Physics 2 Guidebook PDF
Physics 2 Guidebook PDF
Physics 2 Guidebook PDF
c
2007 John and Gay Stewart, The University of Arkansas. Supported in part by the PhysTEC project and the
National Science Foundation.
Contents
1 Electric Charge 2
1.1 Introduction to Electric Charge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Conservation of Charge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Quantization of Charge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3.1 Microscopic Origin of Charge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3.2 Quantization of Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3.3 Determining the Sign of a Charged Object . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Charge in a Macroscopic Object . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4.1 The Structure of Matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.5 What is a Macroscopic Net Charge? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.6 Continuous Charge Distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.6.1 What is a Charge Density? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.6.2 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.6.3 Computing Total Charge from a Density . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2 Electric Force I 10
2.1 Basics of Electric Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.1 Qualitative Exploration of Electric Force . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.2 Properties of the Electric Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2 Effect of Electric Force on Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2.1 Effect of Electric Force on Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2.2 Effect of Electric Force on a Charged Conductor . . . . . . . . . . . . . . . . . . . . . . . 15
2.2.3 Charge Sharing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2.4 Charge Separation on a Conductor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2.5 Polarization of an Insulator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2.6 A Charged Object Attracts an Uncharged Object . . . . . . . . . . . . . . . . . . . . . . 18
2.2.7 Charging by Charge Separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.3 Capacity and Grounding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.1 General Discussion of Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.2 Grounding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.4 Shielding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.4.1 Shielding Effect of a Conductor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.5 Charging by Induction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.5.1 Charging by Induction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.6 Obtaining a Net Charge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.7.1 Summary of the Basic Principles of Electrostatics . . . . . . . . . . . . . . . . . . . . . . 25
2.7.2 Placing the Charge You Want on a Conducting Sphere . . . . . . . . . . . . . . . . . . . 25
1
CONTENTS CONTENTS
3 Electrostatic Devices 28
3.1 Electrophorus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.2 Leaf Electroscope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.2.1 Building and Using a Leaