (wires, traces).
  • Direction indicated by sign relative to a reference direction tied to voltage polarity.
  • Sign convention: positive current = flow of positive charge into the positive terminal of a passive device (resistor, capacitor, inductor); for active devices (batteries), positive current flows out of the positive terminal.

🌐 Why the circuit view breaks down

The "lumped device + thin filament wire" model fails for:

  • Wire and pin interconnects at radio frequencies
  • Circuit board and enclosure grounding
  • Physical phenomena like lightning

In these cases, current is not constrained to thin paths—it spreads over surfaces and volumes.

🧭 The electromagnetic solution

  • Define current as a vector quantity.
  • Consider spatial distributions of current (how current varies with position).
  • Use three mathematical models depending on geometry: line, surface, and volume distributions.

📏 Three types of current distribution

📏 Line current

Line current: current constrained to follow a particular path, specified mathematically as l-hat times I, where l-hat is the direction unit vector.

  • The direction l-hat may vary with position along the path.
  • Example: straight wire → l-hat is constant; coil → l-hat varies with position along the coil.
  • This is the closest analog to circuit theory, but now direction is explicit.

📐 Surface current density

Surface current density J_s: total current divided by the width (or perimeter) over which it is distributed on a surface (units: A/m).

  • Used when current spreads over a surface rather than a line.
  • Formula: J_s = u-hat times (I / (2 pi a)), where I is total current, 2 pi a is the circumference, and u-hat is the flow direction.
  • Example: radio-frequency current on a high-conductivity metal wire of radius a can be modeled as uniform surface current on the wire surface.
  • The current "lives" on the surface, not inside the wire.

📦 Volume current density

Volume current density J: current per unit area passing through a small planar surface within a volume, defined as the limit as the surface area shrinks to zero: J = u-hat times (di / ds) (units: A/m²).

  • J is a function of position within the volume.
  • To find total current through a surface S, integrate: I = integral over S of (J · ds).
  • This is a flux calculation: volume current density integrated over a surface yields total current through that surface.
  • Don't confuse: J is the density (current per area); I is the total current (the integral of J).

🧮 Calculating total current from volume density

🧮 The integration formula

I = integral over S of (J · ds)

  • S is any surface through which you want to measure current.
  • ds is the differential area element (a vector pointing normal to the surface).
  • The dot product J · ds accounts for the angle between current flow and the surface normal.

🔍 Example: wire with circular cross-section

Given:

  • Wire radius a = 3 mm
  • Uniform volume current density J = z-hat times 8 A/m² (pointing along the wire axis)
  • Find net current I through the wire

Solution approach:

  • Choose surface S = cross-section perpendicular to the wire axis.
  • Choose ds to point in the z-hat direction (matching the sign convention).
  • In cylindrical coordinates: ds = z-hat times rho d-rho d-phi.
  • The integral becomes: I = (8 A/m²) times (integral from 0 to a, 0 to 2pi of rho d-rho d-phi).
  • The integral evaluates to pi a² (the cross-sectional area).
  • Result: I = (8 A/m²) times (pi times (0.003 m)²) = 226 micro-amps.

🔄 Surface choice is arbitrary

  • The answer is independent of which cross-sectional surface you choose.
  • Example: a surface tilted 45° from the wire axis would have larger area, but the dot product J · ds would be proportionally smaller, yielding the same result.
  • Practical tip: choose the surface that makes the math simplest (usually perpendicular to the current flow).

🔗 Connection to other concepts

🔗 Relation to convection vs conduction

  • The excerpt on current distributions applies to both convection and conduction currents (defined in Section 6.1).
  • However, the distinction matters for Ohm's Law (Section 6.3), which applies only to conduction current.

🔗 Relation to conductivity

  • Conductivity (Section 6.3) determines how volume current density J responds to an applied electric field.
  • The current distribution formalism (line, surface, volume) describes where current flows; conductivity describes how much current flows in response to a field.
76

Conductivity

6.3 Conductivity

🧭 Overview

🧠 One-sentence thesis

Conductivity is a material property that directly relates how much conduction current flows in response to an applied electric field, without requiring separate calculations of force and charge mobility.

📌 Key points (3–5)

  • What conductivity determines: the conduction current density that arises when an electric field is applied to a material.
  • Ohm's Law for Electromagnetics: current density J (A/m²) equals conductivity σ (S/m) times electric field intensity E (V/m).
  • Range of materials: from perfect insulators (σ = 0, no available charge) to perfect conductors (σ → ∞, zero electric field inside).
  • Common confusion: conductivity σ is a material property (S/m), while resistance R is a device property (Ω); Ohm's Law for devices (V = IR) is a special case of the electromagnetic version.
  • What determines conductivity: both the availability of charge and the mobility of charge within the material's atomic/molecular structure.

🔌 What conductivity is and why it matters

🔌 The core idea

Conductivity is a property of materials that determines conduction current density in response to an applied electric field.

  • Conduction current is the flow of charge in response to an electric field.
  • Although you can calculate the force on charge straightforwardly, that force alone does not tell you the speed at which charge moves.
  • The speed depends on charge mobility, which is determined by the material's atomic and molecular structure.
  • Conductivity packages both availability and mobility into a single parameter, so you don't have to grapple with force and mobility separately.

⚡ Ohm's Law for Electromagnetics

The relationship is:

J = σE (Equation 6.4)

where:

  • E = electric field intensity (V/m)
  • J = volume current density, a vector describing current flow (A/m²)
  • σ = conductivity (S/m, where 1 siemens = 1 Ω⁻¹)

Important: This law applies only to conduction current, not convection current, displacement current, or other forms.

Example: If you apply an electric field of 10 V/m to a material with conductivity 2 S/m, the resulting current density is 20 A/m².

🔄 Relation to circuit Ohm's Law

  • The familiar circuit version is I = V/R (current in amperes = voltage in volts / resistance in ohms).
  • The excerpt states this is "in fact a special case" of the electromagnetic version J = σE.
  • Don't confuse: σ (S/m) is a material property; R (Ω) is a device property (see Section 6.4 for the full derivation).

🌈 Material classes by conductivity

🌌 Perfect insulators (σ = 0)

  • A perfect vacuum ("free space") contains no charge, so conductivity is zero.
  • No current can flow, no matter what electric field is applied.

🛡️ Good insulators (σ ≪ 10⁻¹⁰ S/m)

  • Conductivities so low that resulting currents can usually be ignored.
  • Air is the most important example: conductivity only slightly greater than free space.
  • Lossless dielectrics: a special class well-characterized by permittivity ε alone; typically μ = μ₀ and σ = 0 can be assumed.

🔻 Poor insulators

  • Low conductivity, but high enough that currents cannot be ignored.
  • Example: dielectric material separating conductors in a transmission line—must account for loss per length.
  • Lossy dielectrics: characterized by both ε and σ; typically μ = μ₀.

🔶 Semiconductors (10⁻⁴ to 10⁺¹ S/m)

  • Materials used in integrated circuits.
  • Intermediate conductivities between insulators and conductors.

⚙️ Good conductors (σ > 10⁵ S/m)

  • Very high conductivities.
  • Metals (aluminum, copper, gold alloys) reach conductivities on the order of 10⁸ S/m.
  • Minuscule electric fields give rise to large currents.
  • No significant energy storage, so permittivity is not relevant for good conductors.

⚡ Perfect conductors (σ → ∞)

A perfect conductor is a material for which σ → ∞, E → 0, and subsequently V (the electric potential) is constant.

  • Do not interpret as "any field gives infinite current"—that is not plausible.
  • Instead: E = J/σ → 0 throughout the material, meaning E is zero independently of any current flow.
  • Important consequence: the potential field V is constant throughout the material (recall E = −∇V, so constant V means E = 0).
  • The volume is called an equipotential volume.
  • Useful as an approximation: metals are often modeled as perfectly-conducting equipotential volumes to simplify analysis.

📊 Summary table

Material classConductivity σ (S/m)Key characteristics
Perfect insulator0No charge available (e.g., vacuum)
Good insulator≪ 10⁻¹⁰Currents negligible (e.g., air, lossless dielectrics)
Poor insulatorLow but non-negligibleCurrents matter (e.g., lossy dielectrics)
Semiconductor10⁻⁴ to 10⁺¹Intermediate (e.g., IC materials)
Good conductor> 10⁵Large currents from small fields (e.g., metals)
Perfect conductor→ ∞E = 0, V constant (equipotential volume)

⚠️ Important qualifications and cautions

⚠️ Context-dependent terms

  • Terms like "good conductor" and "poor insulator" are qualitative and context-dependent.
  • What counts as a "good insulator" in one application may be a "poor insulator" in another.

⚠️ Linearity assumption

  • The excerpt assumes materials are linear (as summarized in Section 2.8).
  • Example of non-linearity: air is normally a good insulator, but under sufficiently large potential difference (e.g., lightning), current flows readily.
  • This is called dielectric breakdown.
  • Non-linearity can appear even before breakdown, so be careful with strong electric fields.

⚠️ What conductivity depends on

Conductivity σ depends on:

  1. Availability of charge within the material.
  2. Mobility of charge—how easily charge-bearing constituents can move.

Both are determined by the material's atomic and molecular structure.

77

Resistance

6.4 Resistance

🧭 Overview

🧠 One-sentence thesis

Resistance characterizes how a device's conductivity and geometry together determine the relationship between voltage and current according to Ohm's Law for Devices (V = IR), and it arises specifically from limited conductivity rather than simply from any voltage-current ratio.

📌 Key points (3–5)

  • What resistance is: a property of devices (not materials) that describes conductivity in terms of V = IR, arising from the conductivity of the materials and the device geometry.
  • Common confusion: resistance always produces real-valued impedance, but not all real-valued impedance represents resistance—impedance can be real even in perfect conductors (e.g., transmission lines).
  • How to calculate resistance: for a cylindrical conductor, R = l / (σA), where l is length, σ is conductivity, and A is cross-sectional area.
  • Geometry matters: resistance depends on both material conductivity and device shape—longer length increases resistance, larger cross-section decreases it.
  • Frequency effects: at higher frequencies, current concentrates near the surface, reducing effective cross-sectional area and increasing resistance.

🔍 What resistance means

🔍 Resistance as a device property

Resistance R (Ω) is a characterization of the conductivity of a device (as opposed to a material) in terms of Ohm's Law for Devices; i.e., V = IR.

  • Resistance is not a material property like conductivity σ (S/m).
  • It is a property of complete devices: resistors, wires, and most practical electronic components.
  • Resistance manifests the conductivity of the materials comprising the device, leading to the V = IR relationship.

⚠️ Resistance vs impedance

The excerpt emphasizes a critical distinction:

Resistance is not necessarily the real part of impedance.

Why this matters:

  • Impedance Z is defined as the ratio V/I (or Ṽ/Ĩ in phasor domain).
  • Many devices can be characterized by this ratio, not just resistive devices.
  • Example from the excerpt: a terminated transmission line can have real-valued impedance even when made of perfect conductors.

Key principle:

Resistance results in a real-valued impedance. However, not all devices exhibiting a real-valued impedance exhibit resistance.

The core distinction:

Resistance pertains to limited conductivity, not simply to voltage-current ratio.

  • The real component of complex-valued impedance does not necessarily represent resistance.
  • Resistance specifically describes energy dissipation due to limited conductivity.

🧮 Calculating resistance

🧮 The cylindrical wire example

The excerpt derives resistance for a straight wire of length l, radius a, and conductivity σ:

Setup:

  • Cylinder of material with conductivity σ, cross-sectional radius a, length l along z-axis.
  • Perfectly-conducting plates at the ends (where terminals attach).
  • Battery creates uniform internal electric field E.

Step 1: Find voltage V

  • Use V = −∫_C E · dl from Section 5.8.
  • Path C goes from z = 0 to z = l along the axis.
  • Electric field E = −ẑE_z (points in −ẑ direction, same as current flow).
  • Result: V = E_z · l

Step 2: Find current I

  • Use I = ∫_S J · ds (Section 6.2).
  • Surface S is the cross-section perpendicular to the wire axis.
  • From Ohm's Law for Electromagnetics: J = σE = −ẑσE_z
  • Integrating over circular cross-section: I = σE_z(πa²)

Step 3: Calculate resistance

  • R = V/I = (E_z · l) / [σE_z(πa²)]
  • Simplifies to: R = l / [σ(πa²)]

📐 General formula for cylindrical conductors

Recognizing that πa² is the cross-sectional area A:

R = l / (σA) [Equation 6.7]

The resistance of a right cylinder of material is proportional to length and inversely proportional to cross-sectional area and conductivity.

What this means:

  • Longer wire → higher resistance (proportional to l)
  • Thicker wire → lower resistance (inversely proportional to A)
  • Better conductor → lower resistance (inversely proportional to σ)
  • This formula works for any cross-sectional shape, not just circular.

⚠️ Important assumption

Equation 6.7 assumes that volume current density J is uniform over the cross-section.

When this is valid:

  • Thin wires at "low" frequencies, including DC.
  • Excellent approximation for most common applications.

When this breaks down:

  • At higher frequencies, J is not uniformly distributed.
  • At sufficiently high frequencies, current is limited to the exterior surface (skin effect).
  • Effective area A is reduced → resistance R increases with frequency.
  • Quantifying high-frequency behavior requires concepts beyond electrostatics.

📊 Worked examples

📊 Example 1: Hookup wire

Problem: 22 AWG copper solid-conductor hookup wire, 3 m long. Diameter = 0.644 mm, copper conductivity σ = 58 MS/m. Find resistance.

Solution:

  • Diameter 2a = 0.644 mm → radius a = 0.322 mm
  • Cross-sectional area: A = πa² ≈ 3.26 × 10⁻⁷ m²
  • Given: σ = 58 × 10⁶ S/m, l = 3 m
  • Using R = l/(σA): R = 159 mΩ

📊 Example 2: Steel pipe

Problem: Pipe 3 m long, inner radius 5 mm, outer radius 7 mm. Steel conductivity = 4 MS/m. Find DC resistance.

Solution:

  • Current flows through the annular cross-section (not the hollow interior).
  • Cross-sectional area: A = π(b² − a²) = π(7² − 5²) mm² ≈ 7.54 × 10⁻⁵ m²
  • Given: σ = 4 × 10⁶ S/m, l = 3 m
  • Using R = l/(σA): R ≈ 9.95 mΩ

Key insight: Only the material between inner and outer radii conducts current; the hollow center does not contribute to conductance.

🔗 Connection to perfect conductors

🔗 Perfect conductors as a limiting case

The excerpt mentions perfect conductors as context:

A perfect conductor is a material for which σ → ∞, E → 0, and subsequently V (the electric potential) is constant.

Implications:

  • When conductivity σ approaches infinity, resistance R = l/(σA) approaches zero.
  • Electric field E inside a perfect conductor is zero.
  • The entire volume is an equipotential volume (constant V throughout).
  • Metals are often modeled as perfectly-conducting equipotential volumes to simplify analysis.

Don't confuse: This is an idealization. Real conductors have finite σ and therefore non-zero resistance, even if very small.

78

Conductance

6.5 Conductance

🧭 Overview

🧠 One-sentence thesis

Conductance, the reciprocal of resistance, is useful in engineering analysis when voltage is the stimulus and current is the response, and it simplifies calculations for devices in parallel.

📌 Key points (3–5)

  • Definition: Conductance G (measured in Ω⁻¹ or S) is the reciprocal of resistance R, depending on both material conductivity and device geometry.
  • Why both concepts exist: Conductance is not strictly required since it's just 1/R, but it offers intuitive advantages in certain contexts.
  • When conductance is preferred: When voltage is the independent stimulus and current is the response; when analyzing parallel device combinations (conductances simply add).
  • Common confusion: Conductance vs conductivity—conductance is a device property (like resistance), while conductivity is a material property.
  • Practical application: Conductance per unit length (G′) is used to determine characteristic impedance in transmission lines.

🔄 Relationship to resistance

🔄 Mathematical definition

Conductance G (Ω⁻¹ or S) is the reciprocal of resistance R.

  • The formula is simply: G = 1/R
  • Like resistance, conductance depends on:
    • Material conductivity (σ)
    • Device geometry (shape and dimensions)
  • Units: Siemens (S) or inverse ohms (Ω⁻¹)

🤔 Why have both concepts?

The excerpt addresses a natural question: if conductance is just the reciprocal of resistance, why do we need it?

Short answer: We don't strictly need both—one would suffice mathematically.

Two practical reasons for using conductance:

  1. Intuitive physics: More natural when voltage is the independent "stimulus" and current is the "response"
    • Example: In transmission line models, conductance appears because voltage drives the current leakage
  2. Parallel combinations: Conductances add directly when devices are in parallel
    • Total conductance = sum of individual conductances
    • This is simpler than the reciprocal formula needed for resistances in parallel

🔌 Coaxial structure example

🎯 Problem setup

The excerpt works through determining conductance of a coaxial structure:

  • Two concentric perfectly-conducting cylinders
  • Inner radius: a
  • Outer radius: b
  • Separated by lossy dielectric with conductivity σₛ
  • Goal: find conductance per unit length G′ (S/m)

🛠️ Solution approach

The excerpt presents an alternative method (not the capacitance-analogy approach):

Step-by-step procedure:

  1. Assume a leakage current I between conductors
  2. Determine current density J using geometry
  3. Calculate electric field E from J/σₛ (using Ohm's Law for Electromagnetics)
  4. Integrate E along a path between conductors to get voltage V
  5. Find conductance as G = I/V

📐 Key geometric insight

Current density must flow radially outward with circular symmetry:

  • J = (direction) × I/A
  • Area A = circumference × length = 2πρl
  • Therefore: J flows radially with density I/(2πρl)
  • Density diminishes inversely with area (like flux spreading out)

📊 Final result

The conductance per unit length is:

G′ = (2πσₛ) / ln(b/a)

Where:

  • σₛ = conductivity of the dielectric material
  • b = outer radius
  • a = inner radius
  • ln = natural logarithm

Important properties:

  • Dimensionally correct (units of S/m)
  • Depends only on materials (σₛ) and geometry (a, b)
  • Does NOT depend on current or voltage (linear behavior)
  • Used in transmission line characteristic impedance calculations

🔢 Numerical example: RG-59 cable

Given parameters:

  • Inner conductor radius: 0.292 mm
  • Outer conductor radius: 1.855 mm
  • Polyethylene conductivity: 5.9 × 10⁻⁵ S/m

Result: G′ ≈ 200 μS/m (microsiemens per meter)

🔍 Comparison with resistance

PropertyResistanceConductance
SymbolRG
UnitsΩ (ohms)S (siemens) or Ω⁻¹
RelationshipG = 1/RR = 1/G
Intuitive when...Current is stimulusVoltage is stimulus
Parallel combinationComplex (reciprocals)Simple (direct sum)
Series combinationSimple (direct sum)Complex (reciprocals)

Don't confuse: Conductance (device property, G) with conductivity (material property, σ). Conductance depends on both the material's conductivity AND the device's geometry.

79

Power Dissipation in Conducting Media

6.6 Power Dissipation in Conducting Media

🧭 Overview

🧠 One-sentence thesis

Power dissipation in conducting media arises from the displacement of charge carriers against resistance, converting electrical potential energy into heat at a rate proportional to conductivity and the square of the electric field magnitude.

📌 Key points (3–5)

  • Core mechanism: charge displacement in response to electric fields reduces potential energy at a steady rate, which manifests as power dissipation.
  • Joule's Law: power dissipation is calculated by integrating the dot product of electric field and current density over the volume of interest.
  • Key formula: power equals the integral of conductivity times electric field magnitude squared over volume (P = ∫ σ|E|² dv).
  • Common confusion: perfect conductors (σ → ∞) have zero electric field for any current, so they dissipate no power, even though conductivity is infinite.
  • Energy fate: dissipated power converts to heat (joule heating or ohmic heating), which is how toasters and heaters work and why electronics generate heat.

⚡ Energy and work fundamentals

⚡ Work and potential energy

Work (ΔW): force times distance traversed in response to that force (units: joules).

  • The excerpt distinguishes two perspectives on work:
    • In earlier sections: work represented external energy used to move charge "upstream" (increasing potential energy).
    • Here: ΔW represents internal energy escaping the system as kinetic energy (decreasing potential energy).
  • Sign convention: positive ΔW now means a reduction in potential energy, hence the explicit "+" sign in the formula ΔW = +F · Δl.
  • The displacement of charge in an electric field constitutes a reduction in the system's potential energy.

⚙️ Power as energy rate

  • Power (ΔP) is work divided by time: ΔP = ΔW / Δt = (F · Δl) / Δt (units: watts).
  • If charge is part of a steady current, energy loss occurs at a steady rate.
  • Power dissipation is the term for this energy loss associated with current flow.

🔬 Deriving Joule's Law

🔬 From force to current density

The excerpt builds the power formula step by step:

  1. Force from electric field: F = qE, where q is charge and E is electric field intensity.
  2. Charge in a volume cell: q = ρᵥ Δv, where ρᵥ is volume charge density and Δv is the cell volume.
  3. Substituting: F = ρᵥ Δv E.
  4. Power expression: ΔP = (ρᵥ Δv E · Δl) / Δt = E · (ρᵥ Δl / Δt) Δv.
  5. Recognizing current density: the quantity (ρᵥ Δl / Δt) has units of A/m², which is the volume current density J.
  6. Result: ΔP = E · J Δv.

📐 Joule's Law integral form

Joule's Law: P = ∫ E · J dv (integrated over volume V).

  • In the limit as Δv → 0, we have dP = E · J dv.
  • Integrating over the volume of interest gives the total power dissipation.
  • This is the general expression for power dissipation in conducting media.

🧮 Using Ohm's Law for Electromagnetics

  • Ohm's Law for Electromagnetics: J = σE, where σ is conductivity.
  • Substituting into Joule's Law: P = ∫ E · (σE) dv = ∫ σ|E|² dv.
  • Final formula: P = ∫ σ|E|² dv.
  • Power is proportional to conductivity (σ) and proportional to the electric field magnitude squared (|E|²).

🔌 Special cases and circuit theory

🔌 Perfect conductors

  • For a perfect conductor: σ → ∞ and E → 0 no matter how much current is applied.
  • The formula P = ∫ σ|E|² dv is not helpful because both σ and E change.
  • However, since E → 0 regardless of current, from Joule's Law (P = ∫ E · J dv) we conclude:
    • No power is dissipated in a perfect conductor.
  • Don't confuse: infinite conductivity does not mean infinite power; the electric field drops to zero, so dissipation is zero.

🔌 Connection to circuit theory

The excerpt shows how the general formula reduces to the familiar P = IV:

StepExpressionExplanation
StartP = ∫ σE²ᵤ dvAssume E = ẑEᵤ (field in z-direction)
Factor outP = σE²ᵤ ∫ dvConductivity and field are uniform
VolumeP = σE²ᵤ A lVolume = cross-sectional area A × length l
ReorganizeP = (σEᵤ A)(Eᵤ l)Group terms strategically
Identify IσEᵤ A = ICurrent density × area = total current
Identify VEᵤ l = VElectric field × length = potential difference
ResultP = IVFamiliar circuit theory formula
  • Example: for a cylindrical conductor aligned along the z-axis, the general electromagnetic formula simplifies to the elementary circuit result.

🔥 Physical interpretation and applications

🔥 What happens to the energy

Joule heating (or ohmic heating): the power dissipation associated with current flow in any material that is not a perfect conductor manifests as heat.

  • Charge carriers are displaced through the material, but conductivity is finite because other constituents "get in the way."
  • Energy is used to displace charge-bearing particles, and this motion of particles is observed as heat.
  • Heat is essentially the motion of constituent particles.

🔥 Practical applications

  • Devices that generate heat: toasters, electric space heaters, and many other heating devices operate by converting electrical energy to heat via joule heating.
  • Electronics heat generation: all practical electronic devices generate heat because they are not perfect conductors.
  • This phenomenon is both useful (for heating applications) and a challenge (for thermal management in electronics).
80

Comparison of Electrostatics and Magnetostatics

7.1 Comparison of Electrostatics and Magnetostatics

🧭 Overview

🧠 One-sentence thesis

Electrostatics and magnetostatics exhibit a formal similarity called duality, where corresponding concepts and equations mirror each other in structure and behavior.

📌 Key points (3–5)

  • What duality means: electrostatics and magnetostatics share parallel structures in their sources, fields, equations, and energy storage mechanisms.
  • How the fields correspond: electric field intensity E pairs with magnetic field intensity H; electric flux density D pairs with magnetic flux density B.
  • Maxwell's Equations mirror each other: integral and differential forms show structural parallels between the two domains.
  • Common confusion: the "-statics" suffix suggests only DC conditions, but many magnetostatics principles apply at AC frequencies as well.
  • Why it matters: recognizing duality helps transfer understanding from electrostatics to magnetostatics and reveals deeper symmetries in electromagnetic theory.

🔄 The concept of duality

🔄 What duality is

Duality: the technical term for the formal similarities between electrostatics and magnetostatics.

  • Duality is not just superficial resemblance; it is a structural correspondence between the two theories.
  • The same concept of duality also exists between voltage and current in electrical circuit theory.
  • Understanding one domain helps predict and understand the parallel concepts in the other.

📖 Scope of magnetostatics

  • Magnetostatics is the theory of the magnetic field when its behavior is independent of electric fields.
  • The "-statics" ending means sources are time-invariant (DC conditions).
  • Don't confuse: despite the name, many magnetostatics aspects are applicable at AC frequencies as well.

⚡ Parallel sources and fields

⚡ Sources of the fields

ElectrostaticsMagnetostatics
Static chargeSteady current, magnetizable material
  • In electrostatics, the source is stationary charge.
  • In magnetostatics, the sources are steady (time-invariant) currents and materials that can be magnetized.

🌊 Field intensities and flux densities

QuantityElectrostaticsMagnetostatics
Field intensityE (V/m)H (A/m)
Flux densityD (C/m²)B (Wb/m² = T)
  • Electric field intensity E corresponds to magnetic field intensity H.
  • Electric flux density D corresponds to magnetic flux density B.
  • Units differ, but the conceptual roles are parallel.

🧲 Material relations

ElectrostaticsMagnetostatics
D = εEB = μH
J = σE(current relation)
  • Both domains relate flux density to field intensity through material properties (permittivity ε in electrostatics, permeability μ in magnetostatics).
  • Conductivity σ relates current density J to electric field E in electrostatics.

🧮 Parallel equations

🧮 Maxwell's Equations: integral form

ElectrostaticsMagnetostatics
D · ds = Q_enclB · ds = 0
E · dl = 0H · dl = I_encl
  • The first row shows Gauss's laws: electric flux through a closed surface equals enclosed charge; magnetic flux through a closed surface is always zero.
  • The second row shows circulation laws: electric field circulation is zero (conservative field); magnetic field circulation equals enclosed current.

🧮 Maxwell's Equations: differential form

ElectrostaticsMagnetostatics
∇ · D = ρ_v∇ · B = 0
∇ × E = 0∇ × H = J
  • Divergence equations: electric flux density divergence equals volume charge density; magnetic flux density has zero divergence.
  • Curl equations: electric field has zero curl; magnetic field curl equals current density.

🔗 Boundary conditions

ElectrostaticsMagnetostatics
× [E₁ − E₂] = 0 × [H₁ − H₂] = J_s
· [D₁ − D₂] = ρ_s · [B₁ − B₂] = 0
  • Tangential component conditions: tangential electric field is continuous across boundaries; tangential magnetic field discontinuity equals surface current density.
  • Normal component conditions: normal electric flux density discontinuity equals surface charge density; normal magnetic flux density is continuous.

💾 Energy and dissipation parallels

💾 Energy storage mechanisms

ElectrostaticsMagnetostatics
CapacitanceInductance
w_e = (1/2)εE²w_m = (1/2)μH²
  • Capacitance stores energy in electric fields; inductance stores energy in magnetic fields.
  • Energy density formulas are structurally identical: half the product of material property and field intensity squared.
  • Example: a capacitor stores energy through separated charges creating an electric field; an inductor stores energy through current creating a magnetic field.

💾 Energy dissipation

  • Both domains include resistance as a mechanism for energy dissipation.
  • The table lists resistance under energy dissipation, showing that real materials convert electromagnetic energy to heat in both contexts.

B1-Comparison-of-Electrostatics-and-Magnetostatics

81

Gauss' Law for Magnetic Fields: Integral Form

7.2 Gauss’ Law for Magnetic Fields: Integral Form

🧭 Overview

🧠 One-sentence thesis

Gauss' Law for Magnetic Fields states that magnetic flux through any closed surface is zero, which means magnetic field lines always form closed loops and have no localizable source analogous to electric charge.

📌 Key points (3–5)

  • The law itself: the integral of magnetic flux density B over any closed surface equals zero.
  • What it means for field lines: every magnetic field line entering a closed surface must also exit—field lines cannot begin or end inside the volume.
  • No magnetic "charge": unlike electric fields that start at charged particles, magnetic fields have no particle or structure that acts as a source point.
  • Common confusion: when we say current is the "source" of a magnetic field, we mean only that the field coexists with current, not that field lines are attached to or originate from the current.
  • Duality with electrostatics: this law is one of four Maxwell's Equations and highlights a key difference between magnetostatics and electrostatics.

🧲 The integral form and its meaning

🧲 Mathematical statement

Gauss' Law for Magnetic Fields (Equation 7.1): the flux of the magnetic field through a closed surface is zero.

  • Written as: the closed surface integral of B · ds = 0
  • B is magnetic flux density (units: Wb/m² or Tesla)
  • S is a closed surface
  • ds is the outward-pointing differential surface normal

📐 Units and interpretation

  • B has units of Wb/m² (webers per square meter)
  • Integrating B over a surface gives a quantity with units of Wb (webers), which is magnetic flux
  • The law says this total flux is always zero for any closed surface

🔁 Field lines form closed loops

🔁 What zero flux requires

  • For the magnetic flux through a closed surface to be zero, every field line entering the volume enclosed by S must also exit this volume
  • Field lines may not begin or end within the volume
  • The only way this can be true for every possible surface S is if magnetic field lines always form closed loops
  • This is what we observe in practice

🧲 Bar magnet example

The excerpt describes a two-dimensional bar magnet with two surfaces:

  • Surface S_A: every field line entering S also leaves S, so the flux through S is zero
  • Surface S_B: every field line within S remains in S (loops back), so the flux through S is again zero

In both cases, the net flux is zero because field lines do not terminate—they loop.

⚡ Contrast with electric fields

⚡ No magnetic "charge"

  • GLM implies there can be no particular particle or structure that can be the source of the magnetic field
  • If such a source existed, it would be a start point for field lines, violating the closed-loop requirement
  • This is very different from the electrostatic field: every electric field line begins at a charged particle

🔄 What "source" means for magnetic fields

  • When we say current (for example) is the source of the magnetic field, we mean only that the field coexists with current
  • We do not mean that the magnetic field is somehow attached to the current
  • There is no "localizable" quantity analogous to charge for electric fields associated with magnetic fields

Don't confuse: "source" in magnetism ≠ "source" in electrostatics. Magnetic fields have no point of origin; they only coexist with currents or magnetizable materials.

📊 Duality with electrostatics

📊 Comparison table

The excerpt provides a duality table. Key rows relevant to Gauss' Law:

AspectElectrostaticsMagnetostatics
SourcesStatic chargeSteady current, magnetizable material
Flux densityD (C/m²)B (Wb/m² = T)
Maxwell's Eq. (integral)Closed surface integral of D · ds = Q_enclClosed surface integral of B · ds = 0
Maxwell's Eq. (differential)Divergence of D = ρ_vDivergence of B = 0
Boundary Conditionsn · [D₁ − D₂] = ρ_sn · [B₁ − B₂] = 0

🔍 Key differences

  • Electrostatics: the flux of D through a closed surface equals the enclosed charge Q_encl (nonzero if charge is present)
  • Magnetostatics: the flux of B through a closed surface is always zero (no enclosed "magnetic charge")
  • This reflects the fundamental difference: electric field lines start and end at charges; magnetic field lines form closed loops

🌐 Context and importance

🌐 One of Maxwell's Equations

  • Gauss' Law for Magnetic Fields is one of the four fundamental laws of classical electromagnetics
  • Collectively, these are known as Maxwell's Equations
  • The excerpt notes that GLM emerges from the "flux density" interpretation of the magnetic field (covered in an earlier section)

🧠 Why magnetic fields are "weird"

  • The excerpt emphasizes: there is no localizable quantity (analogous to charge for electric fields) associated with magnetic fields
  • This is described as "just another way in which magnetic fields are weird"
  • The weirdness stems from the absence of magnetic monopoles (isolated north or south poles) in classical electromagnetics
82

7.3 Gauss' Law for Magnetism: Differential Form

7.3 Gauss’ Law for Magnetism: Differential Form

🧭 Overview

🧠 One-sentence thesis

The differential form of Gauss' Law for magnetism states that the divergence of the magnetic field is always zero everywhere, confirming that magnetic field lines form closed loops and that no point in space can be the source of the magnetic field.

📌 Key points (3–5)

  • What the differential form says: the divergence of the magnetic field B equals zero at every point in space.
  • How it connects to the integral form: applying the Divergence Theorem to the integral form (flux through a closed surface equals zero) yields the differential form.
  • Why the integrand must be zero: the integral over any volume equals zero only if the integrand itself is zero everywhere.
  • Common confusion: "source of the magnetic field" does not mean a localized starting point like charge for electric fields; magnetic fields have no localizable source.
  • Physical meaning: magnetic flux per unit volume is always zero, reinforcing that field lines never begin or end at any point.

🔗 From integral to differential form

🔗 Starting with the integral form

The integral form of Gauss' Law for magnetism (from Section 7.2) is:

The magnetic flux through a closed surface S equals zero.

In mathematical terms: the surface integral of B (magnetic flux density) over a closed surface S equals zero.

  • This means every field line entering the enclosed volume must also exit.
  • The integral form applies to any closed surface.

🧮 Applying the Divergence Theorem

The Divergence Theorem connects a volume integral to a surface integral:

The volume integral of (divergence of B) over volume V equals the surface integral of B over the closed surface S bounding V.

  • The right-hand side (surface integral) is zero from the integral form.
  • Therefore, the volume integral of (divergence of B) must also be zero.

🎯 Why the integrand must be zero

The volume integral equals zero regardless of the specific location or shape of V.

  • If the integral is zero for every possible volume, the only way this can be true is if the integrand itself is zero at every point.
  • Conclusion: divergence of B equals zero everywhere in space.

📐 The differential form equation

📐 Statement of the equation

Divergence of B equals zero.

This is the differential (or "point") form of Gauss' Law for Magnetic Fields.

  • "Differential" or "point" form means it applies at every individual point in space, not just over a region.
  • Contrast with the integral form, which applies to a surface or volume as a whole.

🔍 Physical interpretation

The flux per unit volume of the magnetic field is always zero.

  • "Flux per unit volume" is another way to describe divergence.
  • Zero divergence means no net flux is created or destroyed at any point.
  • Example: if you imagine a tiny volume around any point, the amount of magnetic field entering equals the amount leaving.

🚫 No localizable source

🚫 What "no source" means

The differential form implies there is no point in space that can be considered the source of the magnetic field.

  • If a point were a source, the total flux through a surface surrounding it would be greater than zero.
  • But the integral form guarantees flux through any closed surface is zero.
  • Therefore, magnetic field lines cannot begin or end at any point; they must form closed loops.

⚡ Contrast with electric fields

Field typeSourceField line behavior
ElectrostaticCharged particlesField lines begin at charges
MagneticNo localizable sourceField lines form closed loops
  • For electric fields, every field line begins at a charged particle.
  • For magnetic fields, there is no analogous "magnetic charge" or particle where field lines start.
  • Don't confuse: when we say current is the "source" of a magnetic field, we mean only that the field coexists with current, not that the field is attached to or originates from the current.

🌀 Magnetic fields are "weird"

The excerpt emphasizes that there is no localizable quantity analogous to charge for magnetic fields.

  • Charge is a localized quantity that sources electric fields.
  • Magnetic fields have no such localized source.
  • This is described as "another way in which magnetic fields are weird."

🛠️ Why the differential form matters

🛠️ Solving for magnetic fields

Given the differential equation (divergence of B equals zero) and boundary conditions imposed by structure and materials, we can solve for the magnetic field in very complicated scenarios.

  • The differential form is a local equation that applies at every point.
  • Combined with boundary conditions, it allows calculation of the field throughout a region.
  • This is the same reason the differential form of Gauss's Law for electrostatics is useful.
83

7.4 Ampere's Circuital Law (Magnetostatics): Integral Form

7.4 Ampere’s Circuital Law (Magnetostatics): Integral Form

🧭 Overview

🧠 One-sentence thesis

Ampere's circuital law in magnetostatics provides a way to calculate the magnetic field from a known current by relating the line integral of the magnetic field around a closed path to the total current enclosed by that path.

📌 Key points

  • What ACL states: the integral of the magnetic field intensity H around a closed path C equals the current enclosed by that path.
  • How to use it: ACL allows you to calculate the magnetic field when you know the source current, similar to how Gauss's Law works in electrostatics.
  • When it works best: symmetry is generally required to simplify the problem enough to solve the integral equation.
  • Common confusion: the direction of positive current must follow the right-hand rule—thumb points along the integration direction, fingers curl through the surface in the direction of positive current.
  • Scope limitation: this form applies only to magnetostatics (static magnetic fields), not time-varying situations.

🔄 The integral form equation

🔄 What the equation says

The integral of the magnetic field intensity H over a closed path C equals the current enclosed by that path, I_encl.

  • Written as: the line integral of H · dl around closed path C = I_encl.
  • Dimensional check: H has units of A/m; integrating over distance (m) gives units of current (A), which matches I_encl.
  • The enclosed current I_encl can be positive or negative depending on direction.

🖐️ Right-hand rule for direction

  • The direction of positive current flow must be correctly associated with the closed path C.
  • How to apply the rule:
    • Point your right thumb in the direction of integration around C.
    • Your fingers curl through any surface S bounded by C.
    • The direction your fingers point is the direction of positive I_encl.
  • This follows from Stokes' Theorem and is not a coincidence.

🎯 Choice of surface

  • S can be any surface bounded by C, not just the flat "taut" surface.
  • This flexibility is important for applying the law in different geometries.

🔧 Role and limitations

🔧 What ACL does for magnetostatics

  • ACL plays a role in magnetostatics similar to Gauss's Law in electrostatics.
  • It provides a means to calculate the magnetic field given the source current.
  • Example: if you know the current distribution and can choose a suitable path C, you can find H along that path.

⚠️ When ACL is practical

  • Symmetry requirement: generally, symmetry is needed to simplify the problem so the integral equation can be solved.
  • Fortunate cases: many important problems have the necessary symmetry, including:
    • Magnetic field of a line current.
    • Magnetic field inside a straight coil.
    • Magnetic field of a toroidal coil.
    • Magnetic field of a current sheet.
    • Boundary conditions on magnetic field intensity.

🚫 When ACL is not enough

  • If the necessary symmetry is not available, the integral form may not be solvable.
  • In such cases, the differential form of ACL may be required instead.
  • Don't confuse: the integral form is easier when symmetry exists; the differential form is more general but requires solving differential equations.

📐 Comparison with electrostatics

AspectGauss's Law (electrostatics)Ampere's Circuital Law (magnetostatics)
What it relatesElectric field to enclosed chargeMagnetic field to enclosed current
FormSurface integral of E over closed surface = charge/permittivityLine integral of H over closed path = enclosed current
When practicalRequires symmetryRequires symmetry
RoleCalculate E from known chargeCalculate H from known current
  • Both laws provide a way to find fields from sources.
  • Both have similar practical limitations (need symmetry).
  • Both have differential forms for more general cases.

⏱️ Scope reminder

⏱️ Magnetostatics only

  • The form of ACL presented here applies only to magnetostatics.
  • Magnetostatics means static (non-time-varying) magnetic fields.
  • Be aware that time-varying situations require a different form of the law.
  • Don't confuse: this is not the full, time-dependent version of Ampere's Law used in electrodynamics.
84

Magnetic Field of an Infinitely-Long Straight Current-Bearing Wire

7.5 Magnetic Field of an Infinitely-Long Straight Current-Bearing Wire

🧭 Overview

🧠 One-sentence thesis

Ampere's Circuital Law, combined with symmetry arguments, shows that an infinitely-long straight wire carrying steady current produces a magnetic field that circles the wire, increasing linearly inside the wire and decreasing inversely with distance outside.

📌 Key points (3–5)

  • Core method: Use Ampere's Circuital Law (ACL) with a circular integration path to exploit cylindrical symmetry and solve for the magnetic field.
  • Inside vs outside behavior: Inside the wire (radius less than wire radius a), field strength grows proportionally to distance from center; outside, it falls off inversely with distance.
  • Direction rule: The magnetic field forms concentric circles perpendicular to the wire; use the right-hand rule (thumb along current, fingers curl in field direction).
  • Common confusion: The enclosed current differs inside vs outside the wire—inside, only the fraction of total current within the integration path counts.
  • Key requirement: This solution requires the symmetry of an infinite straight wire; without such symmetry, the differential form of ACL is needed.

🔧 Setting up Ampere's Circuital Law

🔧 The integral form for this problem

The relevant form of ACL is:

The line integral of magnetic field H around a closed path C equals the enclosed current I_encl.

Written in words: integral of H · dl around C = I_encl

  • Units check: H has units of amperes per meter (A/m); integrating over distance (m) yields amperes (A), matching current units.
  • The closed path C can be any shape, but choosing one that matches the problem's symmetry simplifies the math dramatically.

🎯 Choosing the integration path

  • The wire is a circular cylinder of radius a aligned along the z-axis.
  • Smart choice: Pick a circular path of radius ρ in the z = 0 plane, centered at the origin.
  • This exploits cylindrical symmetry—the field should "look the same" at every point on this circle.

➕ Determining enclosed current

Two cases based on the integration path radius ρ:

RegionConditionEnclosed currentReason
Outside wireρ ≥ aI_encl = IPath encloses entire wire cross-section
Inside wireρ < aI_encl = I × (ρ²/a²)Only fraction of wire area is enclosed; DC current distributes uniformly
  • Why the fraction: Steady (DC) current distributes uniformly over the wire's cross-section.
  • Area enclosed by path = π ρ², total wire area = π a², so enclosed current scales as the area ratio.

🧮 Solving using symmetry arguments

🧮 Three symmetry observations

The problem's infinite, uniform cylindrical geometry constrains the field:

  1. No z-dependence: Current distribution is uniform and infinite in z-direction → H cannot depend on z → H · ẑ must be zero everywhere.

  2. Rotational symmetry: Problem looks identical after any rotation in φ → magnitude of H cannot depend on φ → magnitude depends only on ρ.

  3. No radial component: Radial symmetry requires H · ρ̂ = 0 (otherwise field wouldn't be radially symmetric) → combined with point 1, H must be entirely in the ±φ̂ direction.

Conclusion: The most general form is H = φ̂ H(ρ), where H(ρ) is some function of radius only.

🧮 Evaluating the integral

Substituting H = φ̂ H(ρ) into the ACL integral:

  • The path element in cylindrical coordinates: dl = φ̂ ρ dφ
  • Integrating from φ = 0 to 2π
  • Result: I_encl = 2π ρ H(ρ)
  • Solving: H(ρ) = I_encl / (2π ρ)

📐 Final magnetic field solutions

📐 Outside the wire (ρ ≥ a)

H = φ̂ I / (2π ρ)

  • Field strength is inversely proportional to distance from the wire center.
  • As you move farther away, the field decreases.
  • The entire current I is enclosed.

📐 Inside the wire (ρ < a)

H = φ̂ I ρ / (2π a²)

  • Field strength is directly proportional to distance from the wire center.
  • As you move from center toward the wire surface, the field increases linearly.
  • Only the fraction of current within radius ρ contributes.

📐 Behavior at the boundary

  • At ρ = a (the wire surface), both formulas give the same value: I / (2π a).
  • Inside: field grows from zero at center to maximum at surface.
  • Outside: field decreases continuously from the surface value.

Don't confuse: The field doesn't simply decrease everywhere—it increases inside the wire, reaches maximum at the surface, then decreases outside.

🧭 Direction and the right-hand rule

🧭 Field line geometry

The magnetic field lines form concentric circles perpendicular to and centered on the wire.

  • Field is φ̂-directed (tangent to circles around the wire).
  • For current flowing in +z direction, field points in +φ̂ direction.
  • Example: If current flows upward, field circles counterclockwise when viewed from above.

🧭 Right-hand rule for quick determination

The magnetic field points in the direction of the curled fingers of the right hand when the thumb is aligned in the direction of current flow.

  • How to use: Point your right thumb along the current direction; your fingers naturally curl in the direction the magnetic field circles the wire.
  • This rule applies to many other magnetic field problems, not just infinite wires.
  • It emerges naturally from the mathematics but provides a quick visual check.

🧭 Connection to integration direction

  • The integration direction around path C and the positive current direction follow the right-hand rule from Stokes' Theorem.
  • When thumb points in integration direction, fingers intersect the surface in the positive current direction.
  • Changing the integration direction should not change the final magnetic field result (the signs adjust consistently).

🔗 Context and limitations

🔗 Role of ACL in magnetostatics

Ampere's Circuital Law plays a role similar to Gauss' Law in electrostatics:

  • Provides a means to calculate magnetic field from source current.
  • Key limitation: Generally requires symmetry to simplify the integral equation enough to solve it.
  • Fortunately, many important problems (line currents, coils, current sheets) have sufficient symmetry.

🔗 When this approach doesn't work

  • If the necessary symmetry is not available, the differential form of ACL may be required instead.
  • This magnetostatic form applies only to steady currents; time-varying electric fields require an additional "displacement current" term.

🔗 Converting to magnetic flux density

If desired, the magnetic flux density B can be obtained from: B = μ H, where μ is the permeability of the material surrounding the wire.

85

Magnetic Field Inside a Straight Coil

7.6 Magnetic Field Inside a Straight Coil

🧭 Overview

🧠 One-sentence thesis

The magnetic field deep inside an ideal straight coil is uniform, aligned with the coil axis, and proportional to both the winding density (turns per unit length) and the current flowing through the wire.

📌 Key points (3–5)

  • Core result: The magnetic field inside a straight coil is given by H ≈ ẑ (N I / l), where N is the number of turns, I is current, and l is coil length.
  • Uniformity: The field magnitude is constant throughout the interior—the same along the axis and near the cylinder wall—but this breaks down near the coil ends due to fringing fields.
  • Direction: The field points along the coil axis (ẑ direction) and follows the right-hand rule: thumb along current flow, fingers curl in the field direction.
  • Common confusion: The analysis applies only deep inside the coil where l ≫ a (length much greater than radius); near the ends, field lines diverge and the simple formula fails.
  • Design levers: To increase the magnetic field, either pack more turns per unit length (increase N/l) or increase the current I.

🧲 Geometry and setup

🧲 Physical structure

  • The coil is circular with radius a and length l.
  • It consists of N turns (windings) of wire wound with uniform winding density.
  • The problem uses cylindrical coordinates with the coil axis aligned along the z-axis.

🧲 Key assumptions

  • The current is steady (DC).
  • The windings are sufficiently closely-spaced that magnetic field lines cannot take "shortcuts" between individual windings—they can only pass through the openings at the coil ends.
  • The analysis restricts attention to the field deep inside the coil, away from the ends, and assumes l ≫ a (coil length much greater than radius).

🧭 Direction of the magnetic field

🧭 Right-hand rule application

  • The excerpt applies the right-hand rule from the previous section (straight wire): when the thumb points in the direction of current flow, the fingers curl in the direction of the magnetic field.
  • For a coil, consider one short segment of one turn: the field contribution from that segment follows the right-hand rule.
  • Summing over all segments in many loops shows the field inside the coil is generally in the +ẑ direction when current I flows as shown.

🧭 Field line structure

Magnetic field lines inside a straight coil with closely-spaced windings form closed loops aligned with the coil axis.

  • Deep inside the coil, field lines run parallel to the axis.
  • Near the ends, field lines diverge from the axis because they must form closed loops—this creates a "fringing field" that is complex and difficult to analyze.
  • The excerpt emphasizes that the simple analysis applies only deep inside, where the field is approximately ẑ H(ρ).

🧭 Radial symmetry

  • The coil's radial symmetry means the field magnitude cannot depend on the azimuthal angle φ.
  • The most general form for the field deep inside is H ≈ ẑ H(ρ), where the direction is ±ẑ and magnitude depends at most on the radial distance ρ.
  • The analysis will show that H does not even depend on ρ—it is uniform.

🔁 Applying Ampere's Circuital Law

🔁 The integral form of ACL

∮_C H · dl = I_encl

  • C is any closed path enclosing the current of interest.
  • I_encl is the total current enclosed by the path.
  • The choice of path is arbitrary, but a path on a constant-coordinate surface simplifies the calculation.

🔁 Chosen integration path

  • The path is a rectangle in the ρ-z plane with four segments: A, B, C, and D.
  • Segment A runs along the z-axis from z₁ to z₂ at radius ρ₁ (inside the coil).
  • Segment C runs parallel to A at radius ρ₂ (outside the coil, taken to infinity).
  • Segments B and D run radially, connecting A and C.
  • The path length in the z direction is l′ = z₂ − z₁.

🔁 Enclosed current calculation

  • Winding density: N / l turns per meter.
  • Number of turns enclosed by path of length l′: (N / l) l′.
  • Total enclosed current: I_encl = (N / l) l′ I.

🧮 Evaluating the integral

🧮 Segment-by-segment analysis

The integral ∮_C H · dl splits into four segments:

SegmentPath elementH · dlContribution
Adl = ẑ dz at ρ₁H(ρ₁) dzH(ρ₁) l′
Bdl = ρ̂ dρ0 (perpendicular)0
Cdl = ẑ dz at ρ₂ → ∞H(ρ₂) dz → 00
Ddl = ρ̂ dρ0 (perpendicular)0

🧮 Why segments B and D vanish

  • For these segments, dl = ρ̂ dρ (radial direction).
  • The field is H = ẑ H(ρ) (z-direction).
  • Therefore H · dl = 0 because ẑ · ρ̂ = 0 (perpendicular directions).

🧮 Why segment C vanishes

  • Let ρ₂ → ∞ (take the path far outside the coil).
  • The field magnitude H(ρ) outside the coil must decrease with distance from the coil.
  • For sufficiently large ρ, H(ρ) becomes negligible, so the integral over segment C becomes negligible.

🧮 Only segment A contributes

  • The integral simplifies to: (N / l) l′ I = ∫(z₁ to z₂) [ẑ H(ρ₁)] · (ẑ dz) = H(ρ₁) l′.
  • Dividing both sides by l′: H(ρ₁) = (N I) / l.
  • The result is independent of ρ₁—the field does not depend on radial position inside the coil.

📐 Final result and implications

📐 The magnetic field formula

H ≈ ẑ (N I / l) inside the coil

  • Direction: aligned with the coil axis (ẑ).
  • Magnitude: winding density (N / l) times current I.
  • Units: current divided by length gives A/m, the correct units for H.

📐 Uniformity of the field

  • The field is uniform inside the coil: the same magnitude along the axis and close to the cylinder wall.
  • This does not apply close to the ends, where the fringing field dominates.
  • Example: A point on the axis and a point near the coil wall at the same z-coordinate experience the same field strength.

📐 Design implications

To increase the magnetic field inside a coil:

  • Increase winding density (N / l): pack more turns into the same length, or use a shorter coil with the same number of turns.
  • Increase current I: drive more current through the wire.

Don't confuse:

  • The formula H = (N I) / l applies only deep inside the coil, far from the ends.
  • Near the ends, field lines diverge and the field is not uniform—the simple formula breaks down.
  • The assumption l ≫ a (length much greater than radius) is essential for the analysis to hold.

📐 Applications

  • Inductors: coils used to store magnetic energy in circuits.
  • Solenoids: coils used as magnets, typically as part of an actuator (a device that converts electrical energy into mechanical motion).
  • Building block: the result provides insight for analyzing more complex magnetic field problems.
86

Magnetic Field of a Toroidal Coil

7.7 Magnetic Field of a Toroidal Coil

🧭 Overview

🧠 One-sentence thesis

A toroidal coil confines its magnetic field almost entirely inside the coil, making the external field negligibly small and proportional only to distance from the central axis, winding density, and current.

📌 Key points (3–5)

  • Principal advantage: magnetic field containment—the field outside a toroidal coil can be made negligibly small, reducing interactions with nearby fields and structures.
  • Inside the coil: the magnetic field depends only on distance from the central axis (ρ) and is proportional to winding density and current; it is independent of vertical position (z).
  • Outside the coil: the magnetic field is zero everywhere because any closed path outside the coil encloses no current.
  • Common confusion: the field inside is not uniform like in a straight coil—it varies with distance from the central axis (ρ), even though it does not vary with z.
  • How to determine direction: use the right-hand rule—curl fingers in the direction of current flow, and the thumb points in the magnetic field direction.

🔧 Geometry and setup

🔧 What a toroid is

A toroid is a cylinder in which the ends are joined to form a closed loop.

  • Think of it as a doughnut-shaped coil.
  • The excerpt describes a circular toroid with inner radius a and outer radius b.
  • The coil consists of N windings (turns) of wire wound with uniform winding density.
  • Toroidal coils are commonly used to form inductors and transformers.

🧭 Coordinate system and variables

  • The problem is easiest to work in cylindrical coordinates.
  • The toroid is centered on the origin in the z = 0 plane.
  • Key variables:
    • ρ: distance from the central (z) axis
    • φ: angular position around the z axis
    • z: vertical position
    • I: steady (DC) current flowing through the coil
    • N: total number of windings

🧲 Magnetic field inside the coil

🧲 Direction of the field

  • The magnetic field inside the coil is aligned with the φ direction (tangent to circles centered on the z axis).
  • This follows from the right-hand rule applied to each short segment of wire:
    • Curl the fingers of your right hand in the direction of current flow.
    • Your thumb points in the direction of the magnetic field.
  • When current I flows as shown, the field points in the +φ direction.
  • Because the problem is rotationally symmetric around the z axis, the magnitude cannot depend on φ.

📐 Deriving the field using Ampere's Circuital Law

The excerpt uses the magnetostatic form of Ampere's Circuital Law (ACL):

The integral of H · dl around a closed path C equals the enclosed current I_encl.

Steps:

  1. Choose a circular path C of radius ρ centered on the origin in the z = z₀ plane.
  2. The path lies entirely inside the coil, so it encloses the current from all N windings.
  3. Enclosed current: I_encl = N × I.
  4. Evaluate the integral around the circle:
    • The field H = φ̂ H(ρ, z₀) is tangent to the path.
    • The path element dl = φ̂ ρ dφ.
    • Integral: N × I = 2π ρ H(ρ, z₀).
  5. Solve for H: H = φ̂ (N × I) / (2π ρ).

📊 Key properties of the internal field

PropertyWhat it means
Depends only on ρMagnitude varies with distance from the central axis, but is independent of z (vertical position) or φ (angular position).
Proportional to winding densityMore windings per unit circumference → stronger field.
Proportional to currentHigher current I → stronger field.
Dimensionally correctCurrent divided by circumference (2π ρ) gives units of A/m, the units of H.

Summary from the excerpt:

The magnetic field inside a toroidal coil depends only on distance from the central axis and is proportional to winding density and current.

🔄 How to increase the field

  • Use more windings (N).
  • Increase the current (I).
  • Example: doubling the number of windings or doubling the current will double the magnetic field strength.

⚠️ Don't confuse with a straight coil

  • In a straight coil (Section 7.6), the field deep inside is uniform and does not vary with position.
  • In a toroidal coil, the field inside varies with ρ (distance from the central axis), even though it is independent of z.
  • The toroidal geometry introduces this ρ-dependence because the circumference changes with ρ.

🚫 Magnetic field outside the coil

🚫 Why the external field is zero

  • Consider any closed path C that lies completely outside the coil.
  • Such a path encloses no current, so I_encl = 0.
  • Ampere's Circuital Law gives: integral of H · dl around C = 0.

Two possibilities:

  1. H is zero everywhere along the path.
  2. H is non-zero but arranged so the integral cancels out.

The excerpt rules out the second possibility:

  • The problem has radial symmetry (looks the same after any rotation around the z axis).
  • If both H and C are radially symmetric, the sign of H · dl should not change over C.
  • Therefore, the only way the integral can be zero is if H itself is zero.

🛡️ Magnetic field containment

  • The principal advantage of toroidal coils over straight coils is magnetic field containment.
  • The magnetic field outside a toroidal coil can be made negligibly small.
  • This reduces concern about interactions between this field and other fields and structures in the vicinity.
  • Example: in a power supply (as shown in the figure), a toroidal inductor prevents its magnetic field from interfering with nearby electronic components.

🔗 Connection to straight coils

🔗 Similarities and differences

The excerpt recommends reviewing Section 7.6 (Magnetic Field Inside a Straight Coil) because:

  • Both use the right-hand rule to determine field direction.
  • Both show that the field is aligned with the axis of the coil (or, in the toroid, tangent to circles around the central axis).
  • Both fields are proportional to winding density and current.

Key difference:

  • Straight coil: field deep inside is uniform (does not vary with position).
  • Toroidal coil: field inside varies with ρ (distance from the central axis).

🧩 Why the toroidal field varies with ρ

  • Each short segment of wire contributes to the field according to the right-hand rule.
  • When you sum contributions from all segments, the result depends on the circumference at radius ρ.
  • The formula H = φ̂ (N × I) / (2π ρ) shows that as ρ increases, the field strength decreases (because the same current is "spread out" over a larger circumference).
87

Magnetic Field of an Infinite Current Sheet

7.8 Magnetic Field of an Infinite Current Sheet

🧭 Overview

🧠 One-sentence thesis

The magnetic field produced by an infinite sheet of uniform current is spatially uniform on each side of the sheet, with equal magnitude but opposite direction above and below the sheet.

📌 Key points (3–5)

  • What the problem models: an infinite sheet of current lying in a plane, with uniform current density flowing in one direction; useful for approximating finite current sheets in practical problems like microstrip transmission lines and ground plane currents.
  • Key result: the magnetic field magnitude is constant everywhere on each side of the sheet, equal to half the surface current density, and flips sign across the sheet.
  • How to solve it: use Ampere's Circuital Law with a rectangular path that crosses the sheet, exploiting symmetry to eliminate unknowns.
  • Common confusion: the field does not depend on distance from the sheet—it is uniform throughout each half-space, unlike fields from line currents or coils.
  • Why symmetry matters: symmetry arguments eliminate field components and coordinate dependencies before any calculation, simplifying the problem dramatically.

🧩 Problem setup and symmetry

🧩 Geometry and current density

  • The current sheet lies in the plane z = 0.
  • Current density is given by J_s = x-hat J_s (units of A/m).
    • This means current flows uniformly in the x direction.
    • The total current crossing any segment of width Δy along the y direction is J_s Δy.
  • The sheet is infinite in extent, so edge effects are absent.

🔍 What symmetry tells us before calculation

The excerpt emphasizes that "we already know quite a bit" from symmetry alone:

  • No y-component: Imagine the sheet as many thin strips parallel to the x axis. Each strip behaves like a straight line current I = J_s Δy. The right-hand rule shows each strip's field has no y-component, so the total field cannot have a y-component either.
  • No z-component: When fields from all strips are summed, the z-components cancel due to symmetry.
  • No x or y dependence: By symmetry, the magnitude of H cannot depend on x or y coordinates.
  • Sign flip: The sign of H(z) must be positive for z < 0 and negative for z > 0 (or vice versa, depending on current direction).

Conclusion from symmetry: The most general form for H is y-hat H(z), with opposite signs above and below the sheet.

🛠️ Alternative approach mentioned

  • It is possible to solve by summing over a continuum of thin current strips (each treated as a line current).
  • The excerpt notes this is "far easier to use Ampere's Circuital Law" instead.
  • The footnote suggests trying the summation approach "if for no other reason than to see how much simpler it is to use ACL instead."

🔁 Applying Ampere's Circuital Law

🔁 Choosing the integration path

Ampere's Circuital Law (ACL): the integral of H · dl around a closed path C equals the enclosed current I_encl.

  • The path must enclose some current to relate J_s and H.
  • A convenient choice: a rectangle lying in the x = 0 plane, centered on the origin (see Figure 7.13 in the excerpt).
  • The rectangle has width L_y in the y dimension and width L_z in the z dimension.
  • Integration direction: counter-clockwise from the perspective shown, consistent with the right-hand rule from Stokes' Theorem for positive J_s.

🧮 Evaluating the enclosed current

  • The enclosed current is simply the surface current density times the width of the path in the y direction:
    • I_encl = J_s L_y
  • This is the current passing through the surface bounded by the rectangular path.

🧮 Evaluating the line integral

Starting from ACL:

  • Integral of [y-hat H(z)] · dl around C = J_s L_y

Key observation: On the vertical sides of the rectangle, H · dl = 0 because:

  • H is y-directed.
  • dl = z-hat dz on the vertical sides.
  • The dot product of perpendicular vectors is zero.

Only horizontal sides contribute:

  • Bottom side (at z = -L_z/2): integral from -L_y/2 to +L_y/2 of [y-hat H(-L_z/2)] · (y-hat dy)
  • Top side (at z = +L_z/2): integral from -L_y/2 to +L_y/2 of [y-hat H(+L_z/2)] · (y-hat dy)

Evaluating these integrals:

  • H(-L_z/2) L_y - H(+L_z/2) L_y = J_s L_y
  • All factors of L_y cancel.

🔑 Using symmetry to solve

From symmetry:

  • H(-L_z/2) = -H(+L_z/2) because:
    1. Symmetry between upper and lower half-spaces.
    2. The sign change between these half-spaces noted earlier.

Substituting this relation:

  • 2 H(-L_z/2) = J_s
  • Therefore H(-L_z/2) = +J_s/2
  • And H(+L_z/2) = -J_s/2

Critical insight: H is independent of L_z. The same value of H is found regardless of the value of L_z chosen for the integration path.

📊 Final result and interpretation

📊 The magnetic field expression

H = ±y-hat J_s/2 for z ≶ 0

In words:

  • The magnetic field intensity is uniform throughout all space, except for a sign change across the sheet.
  • Magnitude is half the surface current density.
  • Direction is +y-hat below the sheet (z < 0) and -y-hat above the sheet (z > 0).

🎯 Key characteristics

PropertyValueWhy
MagnitudeJ_s/2 everywhereIndependent of distance from sheet
Direction below sheet+y-hatRight-hand rule from current direction
Direction above sheet-y-hatOpposite to below
Dependence on x, yNoneSymmetry of infinite sheet
Dependence on zSign onlyUniform within each half-space

🔬 Physical interpretation

  • Uniform field: Unlike a line current (field decreases with distance) or a toroidal coil (field varies with radius), the infinite current sheet produces a field of constant magnitude on each side.
  • Discontinuity at the sheet: The field flips direction across z = 0, but the magnitude remains J_s/2 on both sides.
  • Don't confuse: This is not a field that "falls off" with distance—it is truly uniform in each half-space. The infinite extent of the sheet is responsible for this behavior.

🛠️ Practical applications

The excerpt mentions this solution is useful for:

  • Building block: insight and starting point for more complex problems.
  • Approximation: finite current sheets in practical scenarios, including:
    • Microstrip transmission line currents.
    • Ground plane currents in printed circuit boards.

Example: A large ground plane carrying return current can be approximated as an infinite sheet when analyzing fields far from the edges but close to the plane.

88

Ampere's Law (Magnetostatics): Differential Form

7.9 Ampere’s Law (Magnetostatics): Differential Form

🧭 Overview

🧠 One-sentence thesis

The differential form of Ampere's Circuital Law reveals that volume current density at any point equals the curl of the magnetic field intensity at that point, showing that current is proportional to the spatial rate of change of the magnetic field and perpendicular to it.

📌 Key points (3–5)

  • From integral to differential: the differential form is derived from the integral form of Ampere's Circuital Law using Stokes' Theorem.
  • What the differential form says: curl of H equals J (volume current density) at every point in space.
  • Physical insight: volume current density is proportional to the spatial rate of change of the magnetic field and is perpendicular to the field at that point.
  • Common confusion: the integral form relates field along a closed path to total enclosed current; the differential form relates field behavior at a single point to current density at that point.
  • Why it matters: combined with boundary conditions, the differential equation can solve for magnetic fields in arbitrarily complicated scenarios.

🔄 From integral to differential form

🔄 Starting point: the integral form

Integral form of Ampere's Circuital Law (ACL): the line integral of magnetic field intensity H around a closed path C equals the total current enclosed by that path.

  • Mathematical statement: the integral of H · dl around closed curve C equals I_encl.
  • H is magnetic field intensity.
  • C is the closed curve (path).
  • I_encl is the total current flowing through any surface bounded by C.

🧮 The derivation strategy

The differential form is obtained by:

  1. Applying Stokes' Theorem to convert the line integral into a surface integral.
  2. Expressing enclosed current as an integral of volume current density J.
  3. Equating the two surface integrals.
  4. Concluding that the integrands must be equal.

🔧 Applying Stokes' Theorem

  • Stokes' Theorem states: the surface integral of (curl of H) · ds equals the line integral of H · dl around C.
  • S is any surface bounded by C.
  • ds is the differential surface area combined with the unit vector in the direction determined by the right-hand rule.
  • The right side of this equation is I_encl (from ACL).

📐 Expressing enclosed current

  • I_encl can be written as the surface integral of volume current density J · ds.
  • J has units of A/m² (amperes per square meter).
  • This converts the total current into a density-based integral over the same surface S.

⚖️ Equating integrands

  • After substitution: surface integral of (curl of H) · ds equals surface integral of J · ds.
  • This relationship must hold regardless of the specific location or shape of S.
  • The only way this is possible for all surfaces in all scenarios is if the integrands themselves are equal.
  • Therefore: curl of H equals J.

🧲 The differential form and its meaning

🧲 The equation

Differential form of Ampere's Circuital Law for magnetostatics: curl of H equals J.

  • At any point in space, the curl of the magnetic field intensity equals the volume current density at that point.
  • This is a local relationship (point-by-point), unlike the integral form which is global (over a path and surface).

🔍 Physical interpretation

The differential form tells us:

  • Volume current density at any point is proportional to the spatial rate of change of the magnetic field.
  • Volume current density is perpendicular to the magnetic field at that point.

Why these conclusions?

  • The curl operator involves derivatives with respect to direction (spatial rate of change).
  • The curl of a vector field measures rotation and is perpendicular to the field itself.

🎯 Practical use

  • Combined with boundary conditions (associated with discontinuities in structure and materials), the differential equation can solve for magnetic fields in arbitrarily complicated scenarios.
  • It provides deeper insight into the relationship between current and magnetic field beyond what the integral form reveals.

🔀 Integral vs. differential perspectives

AspectIntegral formDifferential form
What it relatesField along a closed path to total enclosed currentField behavior at a point to current density at that point
ScopeGlobal (entire path and surface)Local (single point)
Mathematical formLine integral of H around C equals I_enclCurl of H equals J
Use caseCalculating field from known current distribution with symmetrySolving field in complex scenarios with boundary conditions

⚠️ Don't confuse

  • The integral form is not "less accurate" than the differential form; they are equivalent but suited to different problems.
  • The differential form does not replace the integral form; it complements it by providing point-wise detail.
  • Example: for symmetric current distributions (infinite wire, infinite sheet), the integral form is often simpler; for complex geometries with material boundaries, the differential form with boundary conditions may be necessary.
89

Boundary Conditions on the Magnetic Flux Density (B)

7.10 Boundary Conditions on the Magnetic Flux Density ( B )

🧭 Overview

🧠 One-sentence thesis

At the interface between two different materials, the normal component of the magnetic flux density B remains continuous, whereas the normal component of the magnetic field intensity H becomes discontinuous when the materials have different permeabilities.

📌 Key points (3–5)

  • What boundary conditions describe: continuities and discontinuities in electromagnetic fields at interfaces between dissimilar materials.
  • The key result for B: the normal (perpendicular) component of B is continuous across any material boundary.
  • The consequence for H: because B = μH, the normal component of H is discontinuous when permeabilities differ between the two regions.
  • Common confusion: B and H behave differently at boundaries—B's normal component stays continuous, but H's normal component jumps if permeabilities are unequal.
  • Why it matters: boundary conditions constrain and help solve for fields away from interfaces.

🔬 Why boundary conditions exist

🔬 Behavior in homogeneous vs. dissimilar media

  • In homogeneous media, electromagnetic quantities vary smoothly and continuously.
  • At an interface between dissimilar media, electromagnetic quantities can be discontinuous.
  • Boundary conditions mathematically describe these continuities and discontinuities.

🎯 Purpose of boundary conditions

Boundary conditions: mathematical descriptions of continuities and discontinuities in fields at interfaces, used to constrain solutions for fields away from these interfaces.

  • They allow us to connect field solutions on either side of a material boundary.
  • Example: if you know the field in region 1 and the boundary condition, you can determine the field in region 2.

📐 Deriving the boundary condition for B

📐 Starting point: Gauss' Law for Magnetic Fields

The derivation begins with Gauss' Law for Magnetic Fields (GLM):

The integral of B · ds over any closed surface S equals zero.

  • This law states that the net magnetic flux through any closed surface is zero.
  • The excerpt uses a cylindrical surface centered at a point on the interface.

🛢️ The cylindrical test surface

The derivation uses a specific geometry:

  • A cylinder centered at the interface.
  • Flat ends parallel to the surface and perpendicular to the unit normal vector n̂.
  • Radius a and total length 2h.
  • The cylinder straddles the boundary between region 1 and region 2.

🔍 The shrinking limit

The key mathematical step involves taking a limit:

  • Reduce h and a together while maintaining h/a ≪ 1 (h much smaller than a).
  • Keep the cylinder centered on the interface.
  • Because h ≪ a, the side area becomes negligible compared to the top and bottom areas.
  • As h → 0, only the top and bottom contributions remain.

⚖️ The final equation

After the limit, the integral simplifies:

  • B₁ · n̂ ΔA + B₂ · (−n̂) ΔA → 0
  • B₁ and B₂ are the magnetic flux densities at the interface in regions 1 and 2.
  • ΔA is the area of the top and bottom sides.
  • The orientation of n̂ is important: it points into region 1.

This yields:

n̂ · (B₁ − B₂) = 0

where n̂ points into region 1.

✅ The boundary condition for B

✅ What it means

The normal (perpendicular) component of B across the boundary between two material regions is continuous.

  • "Continuous" means the component of B perpendicular to the interface does not jump.
  • The value of B's normal component just inside region 1 equals the value just inside region 2.
  • Example: if the normal component of B is 5 units on one side, it is 5 units on the other side.

🔄 Consequence for H

Because B = μH (where μ is permeability):

The normal (perpendicular) component of H across the boundary between two material regions is discontinuous if the permeabilities are unequal.

  • If μ₁ ≠ μ₂, then even though B's normal component is continuous, H's normal component must jump.
  • Don't confuse: B and H are related but behave differently at boundaries.
FieldNormal component behaviorCondition
BContinuousAlways
HDiscontinuousWhen μ₁ ≠ μ₂
HContinuousOnly when μ₁ = μ₂

🧲 Physical interpretation

  • The continuity of B's normal component reflects the fact that magnetic field lines do not begin or end (no magnetic monopoles).
  • The discontinuity in H's normal component arises from the different magnetic properties (permeabilities) of the two materials.
  • Example: a magnetic field line crossing from air into iron will have the same normal B component on both sides, but the normal H component will differ because iron has a much higher permeability than air.
90

Boundary Conditions on the Magnetic Field Intensity (H)

7.11 Boundary Conditions on the Magnetic Field Intensity ( H )

🧭 Overview

🧠 One-sentence thesis

At the boundary between two magnetic materials, the tangential component of H is discontinuous when surface current is present, while the normal component of B remains continuous across the interface.

📌 Key points (3–5)

  • Normal component of B: continuous across any material boundary (no discontinuity).
  • Tangential component of H: discontinuous when surface current flows at the boundary; continuous when no surface current exists.
  • Common confusion: B and H behave oppositely at boundaries—B's normal component is continuous while H's normal component is discontinuous if permeabilities differ; H's tangential component is discontinuous (with current) while B's tangential component is discontinuous (with unequal permeabilities and no current).
  • Mathematical form: the boundary condition is expressed as n̂ × (H₁ − H₂) = Jₛ, where n̂ points into region 1 and Jₛ is surface current density.
  • Physical meaning: any discontinuity in the tangential H field must be supported by surface current flowing perpendicular to that field component.

🧲 Continuity of the normal component of B

🧲 Derivation from Gauss's Law for Magnetism

The excerpt applies Gauss's Law for Magnetism (GLM) to a cylindrical surface centered on the interface between two materials:

  • The integral of B over a closed surface equals zero.
  • As the cylinder height h shrinks to zero while maintaining h/a ≪ 1, the side area becomes negligible.
  • Only the top and bottom surfaces contribute: B₁ · n̂ ΔA + B₂ · (−n̂) ΔA → 0.

📐 The boundary condition for B

The normal (perpendicular) component of B across the boundary between two material regions is continuous.

  • Mathematically: n̂ · (B₁ − B₂) = 0, where n̂ points into region 1.
  • This means the component of B perpendicular to the interface does not jump.
  • Example: if B₁ has a normal component of 5 units pointing into region 1, then B₂ must also have a normal component of 5 units in the same direction.

⚠️ Implication for H's normal component

Because B = μH, and the permeabilities μ may differ between regions:

The normal (perpendicular) component of H across the boundary between two material regions is discontinuous if the permeabilities are unequal.

  • Don't confuse: B's normal component is continuous, but H's normal component is not (unless μ₁ = μ₂).
  • The ratio H₁ₙ/H₂ₙ = μ₂/μ₁ to keep B₁ₙ = B₂ₙ.

🔄 Discontinuity of the tangential component of H

🔄 Derivation from Ampere's Circuital Law

The excerpt applies Ampere's Circuital Law (ACL) to a rectangular loop centered on the boundary:

  • The loop has sides A, B, C, D; sides A and C are parallel to the boundary (length l), sides B and D are perpendicular (length w).
  • As w and l shrink while keeping the loop centered, the contributions from B and D cancel out.
  • Only sides A and C remain: H₁ · (−t̂ Δl) + H₂ · (+t̂ Δl) = Iₑₙcₗ.

📏 Relating enclosed current to surface current density

  • The enclosed current Iₑₙcₗ is the flux of surface current density Jₛ (units A/m) flowing past the line of length Δl.
  • Mathematically: Iₑₙcₗ → Jₛ · (Δl t̂ × n̂), where t̂ is tangent to the boundary and n̂ is normal pointing into region 1.
  • This holds regardless of the particular direction chosen for t̂, as long as it is tangent to the boundary.

🧮 The boundary condition for H (tangential component)

After algebraic manipulation and applying vector identities, the excerpt derives:

n̂ × (H₁ − H₂) = Jₛ

  • This is the most commonly expressed form.
  • It means: the cross product of the normal vector with the difference in H fields equals the surface current density.
  • Interpretation: a discontinuity in the tangential component of H at the boundary must be supported by surface current flowing in a direction perpendicular to this component of the field.

🔌 Special case: no surface current

If there is no surface current, then the tangential component of the magnetic field intensity is continuous across the boundary.

  • When Jₛ = 0, then n̂ × (H₁ − H₂) = 0, so the tangential components of H₁ and H₂ are equal.
  • Example: in a region with no free currents at the interface, H's tangential component does not jump.

⚠️ Implication for B's tangential component

Because B = μH, and the permeabilities may differ:

In the absence of surface current, the tangential component of B across the boundary between two material regions is discontinuous if the permeabilities are unequal.

  • Don't confuse: when there is no surface current, H's tangential component is continuous, but B's tangential component is not (unless μ₁ = μ₂).
  • The ratio B₁ₜ/B₂ₜ = μ₁/μ₂ to keep H₁ₜ = H₂ₜ.

📊 Summary of boundary behavior

📊 Comparison table

Field componentBehavior at boundaryCondition
B normalContinuousAlways (from GLM)
H normalDiscontinuousIf μ₁ ≠ μ₂
H tangentialDiscontinuousIf surface current Jₛ ≠ 0
H tangentialContinuousIf Jₛ = 0
B tangentialDiscontinuousIf Jₛ = 0 and μ₁ ≠ μ₂

🧩 Why boundary conditions matter

  • In homogeneous media, electromagnetic quantities vary smoothly and continuously.
  • At a boundary between dissimilar media, electromagnetic quantities can be discontinuous.
  • Boundary conditions describe these continuities and discontinuities mathematically.
  • They are used to constrain solutions for fields away from the boundaries.

🔍 Key distinctions to remember

  • B vs H at boundaries: they behave oppositely for normal and tangential components.
  • With vs without surface current: the presence of Jₛ determines whether H's tangential component jumps.
  • Equal vs unequal permeabilities: when μ₁ = μ₂, both B and H have the same boundary behavior; when μ₁ ≠ μ₂, their behaviors diverge.
91

Inductance

7.12 Inductance

🧭 Overview

🧠 One-sentence thesis

Inductance is the ability of a current-carrying structure to store energy in a magnetic field, determined by the structure's geometry and the permeability of the surrounding material, not by the current itself.

📌 Key points (3–5)

  • What inductance measures: the relationship between current applied to a structure and the energy stored in the associated magnetic field.
  • How inductance is defined: the ratio of magnetic flux (Φ) to current (I), with a factor N for multiple windings that link the same flux.
  • What determines inductance: geometry of the current-bearing structure and permeability of the intervening medium—current is a stimulus or response, not a determinant.
  • Common confusion: positive reactance does not always mean physical inductance; "pin inductance" or "lead inductance" refers to equivalent circuit behavior (reactance), not actual energy storage in a magnetic field.
  • Connection to circuit theory: the voltage across an inductor equals L times the time-derivative of current (V = L dI/dt).

🔋 Energy storage mechanism

🔋 How coils store energy

  • Current in a coil creates a magnetic field.
  • The magnetic field exerts force on other current-bearing structures (e.g., other windings in the same coil).
  • If the windings are fixed in place, the force cannot do mechanical work, so the coil stores potential energy instead.
  • When the external source is turned off, the stored energy is released: charge continues to flow, propelled by the magnetic force, converting potential energy to kinetic energy until current stops.
  • To restore energy, the external source must be turned back on, restoring current and the magnetic field.

⚡ Two equivalent views

  • Force view: potential energy is associated with the magnetic force applied to current (F = q v × B).
  • Field view: potential energy is stored in the magnetic field itself, associated with the current distribution.
  • Both views are equally valid; the excerpt emphasizes that energy storage depends on the magnetic field.

📐 Definition and calculation

📐 Basic definition (single linkage)

Inductance L = Φ / I (single linkage)

  • Φ (magnetic flux, units: Wb) is the magnetic flux created by the current.
  • I (current, units: A) is the current responsible for this flux.
  • L (inductance, units: H) is the ratio.
  • A device with high inductance generates large magnetic flux for a given current, storing more energy than a device with lower inductance.

🔁 Magnetic flux definition

Φ = ∫_S B · ds

  • B is magnetic flux density (units: Wb/m²).
  • S is a surface over which flux is integrated.
  • ds is the differential surface area vector, normal to S.

🔗 Which surface and direction?

  • The current I must form a closed loop (C).
  • S is any surface bounded by C.
  • The direction of ds follows the right-hand rule (Stokes' Theorem convention).
  • Any surface bounded by C works because magnetic field lines form closed loops; each field line intersects any open surface bounded by C exactly once (called a "linkage").
  • Example: Figure 7.16 shows a closed loop C carrying current I, with surface S bounded by C and ds oriented by the right-hand rule.

🧲 Multiple windings (identical linkages)

L = N Φ / I (identical linkages)

  • Many structures (e.g., coils) consist of multiple loops.
  • Each winding carries the same current I.
  • The magnetic fields of the windings add; flux density inside a coil is proportional to the number of windings N.
  • N counts how many times the same current I creates a unique set of magnetic field lines intersecting S.
  • This is the complete engineering definition of inductance.

Don't confuse: If loops have different shapes (e.g., a cone-shaped coil), determining N Φ becomes more complex and is beyond the scope of this section.

🔌 Connection to circuit theory

🔌 Voltage-current relationship

Starting from L = N Φ / I, rearrange to get:

  • I = N Φ / L

Take the time-derivative of both sides:

  • dI/dt = (N / L) · (dΦ/dt)

From Faraday's Law (Section 8.3):

  • A change in Φ due to a change in current creates an electrical potential equal to −N dΦ/dt over the loop C.
  • Terminal voltage V = +N dΦ/dt (sign convention for passive devices).
  • Therefore, dΦ/dt = V / N.

Substitute into the derivative equation:

V = L dI/dt

This is the expected relationship from elementary circuit theory.

🔗 Mutual inductance

  • Inductance relates changes in current to voltage in the same device.
  • Mutual inductance relates changes in current in one device to voltage in a different device.
  • This occurs when two devices are coupled by the same magnetic field (e.g., transformer coils linked by the same field lines).
  • Voltage across one coil = (time-derivative of current in the other coil) × (mutual inductance).

⚠️ Common misconception

⚠️ "Pin inductance" or "lead inductance"

The misconception: If current does not form a closed loop (e.g., a pin or lead of an electronic component), what is the inductance?

The formal definition does not apply:

  • The ratio Φ / I requires a closed loop.
  • A pin or lead is not a closed loop, so the formal definition (ratio of magnetic flux to current) does not apply.

What "pin inductance" actually means:

  • It refers to the imaginary part of the impedance (reactance) expressed as an equivalent inductance.
  • Any device with positive reactance can be modeled as an equivalent inductance in a circuit diagram.
  • This does not refer to energy storage in a magnetic field; it is a circuit-theory modeling convenience.
  • The actual culprit is often skin effect, not physical inductance.

Key distinction:

Inductance implies positive reactance, but positive reactance does not imply the physical mechanism of inductance.

Example: An engineer might say "this pin has 2 nH of inductance," meaning the pin's reactance can be modeled as a 2 nH inductor in a circuit, not that the pin stores energy in a magnetic field.

🧩 What determines inductance

🧩 Geometry and permeability

FactorEffect
Geometry of current-bearing structuresDetermines how the magnetic field is shaped and distributed
Permeability of the intervening mediumAffects the strength of the magnetic field for a given current
  • Inductance depends on these two factors.
  • Inductance does not depend on current; current is viewed as either a stimulus (input) or response (output).
  • The corresponding response or stimulus is the magnetic flux associated with the current.

🔄 Inductor as a device

An inductor is a device designed to exhibit a specified inductance.

  • Inductors are engineered to have particular geometries and use materials with specific permeabilities to achieve desired inductance values.
92

Inductance of a Straight Coil

7.13 Inductance of a Straight Coil

🧭 Overview

🧠 One-sentence thesis

The inductance of a long straight coil is proportional to the square of the number of windings, cross-sectional area, and permeability, but inversely proportional to length, and depends only on area—not shape—of the cross-section.

📌 Key points (3–5)

  • What the section calculates: inductance of a straight cylindrical coil with uniform winding density, using the ratio of magnetic flux linkages to current.
  • Key approximation: assumes the coil is long enough (length much greater than radius) that the uniform interior field dominates and fringing fields at the ends are negligible.
  • Inductance formula: inductance equals permeability times the square of the number of windings times cross-sectional area, divided by length.
  • Surprising scaling: inductance grows with the square of the number of windings (N²), not linearly, because field strength increases with N and there are N flux linkages.
  • Common confusion: "pin inductance" or "lead inductance" does not refer to actual energy storage in a magnetic field but to positive reactance (equivalent inductance from a circuit perspective), often caused by skin effect rather than true inductance.

🔧 Setup and geometry

🔧 Coil structure

  • The coil is circular with radius a and length l.
  • It consists of N windings of wire wound with uniform winding density.
  • The problem uses cylindrical coordinates with the coil axis aligned along the z axis.

🔧 Key assumption: high winding density

  • The winding density N/l is assumed large enough that magnetic field lines cannot enter or exit between windings.
  • Field lines must traverse the entire length of the coil.
  • This ensures the interior field is uniform and well-described by the standard formula for a long coil.

🧲 Deriving the inductance

🧲 Starting formula

Inductance L is given by L = N Φ / I, where I is current and Φ is the magnetic flux associated with one winding of the coil.

  • Magnetic flux Φ is the integral of magnetic flux density B over the surface S bounded by a single current loop.
  • The surface element ds points in the direction determined by the right-hand rule with respect to positive current flow.

🧲 Magnetic field inside the coil

  • The magnetic flux density deep inside the coil is approximately B ≈ ẑ μ N I / l.
  • This approximation is valid when the coil is long relative to its radius (l much greater than a).
  • Justification: if l is sufficiently large compared to a, energy stored in the fringing field near the ends is negligible compared to energy in the uniform interior field.
  • The alternative (accounting for fringing fields) leads to a much more complicated problem.

🧲 Calculating magnetic flux

  • A natural choice for the surface S is the interior cross-section of the coil in a plane perpendicular to the axis.
  • The direction of ds is +ẑ (fingers of the right hand point this way when current flows as indicated).
  • Flux calculation: Φ ≈ (μ N I / l) × A, where A is the cross-sectional area of the coil.
  • The integral simplifies because the field is uniform over the cross-section.

🧲 Final inductance formula

  • Substituting into L = N Φ / I gives:

    L ≈ μ N² A / l (valid when l is much greater than a)

  • Dimensional check: permeability (H/m) times area (m²) divided by length (m) gives henries (H), as expected.

📐 Understanding the result

📐 Proportionality to N²

  • Inductance is proportional to the square of the number of windings, not just N.
  • Why: field strength increases with N, and independently there are N flux linkages.
  • Example: doubling the number of windings quadruples the inductance.

📐 Dependence on geometry

ParameterEffect on inductanceReason
Permeability μProportionalHigher permeability → stronger field for same current
Cross-sectional area AProportionalLarger area → more flux through each loop
Length lInversely proportionalLonger coil → weaker field for same number of windings
Shape of cross-sectionNo effectOnly the area matters, not the shape

📐 Shape independence

  • Inductance does not depend on the shape of the coil cross-section, only on the area.
  • Example: a circular cross-section and an elliptical cross-section with the same area yield the same inductance (assuming the same N, l, and μ).

⚠️ Limitations and common confusion

⚠️ Approximation validity

  • The result is approximate because it neglects:
    • The non-uniform fringing field near the ends of the coil.
    • The possibility that magnetic field lines escape between windings if winding density is inadequate.
  • Despite these limitations, the formula facilitates useful engineering analysis and design.

⚠️ "Pin inductance" misconception

Inductance implies positive reactance, but positive reactance does not imply the physical mechanism of inductance.

  • Engineers sometimes refer to "pin inductance" or "lead inductance" of an electronic component.
  • A pin or lead is not a closed loop, so the formal definition of inductance (ratio of magnetic flux to current) does not apply.
  • What is actually meant: the imaginary part of the impedance (reactance) expressed as an equivalent inductance.
  • This is a circuit-theory perspective: any device with positive reactance can be modeled as an equivalent inductor in a circuit diagram.
  • Don't confuse: this does not refer to energy storage in a magnetic field; the culprit is often skin effect, not true inductance.
  • Example: a straight wire pin exhibits positive reactance due to skin effect, so it can be modeled as an inductor, but it does not store significant energy in a magnetic field.
93

Inductance of a Coaxial Structure

7.14 Inductance of a Coaxial Structure

🧭 Overview

🧠 One-sentence thesis

The inductance of a coaxial structure can be determined by assuming a current on the inner conductor, integrating the resulting magnetic field to find the magnetic flux between conductors, and then computing the ratio of flux to current.

📌 Key points (3–5)

  • Core method: Model the structure as two concentric perfectly-conducting cylinders, assume current I on the inner conductor, find the magnetic field, integrate to get flux Φ, then calculate inductance as L = Φ / I.
  • Key insight from Ampere's Law: The magnetic field inside a coaxial structure with concentric conductors carrying current I is identical to the field of a line current I in free space—the outer conductor does not change the radial symmetry.
  • Final formula: Inductance per unit length is L′ = (μ / 2π) ln(b / a), depending only on material permeability μ and geometry (radii a and b), not on current.
  • Common confusion: This derivation is similar to finding the capacitance of a coaxial structure (Section 5.24); both use symmetry and integration over fields, but one deals with electric fields and the other with magnetic fields.
  • Practical application: This result is needed to determine the characteristic impedance of coaxial transmission lines and to find lumped-element parameters for transmission line models.

🔧 Physical model and setup

🔧 Geometry of the coaxial structure

The structure consists of:

  • Two concentric perfectly-conducting cylinders with radii a (inner) and b (outer).
  • A homogeneous material with permeability μ separating the conductors.
  • The +z axis placed along the common axis, so cylinders are described by surfaces ρ = a and ρ = b.
  • Length l along the z-axis.

🎯 Purpose of the analysis

  • Determine the inductance of the coaxial structure.
  • This is particularly important for finding the characteristic impedance of coaxial transmission lines (Section 3.10).
  • The analysis applies to a short section of a longer structure, yielding inductance per unit length for lumped-element equivalent circuit models (Section 3.4).

🔄 Analogy to capacitance derivation

The excerpt notes:

The derivation is similar to the derivation of the capacitance of a coaxial structure, addressed in Section 5.24.

  • Both use symmetry and field integration.
  • Reviewing the capacitance derivation may help before attempting this one.

🧲 Magnetic field inside the structure

🧲 Finding the magnetic field intensity

The first step is to determine the magnetic field inside the structure:

  • Assume fringing fields are negligible, so the internal field is constant with respect to z.
  • A current I flowing in the +z direction on the inner conductor creates a magnetic field.
  • From Section 7.5, the magnetic field intensity is:
    • H = (direction φ̂) × I / (2πρ), for aρb.

🔍 Why the outer conductor doesn't change the field

The excerpt emphasizes an important principle using Ampere's Law:

The magnetic field inside a coaxial structure comprised of concentric conductors bearing current I is identical to the magnetic field of the line current I in free space.

Why this is true:

  • Ampere's Law states: the line integral of H around a closed path C equals the enclosed current I_encl.
  • For a circular path with radius between a and b, the enclosed current is simply I.
  • The presence of the outer conductor does not change the radial symmetry.
  • Nothing else remains that can alter the outcome.

Don't confuse: This does not mean the outer conductor is irrelevant—it defines the boundary and return path for current—but it does not alter the field distribution between the conductors.

🧲 Converting to magnetic flux density

To find magnetic flux, we need magnetic flux density B instead of H:

  • They are related by the permeability: B = μ H.
  • Thus: B = (direction φ̂) × μI / (2πρ), for aρb.

🔢 Calculating flux and inductance

🔢 Integrating to find magnetic flux

Magnetic flux Φ is obtained by integrating over the magnetic flux density:

  • Φ = integral over surface S of B · ds.
  • S is any open surface through which all magnetic field lines within the structure must pass.

Choosing the simplest surface:

  • The simplest surface is a plane of constant φ (a constant-coordinate surface perpendicular to the field lines).
  • This is shown as the shaded area in Figure 7.18.

Performing the integration:

  • Integrate from ρ = a to ρ = b and from z = 0 to z = l.
  • The result is: Φ = (μIl / 2π) × ln(b / a).

🔢 Deriving inductance from the definition

Using the definition of inductance (Section 7.12):

L = Φ / I

Substituting the flux:

  • L = [(μIl / 2π) × ln(b / a)] / I.
  • The current I cancels out in numerator and denominator.
  • Final result: L = (μl / 2π) × ln(b / a).

Dimensional check:

  • Units are henries (H), as expected.
  • Permeability μ has units H/m, length l has units m, so (μl) has units H.

📏 Inductance per unit length

For transmission line applications, divide by length l:

  • L′ = (μ / 2π) × ln(b / a).
  • Units: H/m (henries per meter).

🎓 Key properties and example

🎓 What the formula depends on

The inductance depends only on:

  • Material properties: permeability μ.
  • Geometry: length l, inner radius a, outer radius b.

Notably:

  • Inductance does not depend on current I, which would imply non-linear behavior.
  • The formula is linear with respect to all parameters.

📊 Comparison of dependencies

ParameterEffect on inductance
Permeability μDirectly proportional
Length lDirectly proportional
Ratio b/aProportional to ln(b/a)
Current INo dependence (linear device)

🧪 Example: RG-59 coaxial cable

The excerpt provides a concrete example:

Given:

  • Inner conductor radius: a = 0.292 mm.
  • Outer conductor radius: b = 1.855 mm.
  • Spacing material: polyethylene (non-magnetic), so μμ₀.

Solution:

  • Using the formula L′ = (μ / 2π) × ln(b / a).
  • Result: L′ ≈ 370 nH/m (nanohenries per meter).

Don't confuse: This is inductance per unit length, not total inductance; multiply by the cable length to get total inductance.

94

Magnetic Energy

7.15 Magnetic Energy

🧭 Overview

🧠 One-sentence thesis

Magnetic energy stored in an inductor is proportional to inductance and the square of current, and can be calculated either from circuit parameters or by integrating the energy density over the volume containing the magnetic field.

📌 Key points (3–5)

  • Energy storage mechanism: inductors store energy in the magnetic field when current flows, interpreted as potential energy from forces between current-bearing elements.
  • Circuit-level formula: energy stored in an inductor is one-half times inductance times current squared (W_m = ½LI²).
  • Field-level formula: energy density in a magnetic field is one-half times permeability times magnetic field intensity squared (w_m = ½μH²).
  • Common confusion: the circuit formula (½LI²) applies to a specific inductor structure, but the field formula (½μH²) is completely general and works for any volume with any field distribution.
  • Practical relevance: periodic energizing and de-energizing of inductors consumes power in electrical systems, especially power systems.

⚡ Why magnetic energy matters

💡 Energy consumption in electrical systems

  • Any system with inductance uses a fraction of the power supply's energy to energize the magnetic field.
  • In many electronic systems—particularly power systems—inductors are energized and de-energized at regular intervals.
  • Since power is energy per unit time, this periodic cycling consumes power.
  • Energy storage in inductors therefore contributes to the overall power consumption of electrical systems.

🔋 Interpreting stored energy

  • When current flows through an inductor, current-bearing elements exert forces on each other.
  • Since these elements are normally not free to move, this force manifests as potential energy stored in the magnetic field.
  • This concept connects back to the force relationships discussed in Section 7.12.

🔌 Circuit-theory derivation

📐 Starting relationships

The derivation begins with two fundamental circuit relationships:

Voltage-current relationship: v(t) = L × (d/dt)i(t), where v(t) is electrical potential difference (V), i(t) is current (A), and L is inductance (H).

Instantaneous power: p(t) = v(t) × i(t), the product of voltage and current.

🧮 Integration over time

  • Energy (J) equals power (J/s) integrated over time.
  • At some past time t₀, current i(t₀) = 0 and stored energy W_m = 0.
  • As current is applied, W_m increases monotonically.
  • At present time t, current i(t) = I (the final steady current).

The magnetic energy is:

  • W_m = integral from t₀ to t of p(τ) dτ
  • Substituting the voltage-current relationship: W_m = integral of [L × (d/dτ)i(τ)] × i(τ) dτ
  • Changing the integration variable from time τ to current i: W_m = L × integral from 0 to I of i di
  • Evaluating: W_m = ½LI²

🎯 Circuit formula result

Energy stored in an inductor: W_m = ½LI², where L is inductance and I is steady current.

  • Energy increases in proportion to inductance.
  • Energy increases in proportion to the square of current (doubling current quadruples stored energy).

🧲 Field-theory perspective

🔁 Long straight coil example

The excerpt uses a long straight coil to connect circuit and field perspectives. For this structure:

  • Inductance: L = μN²A / l

    • μ = permeability
    • N = number of windings
    • A = cross-sectional area
    • l = length
  • Magnetic field intensity inside: H = NI / l

🌐 Deriving energy density

Substituting the coil formulas into W_m = ½LI²:

  • W_m = ½ × [μN²A / l] × [Hl / N]²
  • Simplifying: W_m = ½μH² × Al

Since the magnetic field inside a long coil is approximately uniform, the energy density is uniform throughout the volume Al:

Magnetic energy density: w_m = ½μH², with units of energy per unit volume (J/m³).

🗺️ General volume integration

The energy density formula provides an alternative method to compute total magnetostatic energy in any structure:

Total magnetic energy in volume V: W_m = integral over V of w_m dv = ½ × integral over V of μH² dv

Key features:

  • Works even if the magnetic field varies with position.
  • Works even if permeability varies with position.
  • Completely general—not limited to coils or uniform fields.

Example: To find energy in a complex structure, define a mathematical volume V containing the field, then integrate ½μH² over that volume.

⚠️ Don't confuse: circuit vs field formulas

FormulaScopeWhen to use
W_m = ½LI²Specific inductor with known L and ICircuit analysis; known inductance
W_m = ∫ ½μH² dvAny volume with any field distributionField analysis; varying fields or permeability
  • The circuit formula is derived from the field formula for the special case of a coil.
  • The field formula (Equations 7.87 and 7.89) is the fundamental, general result.

🔬 Physical interpretation

📈 Energy and permeability

  • Energy stored increases with the permeability μ of the medium.
  • This makes physical sense: inductance is proportional to permeability, so higher permeability means more energy storage for the same current.

🎓 Generality of the result

Although the derivation used the long straight coil as an example:

  • The energy density formula w_m = ½μH² applies to any magnetic field configuration.
  • The volume integral formula applies to any defined volume, regardless of geometry or field uniformity.
  • These are fundamental results in magnetostatics, not limited to idealized structures.
95

Magnetic Materials

7.16 Magnetic Materials

🧭 Overview

🧠 One-sentence thesis

Magnetic materials exhibit permeability significantly different from free space, with ferromagnetic materials showing orders-of-magnitude greater permeability and nonlinear behavior that enables both permanent magnetization and memory storage applications.

📌 Key points (3–5)

  • What defines a magnetic material: permeability μ significantly different from free space μ₀, causing the material itself to become a source of magnetic field when an external field is applied.
  • Three mechanisms: paramagnetism, diamagnetism, and ferromagnetism—each involves quantum mechanical processes at atomic/subatomic levels.
  • Common confusion: paramagnetic vs diamagnetic—both have permeability only slightly different from μ₀ (less than 0.01%), but paramagnetic materials align with the external field and show weak persistence, while diamagnetic materials align opposite to the external field and show no persistence.
  • Ferromagnetic nonlinearity: these materials exhibit saturation (B stops increasing beyond a certain H) and hysteresis (B depends on the history of H, not just the current value).
  • Why it matters: ferromagnetic hysteresis enables permanent magnets and digital memory storage devices like hard drives.

🧲 Classification of magnetic materials

🧲 Defining magnetic materials

A magnetic material: a substance that exhibits permeability μ that is significantly different from the permeability of free space μ₀.

  • Because magnetic flux density B relates to magnetic field intensity H via B = μH, magnetic materials can exhibit much greater B for a given H than other materials.
  • Magnetic materials are "magnetizable": applying an external magnetic field causes the material itself to become a source of magnetic field.
  • Typically metals, semiconductors, or heterogeneous media (e.g., ferrite = iron particles suspended in ceramic).

🔬 Three physical mechanisms

All three mechanisms involve quantum mechanical processes beyond classical physics:

MechanismPermeability vs μ₀Magnetization strengthPersistence
ParamagnetismVery slightly different (< 0.01%)Very weakWeak persistent field
DiamagnetismVery slightly different (< 0.01%)Very weakNo persistent field
FerromagnetismOrders of magnitude greaterStrong, readily magnetizableIndefinite (permanent magnets)

🔄 Paramagnetic and diamagnetic materials

🔄 Shared characteristics

  • Both exhibit permeability only very slightly different from μ₀, typically by much less than 0.01%.
  • Both show very weak and temporary magnetization.
  • Magnetization is typically so weak it is not often a consideration in engineering analysis and design.

⚖️ Key distinction: alignment and persistence

Paramagnetic materials (aluminum, magnesium, platinum):

  • Exhibit a very weak persistent magnetic field.
  • Induced magnetic field is aligned in the same direction as the external (impressed) magnetic field.

Diamagnetic materials (copper, gold, silicon):

  • Do not exhibit a persistent magnetic field.
  • Induced magnetic field is aligned in the opposite direction as the external magnetic field (counter to intuition!).

Don't confuse: both are weak, but the direction of induced field and persistence differ.

⚡ Ferromagnetic materials

⚡ High permeability and strong magnetization

  • Permeability can be many orders of magnitude greater than μ₀ (see Appendix A.2 for example values).
  • Can be readily and indefinitely magnetized → permanent magnets are typically ferromagnetic.
  • Common examples: iron, nickel, cobalt.

📉 Nonlinearity: saturation

Ferromagnetic materials are significantly nonlinear (permeability μ is not constant).

Saturation occurs when:

  • As external field H increases, response field B initially increases according to B = μH.
  • The slope of the B-vs-H curve (which represents μ) is not constant.
  • Once H exceeds a certain value, B no longer significantly increases.
  • Further increases in external field do not significantly increase magnetization, so there is no significant increase in B.

Example: Starting from an unmagnetized state (origin: H = B = 0), increasing H causes B to rise along a curve; eventually B flattens out even as H continues to increase.

🔁 Nonlinearity: hysteresis

Hysteresis: a form of nonlinear behavior where the material's response depends on the history of magnetization, not just the current external field.

How hysteresis manifests:

  • Starting from saturation, reduce the external field H.
  • The rate of decrease in B with respect to H is significantly less than the rate B originally increased.
  • When H is reduced to zero, B is still greater than zero—the material remains magnetized.
  • Applying an external field in the reverse direction eventually zeros and then redirects B.
  • Continuing to decrease H (increase magnitude in reverse direction) reaches saturation again in the opposite direction.
  • Returning to the start condition (H = B = 0, demagnetized with no external field) is not possible.

Don't confuse: the same H value can correspond to different B values depending on whether H is increasing or decreasing and what the previous values were.

🛠️ Engineering implications of ferromagnetism

🛠️ Applications and design challenges

Permanent magnets:

  • Hysteresis is an important consideration in the analysis and design of magnets.
  • The material "remembers" its magnetization history.

High-permeability devices (inductors, transformers):

  • Ferromagnetic materials are used because high permeability is desired.
  • Hysteresis complicates design and imposes limits on device performance.

💾 Memory storage

Hysteresis can be exploited as a form of memory:

  • If B > 0, recent values of H must have been relatively large and positive.
  • If B < 0, recent values of H must have been relatively large and negative.
  • The most recent sign of H can be inferred even if the present value of H is zero.
  • The material "remembers" the past history of its magnetization and thereby exhibits memory.

Example: This is the enabling principle for digital data storage devices, including hard drives.

📐 Magnetostatic energy context

📐 Energy density and total energy

The excerpt begins with a formula for total magnetostatic energy in any structure:

  • Total magnetostatic energy W_m equals the integral of energy density over volume V.
  • Energy density is given by (1/2) μH² (one-half times permeability times H squared).
  • Total energy: W_m = integral over V of (1/2) μH² dv.

Why this matters for magnetic materials:

  • Energy stored by the magnetic field increases with the permeability of the medium.
  • This makes sense because inductance is proportional to permeability.
  • Although derived using a long straight coil example, the formulas are completely general.
96

Comparison of Static and Time-Varying Electromagnetics

8.1 Comparison of Static and Time-Varying Electromagnetics

🧭 Overview

🧠 One-sentence thesis

Time-varying electromagnetics differs from static electromagnetics by coupling electric and magnetic fields through time-derivative terms in Maxwell's Equations, enabling waves to exist even after their sources are removed.

📌 Key points (3–5)

  • Static vs time-varying: in static cases (electrostatics and magnetostatics), electric and magnetic fields are independent; in time-varying cases, they are coupled.
  • Extra terms in Maxwell's Equations: time-varying versions include time derivatives of fields that describe coupling between electric and magnetic fields.
  • Waves as a consequence: coupling allows fields to persist after charges and currents are turned off, creating waves (e.g., signals in transmission lines and antenna radiation).
  • Common confusion: the static equations are not wrong—they are special cases; the time-varying equations reduce to the static forms when fields do not change with time.

⚡ Static vs time-varying field behavior

⚡ Independence vs coupling

  • In static cases: electric and magnetic fields are independent—changes in one do not affect the other.
  • In time-varying cases: electric and magnetic fields are coupled—a changing magnetic field induces an electric field, and vice versa.
  • This coupling is the fundamental difference that drives all other distinctions.

🌊 Waves and source independence

A wave is a field that can continue to exist even after its sources (charges and currents) are turned off.

  • In static electromagnetics, fields disappear when sources are removed.
  • In time-varying electromagnetics, coupling allows fields to sustain themselves.
  • Example: signals propagating away from an antenna continue to exist in space even after the antenna stops transmitting.
  • Example: signals in transmission lines persist as traveling waves.

📐 Maxwell's Equations comparison

📐 Integral form differences

The table in the excerpt highlights differences in blue. Key changes:

EquationStatic formTime-varying formWhat changed
Gauss's law (electric)∮ D · ds = Q_encl∮ D · ds = Q_enclNo change
Faraday's law∮ E · dl = 0∮ E · dl = −∂/∂t ∫ B · dsAdded time derivative of magnetic flux
Gauss's law (magnetic)∮ B · ds = 0∮ B · ds = 0No change
Ampère's law∮ H · ds = I_encl∮ H · dl = I_encl + ∫ ∂/∂t D · dsAdded time derivative of electric flux
  • Faraday's law: the right-hand side is zero in the static case (no induced electric field); in the time-varying case, a changing magnetic flux induces a circulating electric field.
  • Ampère's law: the static case includes only enclosed current; the time-varying case adds a term for the rate of change of electric flux (displacement current).

📐 Differential form differences

The differential forms show the same pattern:

EquationStatic formTime-varying formWhat changed
Divergence of D∇ · D = ρ_v∇ · D = ρ_vNo change
Curl of E∇ × E = 0∇ × E = −∂/∂t BAdded time derivative of B
Divergence of B∇ · B = 0∇ · B = 0No change
Curl of H∇ × H = J∇ × H = J + ∂/∂t DAdded time derivative of D
  • Curl of E: zero in static cases (conservative field); nonzero in time-varying cases when the magnetic field changes.
  • Curl of H: includes only current density J in static cases; includes an additional term for the rate of change of electric displacement in time-varying cases.

🔗 The coupling mechanism

🔗 Time derivatives as coupling terms

  • The terms involving ∂/∂t (time derivatives) are what couple electric and magnetic fields.
  • A changing magnetic field (∂B/∂t) produces a curl in the electric field (∇ × E ≠ 0).
  • A changing electric field (∂D/∂t) contributes to the curl of the magnetic field (∇ × H).
  • Without these time-derivative terms, the fields remain independent.

🔗 Why static equations are special cases

  • When fields do not change with time, all time derivatives are zero.
  • Setting ∂/∂t = 0 in the time-varying equations recovers the static equations.
  • Don't confuse: static electromagnetics is not a separate theory—it is the limit of time-varying electromagnetics when nothing changes.

🎯 Practical implications

🎯 Signals and transmission

  • Time-varying electromagnetics explains how signals propagate in transmission lines.
  • Coupling allows energy to travel as waves, not just as static fields tied to sources.

🎯 Antennas and radiation

  • Antennas rely on time-varying fields to radiate waves into space.
  • Once radiated, these waves propagate independently of the antenna (the source).
  • This is only possible because of the coupling terms in Maxwell's Equations.
97

Electromagnetic Induction

8.2 Electromagnetic Induction

🧭 Overview

🧠 One-sentence thesis

Electromagnetic induction is the phenomenon where a time-varying magnetic field induces an electrical potential difference across a conducting structure, and Lenz's Law predicts that the induced magnetic field always opposes the change in the impressed magnetic field.

📌 Key points (3–5)

  • What electromagnetic induction is: when a time-varying magnetic field creates an induced electrical potential difference (voltage) across a conductor.
  • Lenz's Law predicts direction: the induced current creates a magnetic field that opposes the change in the impressed (external) magnetic field.
  • What is actually induced: the induced quantity is voltage (V), not current directly—current is simply a response to the induced potential.
  • Common confusion: Lenz's Law describes the direction of induced effects but does not explain the underlying physics or give magnitudes; Faraday's Law provides the complete picture.
  • Energy conservation basis: Lenz's Law follows from conservation of energy—if the induced field reinforced the change, positive feedback would require an impossible external energy source.

🧲 The basic phenomenon

🧲 What happens during electromagnetic induction

Electromagnetic induction: when an electrically-conducting structure is exposed to a time-varying magnetic field, an electrical potential difference is induced across the structure.

  • The key requirement is time-varying magnetic field—a static (constant) magnetic field does not induce voltage.
  • The conducting structure can be a coil, loop, or any conductor.
  • Three quantities change together: induced voltage V, induced current I, and induced magnetic field B_ind.

🔬 The coil-and-magnet experiment

The excerpt describes an experiment with a cylindrical coil connected to a resistor and a bar magnet:

  • Magnet motionless: no current flows, no induced field, total field equals only the impressed field B_imp.
  • Magnet moving toward coil: positive current flows, creating induced field B_ind pointing right (opposite to B_imp which points left)—the induced field opposes the increase.
  • Magnet moving away: negative current flows, creating B_ind pointing left (same direction as B_imp)—the induced field opposes the decrease.

Example: When the magnet approaches, the impressed field inside the coil grows stronger; the induced current creates a field pointing the opposite way, acting to reduce the total field increase.

⚖️ Lenz's Law

⚖️ The core principle

Lenz's Law: the current that is induced by a change in an impressed magnetic field creates an induced magnetic field that opposes (acts to reduce the effect of) the change in the total magnetic field.

  • The induced magnetic field always opposes the change, not the field itself.
  • If the impressed field is increasing, the induced field points opposite to reduce the increase.
  • If the impressed field is decreasing, the induced field points in the same direction to reduce the decrease.

🔄 What Lenz's Law does and doesn't tell you

What it does:

  • Quickly determines the direction of induced current flow without detailed mathematics.
  • Useful for practical problem-solving in electromagnetic induction scenarios.

What it doesn't:

  • Does not explain the underlying physics mechanism.
  • Does not give the magnitude of induced voltage or current.
  • Does not clarify which quantity is directly induced versus which are responses.

⚠️ Common confusion: what is actually induced

  • It is easy to think current I is induced directly, but voltage V is the quantity actually induced.
  • Current flow is simply a response to the induced potential difference.
  • This can be verified by replacing the resistor with a high-impedance voltmeter—voltage changes even with negligible current.
  • Don't confuse: people informally say "current is induced," but this is indirect through voltage.

🔌 Transformer example

🔌 How induction works through a transformer

The excerpt provides a transformer example with two coils wound around a common toroidal core:

Setup:

  • Left circuit: battery, switch, and coil.
  • Right circuit: voltmeter and coil.
  • The core contains magnetic flux so flux from either coil reaches the other.

What happens when the switch closes:

  1. Current flows counter-clockwise in the left circuit (bottom terminal entry).
  2. This creates impressed field B_imp oriented counter-clockwise through the core.
  3. The right coil "sees" B_imp increase from zero.
  4. By Lenz's Law, if current could flow in the right circuit, it would flow counter-clockwise to create clockwise B_ind opposing the increase.
  5. Therefore, potential at the bottom of the right coil is higher than at the top.
  6. The voltmeter needle deflects right (temporarily, only during the change).

Key insight: The voltmeter reading returns to zero once current becomes constant—the induced voltage responds only to the change in B_imp, not to a steady field.

🌀 Energy conservation foundation

🌀 Why Lenz's Law must be true

Lenz's Law can be deduced from conservation of energy:

  • If induced field reinforced the change: the sum magnetic field would increase further, causing more induced field, creating positive feedback.
  • Problem with positive feedback: it cannot be sustained without an external energy source, leading to logical contradiction.
  • Conclusion: conservation of energy requires the negative feedback described by Lenz's Law—the induced field must oppose the change.

Example scenario: If moving a magnet toward a coil caused an induced field that attracted the magnet more strongly, the system would accelerate indefinitely without energy input—impossible.

🔗 Connection to Faraday's Law

🔗 The more complete picture

The excerpt introduces Faraday's Law as providing the full explanation:

  • Lenz's Law tells the direction of induced effects.
  • Faraday's Law describes the generation of electric potential by time-varying magnetic flux.
  • Faraday's Law gives the magnitude of induced voltage and explains the underlying physics.
  • For a single loop: V_T equals negative rate of change of magnetic flux (the excerpt shows the beginning of this formula).

Don't confuse: Lenz's Law is an observation useful for quick direction determination; Faraday's Law is the fundamental principle with complete mathematical description.

98

Faraday's Law

8.3 Faraday’s Law

🧭 Overview

🧠 One-sentence thesis

Faraday's Law states that a time-varying magnetic flux through a loop or coil induces an electric potential (emf) proportional to the rate of change of that flux, which is the fundamental mechanism behind transformers and generators.

📌 Key points (3–5)

  • What Faraday's Law describes: the generation of electric potential (emf) by a time-varying magnetic flux, not current directly—current is simply the induced voltage divided by resistance.
  • The mathematical relationship: the induced potential V_T equals the negative time derivative of magnetic flux Φ, multiplied by the number of windings N in a coil.
  • Two ways to generate emf: either vary the magnetic field B over time (transformer emf) or vary the loop's shape/orientation in a constant field (motional emf).
  • Common confusion: Lenz's Law seems to imply that current is induced, but Faraday's Law clarifies that potential is induced; the minus sign in Faraday's Law is Lenz's Law, ensuring the induced field opposes the change.
  • Generality: Faraday's Law is fundamental physics applicable to any closed path, not just wire loops, and describes the electromagnetic induction contribution to potential difference independent of static electric fields.

🔋 The core mechanism

⚡ What Faraday's Law states

Faraday's Law describes the generation of electric potential by a time-varying magnetic flux.

  • For a single loop: V_T = − (∂/∂t)Φ
  • For N identical windings tightly packed: V_T = −N(∂/∂t)Φ
  • The induced potential V_T is often called "emf" (electromotive force), a historical misnomer since it's potential, not force.
  • Each winding contributes a potential, and these add in series.

🧲 Magnetic flux Φ

Magnetic flux Φ (units: Wb) is the integral of magnetic flux density B over any open surface S bounded by the loop: Φ = ∫_S B · ds

  • B is magnetic flux density (units: T or Wb/m²)
  • ds is the differential surface area vector
  • Any surface bounded by the loop works—planar or non-planar—as long as every magnetic field line passing through the loop also passes through S
  • The magnitude of ds is the differential surface element, and its direction is the unit vector n̂ perpendicular to each point on S

🔄 The right-hand rule for orientation

  • There are two possible perpendicular directions for n̂; the choice affects the sign of V_T
  • Convention: Let curve C begin at the "−" terminal of V_T, follow the loop perimeter, and end at the "+" terminal
  • n̂ points in the direction of the curled fingers of the right hand when the thumb aligns with the direction of C
  • This is the same convention used in Stokes' Theorem

🔧 How to apply Faraday's Law (single loop scenario)

📋 Step-by-step procedure

  1. Assign polarity: Choose "+" and "−" terminals for the gap voltage V_T
  2. Determine n̂: Use the right-hand rule with C running from "−" to "+" around the loop perimeter
  3. Calculate flux: B yields magnetic flux Φ via the integral over any convenient open surface S that intersects all field lines through the loop
  4. Apply Faraday's Law: V_T is the negative time derivative of Φ
  5. Find current: Current I = V_T / R, with reference direction from "+" to "−" through the resistor (think of the loop as a voltage source)

⚠️ Key distinction: potential vs. current

  • Electromagnetic induction induces potential (V_T), not current directly
  • Current is simply the induced voltage divided by the loop's resistance
  • Don't confuse: Lenz's Law seems to focus on current, but Faraday's Law clarifies that potential is the primary induced quantity
  • The current I that circulates generates its own magnetic field (B_ind), distinct from the impressed field B, which opposes changes in B

🔀 Two mechanisms for generating emf

🔁 Transformer emf (time-varying B)

  • The magnetic field B itself varies with time while the loop remains stationary
  • Example mechanisms:
    • A permanent magnet is moved (translated or rotated) near the coil
    • A different coil bearing a time-varying current creates the changing B
  • This is the underlying principle of transformers

🔄 Motional emf (time-varying geometry)

  • The perimeter C and surface S change over time while B remains constant
  • Example mechanisms:
    • A wire loop is rotated in a constant magnetic field
    • A loop changes shape in a constant field
  • This is the underlying principle of generators

🔀 Combined mechanisms

  • Transformer and motional emf can exist in various combinations
  • From Faraday's Law's perspective, both are equivalent: in either case, Φ is time-varying, which is all that's required to generate emf

🌐 The minus sign and Lenz's Law

➖ Why the negative sign matters

  • The minus sign in V_T = −(∂/∂t)Φ is Lenz's Law embedded in Faraday's Law
  • It ensures that the induced current I generates a magnetic field that opposes the change in the impressed field B
  • This represents negative feedback, which is required by conservation of energy
  • Don't confuse: If the induced field reinforced the change (positive feedback), it would create a logical contradiction without an external energy source

🌍 Generality and broader implications

🔬 Faraday's Law as fundamental physics

  • Faraday's Law is not specific to wire loops or coils; it applies to any closed path, whether current-bearing or not
  • The computed potential difference is the contribution from electromagnetic induction, existing independently of and in addition to static electric field potentials
  • Example: You can define any closed path and compute the potential difference from traversing it using Faraday's Law

⚡ Connection to Maxwell-Faraday Equation

  • Faraday's Law indicates the contribution of electromagnetic induction to potential difference achieved by traversing a closed path
  • This insight transforms Kirchhoff's Voltage Law (which accounts for electric field only) into the Maxwell-Faraday Equation
  • The Maxwell-Faraday Equation is a general statement about the relationship between instantaneous electric field and the time derivative of magnetic field
99

Induction in a Motionless Loop

8.4 Induction in a Motionless Loop

🧭 Overview

🧠 One-sentence thesis

A motionless loop in a spatially-uniform but time-varying magnetic field generates an induced potential (transformer emf) proportional to the loop's projected area and the rate of change of the magnetic flux density, demonstrating Faraday's Law in action.

📌 Key points (3–5)

  • What this problem demonstrates: transformer emf in a stationary loop—no motion required, only a time-varying magnetic field.
  • Key result: induced potential equals (negative) projected area times the time derivative of magnetic flux density.
  • What matters for induction: the loop's area and orientation, the rate of change of the magnetic field, not the loop's shape or the absolute value of the field.
  • Common confusion: induction generates potential (voltage), not current directly; current flows only in response to the induced potential via Ohm's Law.
  • Verification tool: Lenz's Law confirms that the induced current's magnetic field opposes the change in the impressed field.

🔧 Problem setup and approach

🔧 The physical scenario

  • A single motionless loop of wire lies in the z = 0 plane.
  • The magnetic field is spatially uniform (same everywhere) but time-varying: B = B(t), where B(t) changes with time and is a constant direction.
  • A small gap in the loop allows measurement of induced potential V_T.
  • A resistor R is connected across the gap to allow current I to flow.

🧭 Sign conventions and reference directions

  • The polarity of V_T (+ and − terminals) is chosen arbitrarily; the result will reflect the actual direction.
  • Current I is directed according to the standard passive-device convention (from + to − through the resistor).
  • These are reference directions: a negative result means the actual direction is opposite.
  • The excerpt notes that reversing the terminal polarity and re-solving yields the same answer.

📐 Why shape doesn't matter

  • Because the magnetic field is spatially uniform, only the loop's area matters, not its specific shape.
  • A circular loop and a square loop with the same area and orientation produce the same induced potential.
  • This simplifies the problem: no need to specify radius or exact geometry.

🧮 Applying Faraday's Law

🧮 Starting with Faraday's Law

Faraday's Law: V_T = −N (∂/∂t) Φ

  • N = 1 (single loop).
  • Φ is the magnetic flux through the loop: Φ = ∫_S B · ds, where S is any open surface bounded by the loop.
  • The simplest choice: the planar surface defined by the loop's perimeter.

🔄 Determining the normal direction

  • The surface element is ds = ds, where is the normal to the loop's plane.
  • Which of the two possible normals? Use the right-hand rule (Stokes' Theorem):
    • Point the thumb from the − terminal to the + terminal along the loop's perimeter.
    • The curled fingers point in the direction of .
  • For the loop in the z = 0 plane with the indicated polarity, = +.

🧩 Simplifying the integral

  • Substitute B = B(t) into the flux integral:
    • V_T = − (∂/∂t) ∫_S ( B(t)) · ( ds)
    • Factor out constants: V_T = − ( · ) (∂/∂t) ∫_S B(t) ds
  • Because the field is uniform, B(t) can be extracted from the integral.
  • The loop's shape and orientation are time-invariant, so the remaining integral (the area A) can be extracted from the time derivative.

📊 Final expression for induced potential

V_T = − ( · A) (∂/∂t) B(t)

  • · A is the projected area of the loop.
  • Projected area equals A when the magnetic field is perpendicular to the loop ( = ).
  • Projected area decreases to zero as · → 0 (field parallel to the loop).

Summary statement from the excerpt:

The magnitude of the transformer emf induced by a spatially-uniform magnetic field is equal to the projected area times the time rate of change of the magnetic flux density, with a change of sign.

🔍 Key observations about the result

🔍 What affects the induced potential

FactorEffect on V_TExplanation
Loop area AProportionalLarger area → more flux → larger V_T
OrientationMaximized when perpendicularPeak when plane is ⊥ to field lines; zero when parallel
Rate of change ∂B/∂tProportionalFaster change → larger V_T; constant B → no induction
Loop shapeIrrelevantOnly area matters (for uniform field)

⚠️ No induction without change

  • If B is constant in time, ∂B/∂t = 0, so V_T = 0.
  • Induction requires time-varying flux, not just the presence of a magnetic field.
  • Example: when B(t) = 0 but ∂B/∂t is at maximum, V_T and I are also at maximum (see Example 8.3).

⚡ Current flows in response to induced potential

  • The current in the loop is I = V_T / R (Ohm's Law).
  • Electromagnetic induction induces potential; current is a secondary effect.
  • If the resistor is removed (R → ∞), then I → 0, but V_T remains unchanged.
  • Don't confuse: induction does not directly create current; it creates voltage, which then drives current through any available resistance.

📚 Worked examples

📚 Example 8.2: Linearly increasing magnetic field

Setup:

  • Circular loop, radius a = 10 cm, in the z = 0 plane.
  • Magnetic field B(t), where B(t) increases linearly from 0 to B₀ = 0.2 T over time t₀ = 3 s, then remains constant.
  • Resistor R = 1 kΩ closes the loop.

Solution steps:

  • = + (from right-hand rule with indicated polarity).
  • · A = A (field perpendicular to loop).
  • Area A = π a².
  • ∂B/∂t = B₀ / t₀ for 0 ≤ t ≤ t₀; zero before and after.
  • V_T = − π a² (B₀ / t₀) = −2.09 mV during the ramp-up; zero otherwise.
  • I = V_T / R = −2.09 μA during ramp-up; zero otherwise.

Interpretation:

  • Negative sign means current flows clockwise (opposite the reference direction).
  • Current exists only while B is changing, not before or after.

Lenz's Law check:

  • The induced current creates an induced magnetic field B_ind.
  • By the right-hand rule, B_ind points generally in the − direction inside the loop.
  • This opposes the increasing impressed field (+), consistent with Lenz's Law.

📚 Example 8.3: Sinusoidally varying magnetic field

Setup:

  • Same loop as Example 8.2.
  • Now B(t) = B₀ sin(2π f₀ t), with f₀ = 1 kHz.

Solution steps:

  • ∂B/∂t = 2π f₀ B₀ cos(2π f₀ t).
  • V_T = − 2π² f₀ a² B₀ cos(2π f₀ t).
  • I = V_T / R = − (2π² f₀ a² B₀ / R) cos(2π f₀ t).
  • Substituting values: I = −(39.5 mA) cos[(6.28 krad/s) t].

Interpretation:

  • V_T and I vary sinusoidally because the source B varies sinusoidally.
  • V_T and I are 90° out of phase with B(t): when B(t) = 0, V_T and I are at maximum magnitude.
  • This reinforces: it is the change in B that induces voltage and current, not B itself.

🧪 Key takeaway from examples

  • Linearly changing field → constant induced potential and current (during the change).
  • Sinusoidally varying field → sinusoidal induced potential and current, phase-shifted by 90°.
  • In both cases, no change in B → no induction.
100

8.5 Transformers: Principle of Operation

8.5 Transformers: Principle of Operation

🧭 Overview

🧠 One-sentence thesis

The ratio of coil voltages in an ideal transformer equals the ratio of turns (with sign determined by winding direction), a relationship derived from Faraday's Law through the shared magnetic field linking two coils.

📌 Key points (3–5)

  • What a transformer does: connects two electrical circuits through a shared magnetic field, enabling voltage conversion, impedance transformation, and circuit isolation.
  • Core principle: Faraday's Law—specifically transformer emf—governs operation; the change in magnetic field induces voltage, not the field itself.
  • The transformer law: the voltage ratio V₁/V₂ equals p·(N₁/N₂), where N₁ and N₂ are the number of turns and p = +1 (same winding direction) or −1 (opposite direction).
  • Common confusion: it is the rate of change of the magnetic field B that induces voltage and current, not the instantaneous value of B; voltage and current can be maximum even when B is zero.
  • Design consideration: toroidal cores confine the magnetic field to prevent electromagnetic interference (EMI) and electromagnetic compatibility (EMC) problems.

🔌 What a transformer is and does

🔌 Definition and applications

A transformer is a device that connects two electrical circuits through a shared magnetic field.

  • Applications include:
    • Impedance transformation
    • Voltage level conversion
    • Circuit isolation
    • Conversion between single-ended and differential signal modes
  • The underlying principle is Faraday's Law (transformer emf).

⚡ The key insight: change matters, not magnitude

  • The excerpt emphasizes: "it is the change in B that induces voltage and subsequently current, not B itself."
  • Example from the excerpt: when B(t) = 0 (which occurs twice per period), the induced voltage V_T and current I are at their maximum magnitude, not zero.
  • This happens because the derivative (rate of change) of B is maximum when B crosses zero.
  • Don't confuse: a zero magnetic field does not mean zero induced voltage; what matters is how fast the field is changing at that instant.

🧲 The two-coil experiment: deriving the transformer law

🧲 Experimental setup

The excerpt describes a simple experiment with two coils on a common axis:

  • Both coils wound on the same high-permeability core to contain the magnetic field.
  • Upper coil: N₁ turns.
  • Lower coil: N₂ turns, wound in the opposite direction.
  • Winding pitch is small so all magnetic field lines pass through the coil length; no lines pass between windings.
  • The magnetic field B in the core is assumed equal in both coils due to close spacing and high permeability.

🔄 Part I: voltage applied to upper coil

  • A sinusoidally-varying voltage source V₁⁽¹⁾ is connected to the upper coil.
  • This creates a current, which generates a time-varying magnetic field B in the core.
  • The lower coil (N₂ turns, opposite winding direction) is open-circuited.
  • By Faraday's Law, the induced potential in the lower coil is:
    • V₂⁽¹⁾ = −N₂ · (∂Φ₂/∂t)
    • where Φ₂ is the flux through a single turn of the lower coil.
  • Rewritten: V₂⁽¹⁾ = −N₂ · (∂/∂t) ∫_S B · (−ẑ ds)
  • The direction of ds = −ẑ ds is determined by the chosen polarity for V₂⁽¹⁾.

🔄 Part II: voltage applied to lower coil

  • Now a voltage V₂⁽²⁾ is applied to the lower coil, and the upper coil is open-circuited.
  • V₂⁽²⁾ is adjusted so the induced magnetic flux density is again B (equal to Part I).
  • The induced potential in the upper coil is:
    • V₁⁽²⁾ = −N₁ · (∂Φ₁/∂t)
    • which becomes V₁⁽²⁾ = −N₁ · (∂/∂t) ∫_S B · (+ẑ ds)
  • Shifting the minus sign into the integral: V₁⁽²⁾ = +N₁ · (∂/∂t) ∫_S B · (−ẑ ds)
  • Comparing with Part I, this can be rewritten in terms of the flux in the lower coil from Part I:
    • V₁⁽²⁾ = +N₁ · (∂Φ₂/∂t)
  • Expressing in terms of the potential from Part I:
    • V₁⁽²⁾ = (−N₁/N₂) · (−N₂ · ∂Φ₂/∂t) = (−N₁/N₂) · V₂⁽¹⁾

📐 The general transformer law

  • The excerpt concludes: the potential in the upper coil in Part II is related to the potential in the lower coil in Part I by a simple ratio.
  • If Part II had been done first, the same result would hold with superscripts swapped.
  • Therefore, regardless of termination arrangement:
    • V₁ = −(N₁/N₂) · V₂
  • The minus sign is a consequence of the coils being wound in opposite directions.

🔀 Generalizing for winding direction

The excerpt generalizes the expression:

V₁/V₂ = p · (N₁/N₂)

where:

  • p = +1 when coils are wound in the same direction.
  • p = −1 when coils are wound in opposite directions.

The transformer law: The ratio of coil voltages in an ideal transformer is equal to the ratio of turns with sign determined by the relative directions of the windings.

  • The excerpt notes it is "an excellent exercise" to confirm this by repeating the analysis with changed winding direction for either coil, which yields p = +1.

🛡️ Practical transformer design: toroidal cores

🛡️ Why use a toroidal core

  • A more familiar design: coils wound on a toroidal core (instead of a cylindrical core).
  • Purpose: confines the magnetic field linking the two coils to the core, preventing field lines from extending beyond the device.
  • Importance: confinement prevents fields from outside the transformer from interfering with the magnetic field linking the coils.
    • This avoids electromagnetic interference (EMI) and electromagnetic compatibility (EMC) problems.
  • The principle of operation is "in all other respects the same."

🛡️ Toroidal vs cylindrical cores

Core typeField confinementEMI/EMC risk
CylindricalField lines extend beyond deviceHigher risk of interference
ToroidalField confined to coreLower risk; better isolation
  • Example: in a toroidal transformer, the magnetic flux Φ circulates within the toroidal core, linking both coils without radiating into the surrounding space.

🔗 Connection to circuit theory

🔗 Familiar but with a twist

  • The excerpt notes the transformer law (Equation 8.19) "should be familiar from elementary circuit theory—except possibly for the minus sign."
  • The minus sign arises from the opposite winding directions in the experiment.
  • When coils are wound in the same direction, p = +1, and the familiar positive ratio is recovered.

🔗 Foundation for two-port analysis

  • The excerpt states: "This is the 'transformer law' of basic electric circuit theory, from which all other characteristics of transformers as two-port circuit devices can be obtained."
  • The excerpt references Section 8.6 for follow-up on two-port device characteristics (current and impedance ratios).
  • Don't confuse: the transformer law derived here is for an ideal transformer (no winding resistance, perfect flux containment).
101

Transformers as Two-Port Devices

8.6 Transformers as Two-Port Devices

🧭 Overview

🧠 One-sentence thesis

An ideal transformer scales voltage, current, and impedance according to its turns ratio, enabling signal transformation and isolation between circuits while exhibiting highpass frequency response due to practical limitations at low frequencies.

📌 Key points (3–5)

  • Voltage and current relationships: voltage ratio equals the turns ratio (times a polarity factor), while current ratio is the inverse (times the same polarity factor).
  • Impedance transformation: input impedance scales by the square of the turns ratio, independent of winding direction.
  • Power conservation: in an ideal transformer with no resistance and perfect flux containment, power delivered by the source equals power absorbed by the load.
  • Common confusion: transformers do not work at DC and exhibit poor low-frequency performance due to magnetic saturation; they are AC devices with highpass behavior.
  • Practical applications: transformers isolate DC offsets, convert between single-ended and differential signals (baluns), and provide bandwidth advantages in circuit design.

⚡ Voltage and current relationships

⚡ Voltage ratio

The relationship governing terminal voltages V₁ and V₂ is: V₁/V₂ = p·(N₁/N₂), where N₁ and N₂ are the number of turns in the associated coils and p is either +1 or −1 depending on the relative orientation of the windings.

  • The factor p indicates whether the reference direction of the associated fluxes is the same (+1) or opposite (−1).
  • This ratio depends only on the turns ratio and winding orientation.
  • Example: if the primary coil has twice as many turns as the secondary (N₁/N₂ = 2) and p = +1, then V₁ = 2·V₂.

🔄 Current ratio

The excerpt derives the current relationship from power conservation:

  • Power delivered by the source: V₁·I₁
  • Power absorbed by the load: −V₂·I₂
  • In an ideal transformer (no resistance, perfect flux containment), these must be equal: V₁·I₁ = −V₂·I₂

This yields:

I₁/I₂ = −V₂/V₁ = −p·(N₂/N₁)

  • Notice the current ratio is the inverse of the voltage ratio (with a minus sign from reference conventions).
  • Example: if voltage steps up by a factor of 2, current steps down by a factor of 2.
  • Don't confuse: the minus signs result from the specific reference polarities and directions chosen in the two-port model; other conventions may omit them.

🔌 Impedance transformation

🔌 Input impedance derivation

The excerpt defines:

  • Z₁: input impedance looking into port 1 from the source
  • Z₂: output impedance of port 2, looking out into the load

Starting from Z₁ = V₁/I₁ and substituting the voltage and current ratios:

Z₁ = [+p·(N₁/N₂)·V₂] / [−p·(N₂/N₁)·I₂] = −(N₁/N₂)²·(V₂/I₂)

Since Z₂ = −V₂/I₂ (the minus sign from reference conventions), substitution yields:

Z₁ = (N₁/N₂)²·Z₂

📐 Key impedance properties

PropertyFormulaMeaning
Impedance ratioZ₁/Z₂ = (N₁/N₂)²Scales by the square of the turns ratio
Independence from pNo p factor appearsImpedance transformation does not depend on winding direction
  • Why the square? Because impedance involves both voltage (scales as N₁/N₂) and current (scales inversely), so the ratio compounds.
  • Example: a 1:10 turns ratio (N₁/N₂ = 0.1) transforms a 50-ohm load into a 0.5-ohm input impedance.
  • The excerpt emphasizes: "Remarkably, the impedance transformation depends only on the turns ratio, and is independent of the relative direction of the windings (p)."

🚫 Frequency limitations

🚫 Why transformers don't work at DC

The excerpt explains this through Faraday's Law:

V = −N·(∂Φ/∂t)

  • If magnetic flux Φ is not time-varying, then ∂Φ/∂t = 0, so there is no induced electric potential.
  • "If the magnetic flux Φ isn't time-varying, then there is no induced electric potential, and subsequently no linking of the signals associated with the coils."
  • Don't confuse: transformers require changing flux; a constant (DC) flux produces zero voltage.

⚠️ Magnetic saturation at low frequencies

The excerpt integrates Faraday's Law to show:

Φ(t) = −(1/N)·∫[from t₀ to t] V(τ)dτ + Φ(t₀)

For a sinusoidally-varying voltage:

  • At very low frequencies, the integral accumulates to a large value.
  • Peak flux Φ becomes very large, which means the magnetic field becomes very large.
  • Problem: Most practical high-permeability core materials exhibit magnetic saturation—the rate at which the magnetic field can increase decreases with increasing field magnitude.
  • Result: "a transformer may work fine at (say) 1 MHz, but at (say) 1 Hz the transformer may exhibit an apparent loss associated with this saturation."

📶 Highpass frequency response

Thus, practical transformers exhibit highpass frequency response.

  • The relationships developed "should be viewed as AC expressions, and are not normally valid at DC."
  • This limitation is fundamental, not a design flaw.

🔧 Practical applications

🔧 DC isolation and noise filtering

The highpass behavior can be useful:

  • A transformer can isolate an undesired DC offset from one circuit to another.
  • It can also isolate low-frequency noise in the circuit attached to one coil from the circuit attached to the other coil.
  • The DC-isolating behavior allows the transformer to be used in various signal-conditioning applications.

🔀 Balun (balanced-unbalanced conversion)

A balun is a two-port device that transforms a single-ended ("unbalanced") signal—that is, one having an explicit connection to a datum (e.g., ground)—into a differential ("balanced") signal, for which there is no explicit connection to a datum.

  • Single-ended (unbalanced): signal has an explicit ground reference.
  • Differential (balanced): signal has no explicit ground connection; both conductors carry signal.
  • Differential signals have many benefits in circuit design (e.g., noise immunity).
  • Inputs and outputs to devices must often be in single-ended form, so conversion is needed.
  • Transformer advantage as balun: "it has one very big advantage, namely bandwidth"—transformers can operate over a wide frequency range (within their highpass limits).
  • Example use case: converting a single-ended antenna input to a differential amplifier input, or vice versa.
102

The Electric Generator

8.7 The Electric Generator

🧭 Overview

🧠 One-sentence thesis

An electric generator transforms mechanical rotation into electrical energy by using Faraday's Law: as a loop rotates in a magnetic field, the changing magnetic flux through the loop induces a time-varying voltage.

📌 Key points (3–5)

  • What a generator does: converts mechanical energy (rotation) into electrical energy via electromagnetic induction.
  • Motional emf: the induced voltage arises from changes in geometry (rotation) rather than changes in the magnetic field strength itself.
  • Key mechanism: rotating a loop in a static magnetic field changes the magnetic flux through the loop over time, which induces voltage according to Faraday's Law.
  • Common confusion: maximum voltage occurs when the plane of the loop is parallel to the magnetic field (flux is zero but changing fastest), not when flux is maximum.
  • Practical insight: the shape of the loop doesn't matter—only its area, the rotation frequency, and the magnetic field strength determine the induced voltage.

🔄 How generators work

⚙️ Basic principle

A generator is a device that transforms mechanical energy into electrical energy, typically by electromagnetic induction via Faraday's Law.

  • Mechanical rotation (e.g., from a gasoline engine, steam turbine, wind turbine, or hydroelectric flow) turns a crankshaft.
  • Attached to the crankshaft is a system of coils and/or magnets.
  • As the system rotates, the relative orientations of coils and magnetic field change over time.
  • This changing geometry produces a time-varying magnetic flux, which induces an electric potential.

🔁 Motional emf

Motional emf: induced potential due entirely to changes in geometry (motion) as opposed to changes in the magnitude of the magnetic field.

  • The magnetic field itself is static (constant in time and uniform in space).
  • The voltage arises because the loop is moving (rotating) through this field.
  • Example: coal-fired, steam-fired, hydroelectric generators, and wind turbines all operate on this principle.

🧲 The rudimentary single-loop generator

🔧 Setup

  • A planar loop rotates around the z-axis at frequency f₀ (one complete revolution takes 1/f₀ seconds).
  • The magnetic flux density B is constant in magnitude (B₀) and direction (unit vector ), uniform everywhere.
  • The loop is shown as circular, but the shape doesn't matter—only the area A of the loop matters.
  • The induced voltage V_T is measured across the loop terminals with a reference polarity.

📐 Why shape doesn't matter

  • Because the magnetic field is time-invariant and spatially uniform, the specific shape of the loop is irrelevant.
  • Only the area of the loop affects the result.
  • The simplest choice for the surface S (used in flux calculation) is the planar area bounded by the loop.

🧮 Calculating the induced voltage

📏 Faraday's Law applied

The induced potential is given by:

V_T = − (time derivative of magnetic flux Φ)

where magnetic flux Φ is:

Φ = integral over surface S of B · ds

  • The differential surface element ds = ds, where is the unit normal determined by the right-hand rule from the reference polarity of V_T.
  • Substituting: Φ = [ · ] B₀ A

🔄 Time-varying normal vector

  • At time t = 0, = +ŷ (pointing in the +y direction).
  • At t = 1/(4f₀), the loop has rotated one-quarter turn, so = −.
  • At t = 1/(2f₀), the loop has rotated half a turn, so = −ŷ.
  • The general expression for the rotating normal is:
    • (t) = − sin(2πf₀t) + ŷ cos(2πf₀t)

🔢 Final voltage expression

After taking the time derivative and simplifying:

V_T = +2πf₀ A B₀ [ · ρ̂(t)]

where ρ̂(t) = + cos(2πf₀t) + ŷ sin(2πf₀t) is the radial basis vector in cylindrical coordinates.

  • The voltage varies sinusoidally with frequency f₀.
  • Peak voltage magnitude: max |V_T(t)| = 2πf₀ A B₀
  • This maximum is achieved when the magnetic field B is polarized entirely in the x-y plane (perpendicular to the rotation axis z).

📊 Example calculation

Setup: circular loop of radius a = 1 cm, rotating at 1000 revolutions per second (f₀ = 1 kHz), in a static uniform magnetic field of 1 mT in the + direction.

ParameterValue
Frequency f₀1 kHz
Magnetic field B₀1 mT
Direction +
Loop area Aπ(0.01 m)²

Result: V_T(t) ≈ (1.97 mV) cos[(6.28 krad/s) t]

🔍 Key insights and common confusions

⚡ When is voltage maximum?

Don't confuse flux magnitude with rate of change:

  • Maximum voltage occurs when the plane of the loop is parallel to B.
    • At this instant, · (t) = 0, so Φ(t) = 0 (flux is zero).
    • But this is when Φ(t) is changing most rapidly (maximum time derivative).
  • Zero voltage occurs when the plane of the loop is perpendicular to B.
    • At this instant, · (t) = 1, so |Φ(t)| is maximum.
    • But the time derivative is zero (flux is momentarily not changing).

Why? Faraday's Law depends on the rate of change of flux, not the flux itself.

🔄 Relative motion

  • It doesn't matter whether the loop rotates in a stationary magnetic field or the magnetic field rotates around a stationary loop—the physics is the same.
  • In practical generators, both coils and magnetic fields (from magnets or other coils) may rotate.

🎯 Optimizing power generation

  • To maximize power, choose the magnetic field direction to maximize | · ρ̂(t)|.
  • This means B should be polarized entirely in the x-y plane (perpendicular to the rotation axis), with B · = 0.
103

8.8 The Maxwell-Faraday Equation

8.8 The Maxwell-Faraday Equation

🧭 Overview

🧠 One-sentence thesis

The Maxwell-Faraday Equation generalizes Kirchhoff's Voltage Law to time-varying magnetic fields, revealing that electric and magnetic fields are coupled not only in integral form but also at every point in space.

📌 Key points (3–5)

  • What it generalizes: KVL (Kirchhoff's Voltage Law) works only when magnetic flux is constant; the Maxwell-Faraday Equation extends it to time-varying magnetic flux.
  • Integral form: the electric potential around a closed path equals the negative time rate of change of magnetic flux through that path (Faraday's Law written more explicitly).
  • Differential form: the curl of the electric field at any point equals the negative time rate of change of the magnetic field at that point.
  • Common confusion: the integral form looks like just a restatement of Faraday's Law, but the differential form reveals the deeper coupling—spatial changes in electric field are linked to temporal changes in magnetic field at each point.
  • Why it matters: electric and magnetic fields are coupled everywhere in space, not just in aggregate over loops and surfaces.

🔄 From KVL to the Maxwell-Faraday Equation

⚡ Kirchhoff's Voltage Law (KVL) in electrostatics

KVL states that in the absence of a time-varying magnetic flux, the electric potential accumulated by traversing a closed path C is zero.

  • Mathematically: the line integral of the electric field around a closed path C equals zero.
  • This holds only when magnetic flux is constant (magnetostatic case).
  • KVL is a principle of electrostatics encountered in earlier sections.

🧲 Faraday's Law and the conflict

  • Faraday's Law says: the voltage V around a closed path equals the negative time derivative of the magnetic flux through any open surface S bounded by that path.
  • The flux is the surface integral of the magnetic field B over S.
  • The relative orientations of the path C and the surface element follow the Stokes' Theorem convention.
  • Agreement with KVL: if magnetic flux is constant, Faraday's Law gives V = 0, matching KVL.
  • Conflict: if magnetic flux is time-varying, Faraday's Law gives a nonzero V, which contradicts KVL.

🔗 The resolution: setting them equal

  • The correction is to equate the two expressions: the line integral of E around C equals the negative time derivative of the flux of B through S.
  • This combined expression is the Maxwell-Faraday Equation (MFE) in integral form.
  • It states that the electric potential around a closed path is entirely due to electromagnetic induction via Faraday's Law.

📐 The integral form

📐 What it says

The integral form of the Maxwell-Faraday Equation states that the electric potential associated with a closed path C is due entirely to electromagnetic induction, via Faraday's Law.

  • Left side: line integral of the electric field E around the closed path C.
  • Right side: negative time derivative of the surface integral of the magnetic field B over any open surface S bounded by C.
  • The excerpt notes this might seem like just a verbose restatement of Faraday's Law—which is true.

🤔 Why it seems redundant

  • At first glance, the integral form appears to be Faraday's Law written in a slightly different way.
  • The excerpt acknowledges: "one might argue that all we have done is simply to write Faraday's Law in a slightly more verbose way. This is true."
  • The real power comes from the differential form.

🔬 The differential form

🔬 Deriving it with Stokes' Theorem

  • Apply Stokes' Theorem to the left-hand side (line integral of E around C) to convert it into a surface integral of the curl of E over S.
  • Exchange the order of integration and differentiation on the right-hand side.
  • Now both sides are surface integrals over the same surface S.
  • The surface S is arbitrary: it can be any mathematically-valid open surface anywhere in space, of any size and orientation.
  • For the equation to hold universally, the integrands on both sides must be equal at every point in space.

⚙️ The result: curl of E equals negative time derivative of B

The differential form of the Maxwell-Faraday Equation relates the change in the electric field with position to the change in the magnetic field with time.

  • Mathematically: the curl of the electric field E equals the negative partial time derivative of the magnetic field B.
  • What curl means: the curl of E is a spatial derivative of E (a measure of how E changes with position).
  • What the equation constrains: spatial derivatives of E are simply related to the rate of change of B.

💡 Why this is new and useful

  • The differential form reveals that electric and magnetic fields are coupled at each point in space, not just in aggregate over loops and surfaces.
  • This is "arguably new and useful information" compared to the integral form.
  • It shows a local, point-by-point relationship between the two fields.

🧩 Key distinctions

🧩 Integral vs differential form

FormWhat it describesKey insight
IntegralElectric potential around a closed path due to changing magnetic flux through a surfaceGeneralizes KVL to time-varying fields; looks like Faraday's Law
DifferentialCurl of E at a point equals negative time rate of change of B at that pointElectric and magnetic fields are coupled locally at every point in space

🧩 Magnetostatic vs time-varying

  • Magnetostatic case: magnetic flux is constant, so the time derivative is zero; the Maxwell-Faraday Equation reduces to KVL (line integral of E is zero).
  • Time-varying case: magnetic flux changes with time, so the right-hand side is nonzero; electric potential around a closed path is nonzero due to induction.
  • Don't confuse: KVL is a special case of the Maxwell-Faraday Equation, valid only when fields are static.
104

Displacement Current and Ampere's Law

8.9 Displacement Current and Ampere’s Law

🧭 Overview

🧠 One-sentence thesis

Ampere's Law must be generalized to include displacement current when sources are time-varying, coupling electric and magnetic fields at every point in space.

📌 Key points (3–5)

  • The problem with classical Ampere's Law: it fails for time-varying currents when applied to surfaces that don't intersect physical current (e.g., between capacitor plates).
  • Displacement current: a new term representing the time rate of change of electric flux density, which has units of current but is not conduction current.
  • Common confusion: displacement current is not actual charge flow—it's a mathematical quantity that accounts for changing electric fields and happens to have current units.
  • Coupling of fields: the generalized Ampere's Law shows that spatial changes in magnetic field relate to both conduction current and time changes in electric field.
  • Why it matters: this generalization is essential for Maxwell's Equations and explains electromagnetic behavior in AC circuits and capacitors.

⚠️ The failure of classical Ampere's Law

⚠️ What classical Ampere's Law says

Classical Ampere's Law: the line integral of magnetic field intensity H around a closed path C equals the enclosed current I_encl.

  • Written as: the line integral of H dot dl around C equals I_encl
  • This works fine for steady (DC) currents
  • The law relies on Stokes' Theorem: the line integral can be converted to a surface integral over any open surface S bounded by C

🔌 The capacitor problem

The excerpt demonstrates the failure using a parallel-plate capacitor scenario:

  • Setup: AC current flows in wires connected to a capacitor
  • Surface S1: passes through the wire → intersects current I → Ampere's Law gives correct result
  • Surface S2: passes between capacitor plates → intersects zero current → Ampere's Law gives zero even though magnetic field H exists

Key insight: No current flows directly between capacitor plates, yet current exists in the wires and creates a magnetic field. The choice of surface S shouldn't matter mathematically, but classical Ampere's Law gives different answers depending on which surface you choose.

🧩 Why it works for continuous wires

In a simple wire scenario (no capacitor):

  • Any valid surface bounded by path C will intersect the same current I
  • The choice of surface doesn't matter
  • Classical Ampere's Law works even when current is time-varying

Don't confuse: The problem isn't time-variation itself—it's time-variation combined with geometries where current doesn't physically cross certain surfaces.

🔧 The solution: displacement current

🔧 What displacement current is

The excerpt proposes adding a new term I_d to Ampere's Law:

  • New form: line integral of H dot dl around C equals I_c plus I_d
  • I_c is the enclosed conduction current (the original term)
  • I_d is displacement current (the new term)

Definition of displacement current:

I_d equals the surface integral of (partial derivative of D with respect to time) dot ds, where D is electric flux density and S is the open surface bounded by C.

📋 Properties of displacement current

What we know about I_d from the excerpt:

PropertyDescription
UnitsAmperes (A), same as current
DC behaviorZero when fields are constant
AC behaviorNon-zero when fields vary with time
Physical meaningTime derivative of some quantity
Field dependenceRelated to electric field E (through D = ε E)

🚫 What displacement current is NOT

The excerpt emphasizes an important clarification:

  • Not conduction current: no actual charge flows through the capacitor
  • Misleading name: "displacement current" suggests charge movement, but that's not what happens
  • More general interpretation: represents time variation in charge distribution relative to surface S, regardless of the path charges take

Don't confuse: Displacement current with actual moving charges. It's a mathematical quantity that accounts for changing electric fields and happens to have current units.

🔗 Why it must involve electric field

The excerpt provides reasoning:

  • The Maxwell-Faraday Equation shows that spatial derivatives of E relate to time derivatives of H
  • E and H are coupled in time-varying cases
  • Since Ampere's Law involves H but hadn't yet included E, the missing term must depend on electric field
  • This coupling between E and H must work in both directions

📐 The generalized Ampere's Law

📐 Integral form

The complete time-varying form:

  • Line integral of H dot dl around C equals I_c plus the surface integral of (partial derivative of D with respect to time) dot ds
  • Where I_c is the surface integral of current density J dot ds
  • Both integrals are over the same surface S bounded by path C

Example: For the capacitor problem, surface S2 between the plates has zero conduction current but non-zero displacement current because the electric field between plates changes with time in AC operation.

📐 Differential (point) form

The excerpt derives the differential form using Stokes' Theorem:

  • Curl of H equals J plus partial derivative of D with respect to time
  • This is the form that "unleashes most of the utility" according to the excerpt

What it means in plain language:

The differential form relates the change in magnetic field with position to the change in electric field with time, plus current.

🔄 Coupling at every point

Key conclusion from the excerpt:

  • Spatial derivatives of H (the curl) are constrained by both J and the time derivative of D
  • Electric and magnetic fields become coupled at each point in space when sources are time-varying
  • This is parallel to the Maxwell-Faraday Equation, which couples fields in the opposite direction

🎯 Physical interpretation

🎯 The capacitor case resolved

How displacement current solves the original problem:

  • Surface S1 (through wire): intersects conduction current I_c; displacement current I_d is negligible
  • Surface S2 (between plates): zero conduction current, but displacement current I_d is non-zero because electric field between plates varies with time
  • Both surfaces now give the same total current (I_c + I_d), resolving the inconsistency

🎯 Charge redistribution perspective

The excerpt offers an alternative interpretation:

  • At one instant, charge is distributed one way; at another instant, it's distributed differently
  • If you define current as "time variation in charge distribution relative to S, regardless of path," then I_d qualifies as current
  • This is somewhat philosophical—safer to think of displacement current as a separate electromagnetic quantity with current units

⚡ When each term dominates

ScenarioI_c (conduction)I_d (displacement)
DC in wireNon-zeroZero (no time variation)
AC in wireDominatesSmall
Between capacitor platesZeroNon-zero in AC
Empty space with changing E fieldZeroNon-zero

Don't confuse: The two terms represent fundamentally different physics—one is actual charge flow, the other is changing electric field configuration.

105

Maxwell's Equations in Differential Phasor Form

9.1 Maxwell’s Equations in Differential Phasor Form

🧭 Overview

🧠 One-sentence thesis

Converting Maxwell's Equations from time-domain to phasor form replaces time derivatives with multiplication by jω, dramatically simplifying electromagnetic analysis while preserving all information.

📌 Key points (3–5)

  • Core transformation: time-domain Maxwell's Equations are converted to phasor form by representing fields as phasors (complex exponentials with time dependence e^(jωt)).
  • Key simplification: time derivative operators (∂/∂t) are replaced by multiplication by jω, reducing partial differential equations to spatial derivatives only.
  • What stays the same: Gauss' Law retains identical form in phasor representation; divergence operations are unaffected.
  • What changes: Maxwell-Faraday Equation and Ampere's Law gain jω terms where time derivatives appeared; curl operations remain but act on phasor quantities.
  • Why it matters: without this simplification, much of "basic" engineering electromagnetics would be intractable; no information is lost because Fourier analysis guarantees completeness.

🔄 The phasor transformation method

🔄 What phasor representation means

Phasor quantities: complex-valued representations of time-varying fields through the relationship: physical field = Re{phasor × e^(jωt)}.

  • Any time-domain field (e.g., D, E, B, H) is written as the real part of a phasor multiplied by e^(jωt).
  • Example: D = Re{D̃ e^(jωt)}, where D̃ is the phasor (complex amplitude).
  • The phasor D̃ is constant with respect to time; all time variation is captured in e^(jωt).

🔧 How the conversion works

The excerpt demonstrates a systematic procedure applied to each Maxwell equation:

  1. Substitute phasor representations for all time-domain quantities.
  2. Exchange order of operations: because divergence/curl and "Re{}" are real-valued linear operators, they can be swapped (Claim 2 from Section 1.5).
  3. Separate spatial and temporal operations: spatial derivatives (∇·, ∇×) don't affect e^(jωt); pull time dependence outside spatial operators.
  4. Simplify time derivatives: ∂/∂t applied to e^(jωt) yields jω e^(jωt).
  5. Drop common e^(jωt) factor: equality of phasors follows from equality of time-domain expressions (Claim 1 from Section 1.5).

📐 Gauss' Law transformation

📐 Starting equation

Time-domain Gauss' Law:

  • ∇ · D = ρ_v (Equation 9.1)

📐 Step-by-step conversion

  • Substitute D = Re{D̃ e^(jωt)} and ρ_v = Re{ρ̃_v e^(jωt)}.
  • Result: ∇ · [Re{D̃ e^(jωt)}] = Re{ρ̃_v e^(jωt)}.
  • Swap "Re" and "∇·": Re{∇ · [D̃ e^(jωt)]} = Re{ρ̃_v e^(jωt)}.
  • Since divergence is with respect to position (not time), pull e^(jωt) outside: Re{[∇ · D̃] e^(jωt)} = Re{ρ̃_v e^(jωt)}.
  • Equate phasors: ∇ · D̃ = ρ̃_v (Equation 9.10).

📐 Key observation

The phasor form of Gauss' Law is identical to the time-domain form—only the tilde notation changes.

  • No jω term appears because divergence involves only spatial derivatives.
  • Don't confuse: this simplicity is specific to Gauss' Law; other equations gain jω terms.

⚡ Maxwell-Faraday Equation transformation

⚡ Starting equation

Time-domain Maxwell-Faraday Equation (MFE):

  • ∇ × E = −∂B/∂t (Equation 9.2)

⚡ Step-by-step conversion

  • Substitute E = Re{Ẽ e^(jωt)} and B = Re{B̃ e^(jωt)}.
  • Result: ∇ × [Re{Ẽ e^(jωt)}] = −∂/∂t [Re{B̃ e^(jωt)}].
  • Swap "Re" with curl and time-derivative: Re{∇ × [Ẽ e^(jωt)]} = −Re{∂/∂t [B̃ e^(jωt)]}.
  • Left side: e^(jωt) doesn't depend on position, so Re{[∇ × Ẽ] e^(jωt)}.
  • Right side: B̃ is constant with respect to time (it's a phasor), so ∂/∂t acts only on e^(jωt):
    • −Re{B̃ ∂/∂t e^(jωt)} = −Re{B̃ jω e^(jωt)} = Re{[−jω B̃] e^(jωt)}.
  • Equate phasors: ∇ × Ẽ = −jω B̃ (Equation 9.17).

⚡ The crucial simplification

Time derivative operator ∂/∂t is replaced by multiplication by jω.

  • This transforms a partial differential equation (spatial + temporal derivatives) into one involving only spatial derivatives.
  • Example: instead of computing ∂B/∂t at every instant, multiply the phasor B̃ by jω.
  • The excerpt emphasizes: "This is a tremendous simplification since the equations now involve differentiation over position only."

⚡ Why no information is lost

  • Fourier analysis guarantees that any time-varying signal can be decomposed into sinusoidal components.
  • Solving for each frequency ω separately (via phasors) and superposing results recovers the full time-domain solution.
  • The excerpt notes: "no information is lost in this simplification."

🧲 Remaining equations

🧲 Gauss' Law for Magnetism

Time-domain: ∇ · B = 0 (Equation 9.3)

Phasor form: ∇ · B̃ = 0 (Equation 9.18)

  • Identical structure to time-domain (no jω term).
  • Same reasoning as Gauss' Law: divergence is purely spatial.

🧲 Ampere's Law

Time-domain: ∇ × H = J + ∂D/∂t (Equation 9.4)

Phasor form: ∇ × H̃ = J̃ + jω D̃ (Equation 9.19)

  • Time derivative ∂D/∂t becomes jω D̃.
  • Similar transformation to Maxwell-Faraday Equation.
  • The excerpt leaves detailed derivation "as an exercise for the reader."

📋 Summary of all four phasor equations

EquationTime-domain formPhasor formKey change
Gauss' Law∇ · D = ρ_v∇ · D̃ = ρ̃_vNone (identical structure)
Maxwell-Faraday∇ × E = −∂B/∂t∇ × Ẽ = −jω B̃∂/∂t → jω
Gauss' Magnetism∇ · B = 0∇ · B̃ = 0None (identical structure)
Ampere's Law∇ × H = J + ∂D/∂t∇ × H̃ = J̃ + jω D̃∂/∂t → jω

📋 Pattern recognition

  • Divergence equations (Gauss' Laws): no time derivatives in time-domain → no jω in phasor form.
  • Curl equations (MFE, Ampere): time derivatives in time-domain → jω multiplication in phasor form.
  • All equations now involve only spatial differentiation (∇·, ∇×) and algebraic operations (multiplication by jω).

🎯 Why this matters

🎯 Practical impact

The excerpt states: "Without this kind of simplification, much of what is now considered 'basic' engineering electromagnetics would be intractable."

  • Solving coupled partial differential equations in time and space is extremely difficult.
  • Phasor form reduces the problem to spatial derivatives plus algebra.
  • Engineers can analyze electromagnetic systems frequency-by-frequency, then superpose results.

🎯 What to remember

  • Phasor form is not an approximation—it is an exact reformulation for sinusoidal steady-state analysis.
  • The transformation relies on linearity of Maxwell's Equations and operators (∇·, ∇×, ∂/∂t, Re{}).
  • Don't confuse: phasors assume sinusoidal time dependence at a single frequency ω; for arbitrary time signals, use Fourier superposition.
106

Wave Equations for Source-Free and Lossless Regions

9.2 Wave Equations for Source-Free and Lossless Regions

🧭 Overview

🧠 One-sentence thesis

Maxwell's Equations can be simplified into wave equations for electric and magnetic fields when the region is source-free and lossless, revealing that the fields propagate as waves governed by the same differential equation and differ only by a constant impedance factor.

📌 Key points (3–5)

  • Source-free simplification: removing charges (rho_v = 0) and currents (J = 0) from Maxwell's Equations eliminates complexity and makes wave equations tractable.
  • Lossless requirement: source-free regions automatically exclude conduction current, meaning conductivity sigma must be zero, so no power dissipates as the wave travels.
  • Wave equations for E and H: both electric field E-tilde and magnetic field H-tilde satisfy the same homogeneous differential equation with parameter beta squared.
  • Common confusion: "source-free" does not just mean "no external sources"; it also implicitly means "lossless" because requiring J = 0 forces sigma = 0.
  • Wave impedance preview: because E-tilde and H-tilde solve the same equation, they differ only by a constant factor (an impedance in ohms), analogous to characteristic impedance in transmission lines.

🔧 Simplifying Maxwell's Equations

🔧 Phasor form and time differentiation

  • The excerpt starts from the differential "point" phasor form of Maxwell's Equations developed in Section 9.1.
  • Advantage of phasors: differentiation with respect to time is replaced by multiplication by j omega (imaginary unit times angular frequency).
  • This replacement makes the equations algebraic rather than differential in time, greatly simplifying analysis.

🚫 Source-free conditions

Source-free region: a region containing no net charge and no current, so rho_v = 0 and J = 0.

  • Applying these conditions to Maxwell's Equations removes the source terms on the right-hand sides.
  • The four equations become:
    • Divergence of D-tilde = 0
    • Curl of E-tilde = negative j omega B-tilde
    • Divergence of B-tilde = 0
    • Curl of H-tilde = positive j omega D-tilde

🧲 Choosing E and H over D and B

  • For homogeneous, isotropic, and linear media, D-tilde = epsilon E-tilde and B-tilde = mu H-tilde, where epsilon and mu are real-valued constants.
  • Because D and B are simply scaled versions of E and H, it is sufficient to work with either pair.
  • Engineering convention: use E-tilde and H-tilde.
  • After substitution, the equations involve only E-tilde, H-tilde, epsilon, mu, and omega.

🔒 Why source-free means lossless

🔒 Definition of loss

Loss: reduction in the magnitude of electric and magnetic fields with increasing distance, due to power dissipation in the medium.

  • Loss occurs when conductivity sigma is greater than zero.
  • Ohm's Law for Electromagnetics (J-tilde = sigma E-tilde) requires that power in the electric field be transferred into conduction current, which is then lost to the wave.

🔒 The implicit constraint

  • By requiring J = 0 in the source-free condition, we preclude any conduction current.
  • This implicitly specifies sigma = 0, meaning the medium must be lossless.
  • Evidence: the final equations (9.28–9.31) contain mu and epsilon but not sigma.
  • Don't confuse: "source-free" is not just about external sources; it also rules out internal loss mechanisms.

🔒 Complex permittivity (advanced note)

  • The excerpt mentions that there is a way to handle lossy regions using these equations by redefining epsilon as a complex-valued quantity.
  • The imaginary part of complex epsilon represents loss.
  • This technique is not covered in this section, but readers should be aware that complex permittivity indicates loss.

🌊 Deriving the wave equations

🌊 Starting with Faraday's Law

  • Begin with the curl of E-tilde = negative j omega mu H-tilde (Equation 9.29).
  • Take the curl of both sides:
    • Left side: curl of (curl of E-tilde)
    • Right side: negative j omega mu times (curl of H-tilde)

🌊 Substituting Ampère's Law

  • On the right side, eliminate curl of H-tilde using Equation 9.31 (curl of H-tilde = positive j omega epsilon E-tilde).
  • This gives: negative j omega mu times (positive j omega epsilon E-tilde) = positive omega squared mu epsilon E-tilde.
  • Note: j squared = negative 1, so the product of two j's yields a positive real term.

🌊 Vector identity for curl of curl

  • Apply the vector identity: curl of curl of A = gradient of (divergence of A) minus Laplacian of A.
  • For E-tilde: curl of curl of E-tilde = gradient of (divergence of E-tilde) minus Laplacian of E-tilde.
  • Use Equation 9.28 (divergence of E-tilde = 0) to eliminate the first term.
  • Result: curl of curl of E-tilde = negative Laplacian of E-tilde.

🌊 Final wave equation for E

  • Substitute back into the original equation and rearrange:
    • Laplacian of E-tilde plus omega squared mu epsilon E-tilde = 0.
  • This is a homogeneous differential equation, as expected for a source-free region.

🌊 Defining beta

Beta: the phase propagation constant, defined as omega times the square root of mu epsilon.

  • Units: 1/m (or rad/m).
  • Purpose: beta captures the combined contribution of frequency, permittivity, and permeability in one constant.
  • It indicates the rate at which the phase of the propagating wave progresses with distance.
  • Connection to transmission lines: beta is analogous to the phase propagation constant in transmission line theory (Section 3.8).

🌊 Compact form and duality

  • Using beta, the wave equation for E-tilde becomes: Laplacian of E-tilde plus beta squared E-tilde = 0 (Equation 9.39).
  • The wave equation for H-tilde is derived using essentially the same procedure (left as an exercise).
  • By duality in Equations 9.28–9.31, the result is: Laplacian of H-tilde plus beta squared H-tilde = 0 (Equation 9.40).

🔗 Relationship between E and H

🔗 Same differential equation

  • Both E-tilde and H-tilde satisfy the same homogeneous differential equation.
  • Implication: E-tilde and H-tilde cannot differ by more than a constant factor and a direction.

🔗 Wave impedance

  • Examining units: E-tilde has units of V/m, H-tilde has units of A/m.
  • The ratio of V/m to A/m is ohms (Ω), indicating the constant factor is an impedance.
  • This factor is known as the wave impedance (addressed in Section 9.5).
  • Analogy: wave impedance is analogous to the characteristic impedance of a transmission line (Section 3.7).

🔗 Summary of wave equations

FieldWave EquationNotes
E-tildeLaplacian of E-tilde plus beta squared E-tilde = 0Homogeneous, for isotropic, homogeneous, lossless, source-free material
H-tildeLaplacian of H-tilde plus beta squared H-tilde = 0Same form as E-tilde; solutions differ by wave impedance
  • Both equations apply only to regions that are isotropic, homogeneous, lossless, and source-free.
  • The parameter beta encodes the propagation characteristics of the medium.
107

Types of Waves

9.3 Types of Waves

🧭 Overview

🧠 One-sentence thesis

Electromagnetic waves can be classified by the geometry of their phasefronts—spherical, cylindrical, or plane—with plane waves being especially important because all waves appear locally planar when observed over a small region far from the source.

📌 Key points (3–5)

  • What defines wave types: the shape formed by surfaces of constant phase (phasefronts).
  • Three geometries: spherical (concentric spheres), cylindrical (concentric cylinders), and plane (parallel planes).
  • When each geometry applies: spherical for small sources, cylindrical for line-shaped sources, plane for reflector structures or local observation.
  • Common confusion: the "locally planar" approximation—spherical waves look planar when viewed over a small region far from the source, just as Earth looks flat to an observer on the ground.
  • Why plane waves matter: they simplify analysis and are broadly applicable due to the locally planar approximation.

🌊 What phasefronts are and why they matter

🌊 Definition of phasefronts

Phasefronts: surfaces of constant phase.

  • Wave types are defined by the shape of these surfaces, not by field magnitude.
  • The magnitude of the field on a phasefront may vary significantly, but the geometry of the phasefront determines the wave type.
  • Analogy: electromagnetic waves behave like sound waves, which also exhibit these geometries.

🔍 Why geometry matters

  • Different phasefront geometries correspond to different source shapes and observation scales.
  • The geometry determines how the wave spreads and how analysis can be simplified.

🔵 Spherical waves

🔵 Geometry and characteristics

  • Phasefronts form concentric spheres (Figure 9.1 in the excerpt).
  • The wave radiates outward from a central point.

📏 When to use the spherical model

  • Waves are well-modeled as spherical when the source dimensions are small relative to the observation scale.
  • Example: an antenna with dimensions of 10 cm, observed in free space over a scale of 10 km, produces phasefronts that are very nearly spherical.

🔶 Cylindrical waves

🔶 Geometry and characteristics

  • Phasefronts form concentric cylinders (Figure 9.2 in the excerpt).
  • In one dimension the phasefronts are circular; in the perpendicular direction they are planar.

📏 When to use the cylindrical model

  • A cylindrical wave is often a good description of the wave emerging from a line-shaped source.

🟦 Plane waves

🟦 Geometry and characteristics

  • Phasefronts are planar, forming parallel planes (Figure 9.3 in the excerpt).
  • This is the simplest geometry for analysis.

📏 Two conditions for plane wave models

ConditionDescriptionExample
Structures that produce planar phasefrontsSome structures give rise to waves with planar phasefronts over a limited areaWave radiated by a parabolic reflector (Figure 9.4)
Locally planar approximationAll waves are well-modeled as plane waves when observed over a small region sufficiently far from the sourceSpherical waves appear planar over a small portion of the spherical phasefront (Figure 9.5)

🌍 The "locally planar" approximation

  • Key insight: spherical waves are "locally planar" when observed over a small portion of the spherical phasefront.
  • Analogy: the Earth seems "locally flat" to an observer on the ground, even though it is clearly spherical to an observer in orbit.
  • Don't confuse: a wave can be globally spherical but locally planar—the approximation depends on the observation scale.

⚙️ Why plane waves are especially important

  • The "locally planar" approximation is broadly applicable and simplifies analysis.
  • Most waves can be treated as plane waves under the right observation conditions.
  • The excerpt emphasizes: "Plane waves (having planar phasefronts) are of particular importance due to wide applicability of the 'locally planar' approximation."

📋 Summary comparison

Wave typePhasefront shapeTypical sourceWhen to use
SphericalConcentric spheresSmall source relative to observation scaleSource dimensions ≪ observation distance
CylindricalConcentric cylinders (circular in one dimension, planar in perpendicular)Line-shaped sourceLine sources
PlaneParallel planesParabolic reflector or any source viewed locallyLimited area in front of reflector; small region far from any source
108

Uniform Plane Waves: Derivation

9.4 Uniform Plane Waves: Derivation

🧭 Overview

🧠 One-sentence thesis

Uniform plane waves—where electric and magnetic fields have constant magnitude and phase over parallel planes—are solutions to Maxwell's wave equations that propagate perpendicular to those planes, with the electric field, magnetic field, and direction of propagation mutually perpendicular and related by the wave impedance of the medium.

📌 Key points (3–5)

  • What uniform plane waves are: waves for which both electric field (E) and magnetic field (H) have constant magnitude and phase in a specified plane (e.g., planes of constant z).
  • Why they matter: despite being a special case, they are broadly applicable as building blocks in unguided propagation, transmission lines, and waveguides, and all waves appear "locally planar" when observed over small regions far from the source.
  • Key constraint from Maxwell's Equations: if a wave is uniform over a plane, the electric and magnetic field vectors must lie in that plane (the z-components are zero), and propagation occurs perpendicular to the plane.
  • Relationship between E and H: the fields are perpendicular to each other and to the direction of propagation; their magnitudes are related by the wave impedance η = sqrt(μ/ε), which has units of ohms.
  • Common confusion: the fields do not point in the direction of propagation—they lie in the plane perpendicular to propagation; the direction of propagation is given by E × H.

🌊 Types of waves and the locally planar approximation

🌐 Spherical, cylindrical, and plane waves

The excerpt describes three wave types by their phasefront geometry:

Wave typePhasefront shapeExample source
SphericalConcentric spheresPoint source
CylindricalConcentric cylindersLine-shaped source
PlaneParallel planesParabolic reflector (over limited area)

🔍 When plane waves are a good model

Plane waves arise in two conditions:

  1. Structures that produce planar phasefronts: for example, a parabolic reflector antenna creates waves with planar phasefronts over a limited area in front of the reflector.
  2. "Locally planar" approximation: all waves (including spherical and cylindrical) are well-modeled as plane waves when observed over a small region located sufficiently far from the source.

Analogy: The Earth seems "locally flat" to an observer on the ground, even though it is clearly spherical from orbit.

Why it matters: The locally planar approximation is broadly applicable and simplifies analysis, making plane waves particularly important.

🧮 Setting up the uniform plane wave problem

📐 Starting from Maxwell's wave equations

The derivation begins with the phasor-domain wave equations (from Section 9.2):

  • Nabla-squared E-tilde plus beta-squared E-tilde equals zero
  • Nabla-squared H-tilde plus beta-squared H-tilde equals zero

where beta = omega times sqrt(mu times epsilon), assuming unbounded, homogeneous, isotropic, lossless, and source-free media.

Uniform plane wave: a wave for which both E-tilde and H-tilde have constant magnitude and phase in a specified plane.

🎯 Choosing the plane of constant z

The excerpt assumes the plane over which E and H are constant is a plane of constant z (perpendicular to the z-axis).

  • No loss of generality: you could choose a plane of constant x or y and then exchange variables, or rotate coordinates for any other planar orientation.
  • Why this works: the physics does not depend on the orientation of the plane—if it did, the medium would not be isotropic.

🔒 Mathematical constraint

The condition "constant magnitude and phase over planes of constant z" translates to:

  • Partial derivative of E-tilde with respect to x = 0
  • Partial derivative of E-tilde with respect to y = 0
  • Partial derivative of H-tilde with respect to x = 0
  • Partial derivative of H-tilde with respect to y = 0

This means the fields do not vary in the x or y directions; they can only vary in the z direction.

🧩 Solving for the field components

⚡ Electric field components

Breaking E-tilde into Cartesian components (x, y, z), the wave equation becomes three separate equations (one per component).

Applying the constraint that derivatives with respect to x and y are zero, the equations simplify to:

  • Second partial derivative of E_x with respect to z plus beta-squared E_x = 0
  • Second partial derivative of E_y with respect to z plus beta-squared E_y = 0
  • Second partial derivative of E_z with respect to z plus beta-squared E_z = 0

🚫 The z-component must be zero

Using Ampere's Law (curl of H = j omega epsilon E) and taking the dot product with the z-direction unit vector:

  • The left side involves derivatives of H_x and H_y with respect to y and x, which are zero by the constraint.
  • Therefore, E_z must be zero.

The same procedure applied to H (using the Maxwell-Faraday Equation) shows H_z is also zero.

Key conclusion:

If a wave is uniform over a plane, then the electric and magnetic field vectors must lie in this plane.

This is a direct consequence of Maxwell's Equations requiring the electric field to be proportional to the curl of the magnetic field and vice versa.

📊 General solutions for E_x and E_y

The solutions to the simplified wave equations are:

  • E_x = E⁺_x0 times exp(-j beta z) plus E⁻_x0 times exp(+j beta z)
  • E_y = E⁺_y0 times exp(-j beta z) plus E⁻_y0 times exp(+j beta z)

where E⁺_x0, E⁻_x0, E⁺_y0, E⁻_y0 are complex-valued constants determined by boundary conditions or sources outside the region.

Interpretation:

  • Terms with exp(-j beta z) describe propagation in the +z direction.
  • Terms with exp(+j beta z) describe propagation in the -z direction.

These are the same equations encountered in lossless transmission lines.

🧭 Direction of propagation

If a wave is uniform over a plane, then possible directions of propagation are the two directions perpendicular to the plane.

Since the electric and magnetic field vectors lie in the plane, we conclude:

The direction of propagation is perpendicular to the electric and magnetic field vectors.

This conclusion is generally true (not limited to uniform plane waves). Any wave can be interpreted as a linear combination of uniform plane waves, so perpendicular orientation is inescapable.

🧲 Magnetic field components

The same procedure yields the uniform plane wave solution for H:

  • H = x-hat H_x plus y-hat H_y
  • H_x = H⁺_x0 times exp(-j beta z) plus H⁻_x0 times exp(+j beta z)
  • H_y = H⁺_y0 times exp(-j beta z) plus H⁻_y0 times exp(+j beta z)

The solution is essentially the same as for E, but the arbitrary constants may have different values.

🔗 Relating the electric and magnetic fields

🤔 Two considerations

Maxwell's curl equations make clear there must be a relationship between E and H:

  1. Magnitude and phase: E and H are solutions to the same wave equation, so they may differ by no more than a multiplicative constant. Since E has units V/m and H has units A/m, this constant must have units of ohms (an impedance).
  2. Direction: the direction of E must be related to the direction of H.

🧮 Deriving the relationship

Assume the electric field points in the +x direction and propagates in the +z direction:

  • E-tilde = x-hat E_0 exp(-j beta z)

(No loss of generality—any uniform plane wave can be described this way by rotating coordinates.)

Using the Maxwell-Faraday Equation (curl of E = -j omega mu H), solve for H:

  • H-tilde = (curl of E) / (-j omega mu)

Applying the curl operator in Cartesian coordinates:

  • The x-component of H is zero (because E_y and E_z are zero).
  • The y-component involves the partial derivative of E_x with respect to z.
  • The z-component of H is zero (H must be perpendicular to the direction of propagation).

Result:

  • H-tilde = y-hat (beta / omega mu) E_0 exp(-j beta z)

🔄 Perpendicularity of E, H, and propagation direction

H points in the +y direction, so:

  • E and H are both perpendicular to the direction of propagation (+z).
  • E and H are perpendicular to each other.
  • Just as x-hat cross y-hat equals z-hat, E-tilde cross H-tilde points in the direction of propagation.

Summary:

E-tilde, H-tilde, and the direction of propagation (E × H) are mutually perpendicular.

⚡ Wave impedance

The factor relating E and H is beta / (omega mu), which simplifies to:

  • beta / (omega mu) = (omega sqrt(mu epsilon)) / (omega mu) = 1 / sqrt(mu / epsilon)

The factor sqrt(mu / epsilon) has units of ohms and is known as the wave impedance or intrinsic impedance of the medium, denoted η (eta):

Wave impedance: η = sqrt(μ / ε)

The ratio of the electric field intensity to the magnetic field intensity is the wave impedance η (units of Ω). In lossless media, η is determined by the ratio of permeability to permittivity of the medium.

In free space:

  • η₀ = sqrt(μ₀ / ε₀) ≈ 377 Ω

📝 Final solution

If E is as given (pointing in +x, propagating in +z):

  • E-tilde = x-hat E_0 exp(-j beta z)
  • H-tilde = y-hat (E_0 / η) exp(-j beta z)

Don't confuse: The fields do not point in the direction of propagation; they lie in the plane perpendicular to it. The direction of propagation is given by E × H.

109

Uniform Plane Waves: Characteristics

9.5 Uniform Plane Waves: Characteristics

🧭 Overview

🧠 One-sentence thesis

Uniform plane waves in lossless media exhibit predictable characteristics—including wave impedance, wavelength, and phase velocity—that depend only on the material properties and determine how electromagnetic energy propagates through space.

📌 Key points (3–5)

  • Wave impedance: the ratio of electric field intensity to magnetic field intensity, determined by the medium's permeability and permittivity.
  • Phase velocity: the speed at which points of constant phase travel, maximum in free space (speed of light c) and slower in all other materials.
  • Wavelength and frequency relationship: at a given frequency, wavelength is shorter in materials than in free space due to slower phase velocity.
  • Common confusion: phase velocity is not the same in all media—it depends on material properties and is always less than or equal to c.
  • Plane wave relationships: simple equations relate electric and magnetic fields through wave impedance and direction of propagation, applicable even beyond ideal plane waves.

🌊 Wave impedance fundamentals

🔌 What wave impedance means

Wave impedance η: the ratio of electric field intensity to magnetic field intensity (units: Ω).

  • In lossless media, η equals the square root of (permeability / permittivity).
  • Written as: η = √(μ/ε)
  • This is also called the "intrinsic impedance" of the medium.

🌍 Free space impedance

  • Free space has a special wave impedance denoted η₀.
  • Value: η₀ = √(μ₀/ε₀) ≈ 377 Ω
  • This is a fundamental constant for electromagnetic waves in vacuum.

📐 Field structure and relationships

📏 Mutual perpendicularity

  • The electric field, magnetic field, and direction of propagation are mutually perpendicular.
  • The direction of propagation points in the same direction as the cross product E × H.
  • Example: if E points in +x direction and wave travels in +z direction, then H points in +y direction.

🔄 Plane wave relationships

The excerpt provides simple formulas to find one field from the other:

GivenFindFormula
E and direction k̂HH = (1/η) k̂ × E
H and direction k̂EE = -η k̂ × H
  • These work for both actual fields and their phasor representations.
  • They apply at each point in space, even for non-planar or non-uniform waves.
  • Very useful for quickly determining field relationships when propagation direction is known.

⚡ Phase and magnitude

  • Both E and H have the same phase and frequency.
  • Both depend on position and time through the same function: cos(ωt - βz + ψ).
  • Only the phase (not magnitude) varies with position z in lossless media.
  • The fields are real-valued physical quantities; phasors are mathematical representations.

🚀 Phase velocity and propagation speed

🏃 What phase velocity measures

Phase velocity vₚ: the speed at which any point of constant phase appears to travel along the direction of propagation.

  • Calculated as: vₚ = ω/β (angular frequency / phase propagation constant).
  • In lossless media: vₚ = 1/√(με).
  • Depends only on material properties, not on frequency.

🌟 Speed of light as maximum

  • In free space: c = 1/√(μ₀ε₀) ≈ 3.00 × 10⁸ m/s.
  • This is called "speed of light" but applies to electromagnetic fields at any frequency, not just optical.
  • Phase velocity is maximum in free space.
  • In any other material: vₚ is slower by a factor of 1/√(μᵣεᵣ), where subscript r denotes relative values.

📉 Material effects

  • Permittivity ε and permeability μ of any material are greater than vacuum values.
  • Therefore phase velocity in any material is less than c.
  • Example from excerpt: polyethylene with εᵣ ≈ 2.3 has vₚ ≈ 1.98 × 10⁸ m/s, about two-thirds the speed of light.

📏 Wavelength characteristics

🌊 Definition and calculation

Wavelength λ: the distance over which the wave's phase increases by 2π radians.

  • Formula: λ = 2π/β
  • The wave is periodic in space; it repeats every distance λ.
  • At a fixed point in time, the field has the same value each time z changes by λ.

🔗 Relationship to frequency and velocity

  • Key formula: λ = vₚ/f (phase velocity / frequency).
  • Given any two of f, λ, and vₚ, you can solve for the third.
  • Inverse relationship: λ and vₚ are inversely proportional at fixed frequency.

📐 Wavelength in materials vs free space

  • At a given frequency, wavelength in any material is shorter than in free space.
  • This follows from slower phase velocity in materials.
  • Example: at 1 GHz in polyethylene, λ ≈ 19.8 cm, about two-thirds of free-space wavelength.
  • Don't confuse: shorter wavelength doesn't mean less energy—it's a consequence of slower propagation.

🧮 Practical analysis approach

🔍 Working with phasors

  • Phasors (denoted with tilde, e.g., Ẽ) represent physical fields but are not the actual field values.
  • Actual physical field: E = Re{Ẽ exp(jωt)}.
  • Phasors simplify calculations; convert back to real fields for physical interpretation.

🎯 Example workflow (radially-directed wave)

The excerpt demonstrates analysis in cylindrical coordinates:

  1. Identify direction of propagation (k̂ = +ρ̂ in the example).
  2. Write phasor for electric field with appropriate phase constant β.
  3. Use plane wave relationship to find magnetic field: H̃ = (1/η) k̂ × Ẽ.
  4. Verify: check that Ẽ × H̃ points in propagation direction.
  5. Convert phasor to physical field using Re{Ẽ exp(jωt)}.

⚠️ Coordinate system flexibility

  • Plane wave solutions work in any coordinate system.
  • The example uses cylindrical coordinates (ρ, φ, z) for a radially-directed plane wave.
  • The wave remains planar; cylindrical coordinates are just convenient for that geometry.
110

Wave Polarization

9.6 Wave Polarization

🧭 Overview

🧠 One-sentence thesis

Polarization describes how the electric field vector of a wave is oriented in space and time, with linear polarization maintaining a constant direction and circular polarization rotating continuously in the plane perpendicular to propagation.

📌 Key points (3–5)

  • What polarization measures: the orientation of the electric field vector as the wave propagates through space or evolves in time at a fixed point.
  • Linear polarization: the electric field always points in the same direction; common in straight wire antennas like dipoles.
  • Circular polarization: the electric field vector rotates with constant magnitude; useful when transmitter/receiver orientations vary or when the medium rotates the field.
  • Common confusion: left vs right circular polarization—determined by the direction of rotation relative to propagation direction (use left-hand rule for LCP, right-hand for RCP).
  • Why it matters: polarization choice affects system design; circular polarization solves orientation problems in satellite communications and compensates for effects like Faraday rotation.

📐 Linear polarization fundamentals

📐 What linear polarization means

A wave exhibits linear polarization if the direction of the electric field vector does not vary with either time or position.

  • The electric field always points in the same direction as the wave propagates.
  • Example: a wave with electric field in the +x direction remains in the +x direction everywhere along its path.

🔷 Mathematical representation

For a uniform plane wave propagating in the +z direction in lossless media:

  • Single-component form: E with tilde = x-hat E_x exp(−jβz) or y-hat E_y exp(−jβz)
  • General form: E with tilde = ρ-hat E_ρ exp(−jβz), where ρ-hat can be any direction perpendicular to z-hat.
  • In Cartesian coordinates: ρ-hat = x-hat cos φ + y-hat sin φ, so the field can be written as a combination of x and y components with angle φ determining the orientation.

📡 How linear polarization arises

  • From the source: when the source itself is linearly polarized.
  • Common example: straight wire antennas such as dipoles or monopoles radiate linearly polarized waves.
  • Optical frequencies: linear polarization can be created by passing a plane wave through a polarizer.

🔄 Circular polarization fundamentals

🔄 What circular polarization means

A wave exhibits circular polarization if the electric field vector rotates with constant magnitude.

  • The direction of the electric field changes continuously, tracing a circle in the plane perpendicular to propagation.
  • The magnitude of the field remains constant while the direction rotates.

🧮 How circular polarization is created

Circular polarization results from combining two orthogonal linearly polarized waves with a π/2 phase shift:

  • Start with E_x with tilde = x-hat E_x exp(−jβz) and E_y with tilde = y-hat E_y exp(−jβz).
  • If E_x and E_y have the same phase, the result is linear polarization.
  • If E_x and E_y have different phases by π/2, the result is circular polarization.
  • Example: Let E_x = E_0 and E_y = +jE_0 (phase-shifted by +π/2). The physical fields become:
    • E_x component: |E_0| cos(ωt − βz)
    • E_y component: |E_0| cos(ωt − βz + π/2)
  • These are π/2 radians out of phase, so when one reaches maximum, the other is zero, causing the total field vector to rotate.

🤚 Left vs right circular polarization

The direction of rotation distinguishes two types:

TypeDefinitionHand rule
Left circular (LCP)Field rotates in the direction of left-hand fingers when thumb points along propagationPoint left thumb in propagation direction; fingers curl in rotation direction
Right circular (RCP)Field rotates in the direction of right-hand fingers when thumb points along propagationPoint right thumb in propagation direction; fingers curl in rotation direction
  • Mathematical distinction:
    • LCP: E_y = +jE_x (positive phase shift)
    • RCP: E_y = −jE_x (negative phase shift)
  • Don't confuse: the handedness is defined relative to the direction of propagation, not the observer's viewpoint.

📡 Practical applications of circular polarization

Circular polarization is useful when:

  • Variable geometry: relative orientations of transmitter and receiver change over time.
  • Field rotation by medium: the propagation medium can rotate the electric field vector.

Example: GPS satellites

  • GPS satellites transmit circular polarization.
  • Reasons:
    1. Variable geometry of the space-to-earth radio link (satellites move relative to receivers).
    2. Earth's ionosphere causes Faraday rotation (rotates the electric field vector).
  • If GPS used linear polarization, receivers would need to continuously adjust antenna orientation to optimally receive the signal.

🛠️ How to generate circular polarization

  • Antenna pairs: use two perpendicularly-oriented dipoles fed with the same signal but with a 90° phase shift.
  • Intrinsic antennas: use antennas that are inherently circularly polarized, such as helical antennas.

🌀 Elliptical polarization

🌀 What elliptical polarization is

Elliptical polarization results when the x and y components of the electric field do not have equal magnitude.

  • The electric field vector still rotates, but traces an ellipse rather than a circle.
  • Typically not an intended condition; usually observed as degradation in systems designed for linear or circular polarization.

⚠️ When elliptical polarization occurs

  • Most "circularly polarized" antennas produce true circular polarization only in one direction.
  • In all other directions, they produce various degrees of elliptical polarization.
  • Don't confuse: elliptical polarization is a general case; linear and circular are special cases (linear when one component is zero; circular when both components have equal magnitude and π/2 phase difference).
111

Wave Power in a Lossless Medium

9.7 Wave Power in a Lossless Medium

🧭 Overview

🧠 One-sentence thesis

In lossless media, electromagnetic waves carry power that is best quantified as time-average power density (power per unit area) rather than total power, and this density can be calculated from the electric and magnetic field magnitudes using the Poynting vector.

📌 Key points (3–5)

  • Why power density matters: uniform plane waves have infinite total power because fields are constant over infinite planes, so we measure power per unit area (W/m²) instead.
  • The Poynting vector: instantaneous power density is the magnitude of E × H, which points in the direction of propagation.
  • Time-average power density: for sinusoidal waves, the time-average is (magnitude of E₀ squared) divided by (2 times the medium impedance η).
  • Common confusion: the formula uses peak magnitude by default; if given rms values, the factor of 1/2 disappears because peak is √2 times rms.
  • Analogy to circuits: the wave power formula resembles the circuit formula for power across a resistor, with impedance η playing the role of resistance.

🔋 Why power density instead of total power

🔋 The infinite-power problem

  • A uniform plane wave has electric and magnetic fields that are constant across an entire infinite plane perpendicular to propagation.
  • If power passing through any finite area is greater than zero, then integrating over an infinite plane gives infinite total power.
  • In practice, all real plane waves are only "locally planar" (uniform over a limited region), so the infinite-power scenario never actually occurs.

📏 The solution: spatial power density

Spatial power density (or simply "power density"): power per unit area, measured in W/m².

  • This quantity remains finite and meaningful even for uniform plane waves.
  • If you need total power through a finite area, integrate the power density over that area.
  • Don't confuse: "power density" can also mean power spectral density (W/Hz) or power flux density (W/(m²·Hz)) in other contexts; here it strictly means spatial power density (W/m²).

⚡ The Poynting vector and instantaneous power

⚡ Definition and direction

The Poynting vector is defined as:

S = E × H (the cross product of electric field intensity and magnetic field intensity)

  • The magnitude |S| gives the instantaneous power density.
  • The direction of E × H is the direction of propagation, which is the direction power flows.
  • Units check: (V/m) times (A/m) equals (V·A/m²) = W/m², which is correct for power density.

⚡ For a uniform plane wave

  • Starting with an electric field phasor propagating in the +z direction with magnitude E₀ and phase ψ.
  • The associated magnetic field has magnitude (E₀ divided by η), where η = √(μ/ε) is the real-valued impedance of the lossless medium.
  • The instantaneous Poynting vector becomes: S = (magnitude of E₀ squared divided by η) times cosine-squared of (ωt − βz + ψ), pointing in the +z direction.
  • This instantaneous value oscillates rapidly with time and position.

📊 Time-average power density

📊 Why time-average matters

  • Instantaneous power density fluctuates at twice the wave frequency (because of the cosine-squared term).
  • For practical applications, we care about the average power over one period T of the wave.
  • This is calculated by integrating |S| over one period and dividing by T.

📊 The key formula

After integrating the cosine-squared term (which averages to 1/2 over one period):

S_ave = (magnitude of E₀ squared) / (2η)

  • Units: (V/m)² divided by Ω equals W/m², as expected.
  • This is the time-average power density for a sinusoidally-varying uniform plane wave in lossless media.

📊 Analogy to circuit theory

The wave power formula resembles familiar results:

ContextFormulaNotes
Circuit resistorP_ave = (magnitude of Ṽ squared) / (2R)Voltage phasor across resistance R
Transmission lineP_ave = (magnitude of V₀⁺ squared) / (2Z₀)Voltage wave with characteristic impedance Z₀
Plane waveS_ave = (magnitude of E₀ squared) / (2η)Electric field with medium impedance η
  • In all three cases, the impedance (R, Z₀, or η) plays an analogous role.

⚠️ Peak vs RMS magnitudes

⚠️ The common pitfall

  • The formulas above assume E₀ is the peak magnitude of the electric field.
  • However, field quantities are often given as root mean square (rms) values instead.
  • Peak magnitude is √2 times larger than rms magnitude.

⚠️ Adjusting the formula

When using rms values:

  • If you have E₀,rms and use the standard formula, first multiply by √2 to get peak: E₀,peak = √2 · E₀,rms.
  • Alternatively, use the rms-adapted formula: S_ave = (E₀,rms squared) / η (the factor of 1/2 disappears because (√2)² = 2 cancels it).

⚠️ Example scenario

A radio wave in a rural area (approximated as a uniform plane wave) has electric field intensity 10 μV/m rms. Assuming propagation in air (η ≈ 377 Ω):

  • Method 1 (convert to peak): S_ave = (√2 · 10×10⁻⁶)² / (2·377) ≈ 2.65×10⁻¹³ W/m².
  • Method 2 (use rms formula): S_ave = (10×10⁻⁶)² / 377 ≈ 2.65×10⁻¹³ W/m².
  • Both methods yield 0.265 picowatts per square meter.

🧮 Phasor form of the Poynting vector

🧮 Direct calculation from phasors

For convenience when working with phasor representations, there is an alternative Poynting vector that directly yields time-average power:

S_ave = (1/2) times the real part of (Ẽ × H̃)*

  • The tilde denotes phasor form, and the asterisk denotes complex conjugate.
  • This formula operates directly on field phasors without converting back to time-domain.

🧮 Verification for uniform plane waves

  • For a plane wave with electric field phasor Ẽ = x̂ E₀ exp(−jβz) and magnetic field phasor H̃ = ŷ (E₀/η) exp(−jβz):
  • Applying the phasor Poynting formula gives S_ave = ẑ (magnitude of E₀ squared) / (2η), pointing in the propagation direction.
  • This matches the result from time-domain integration, confirming the formula's validity.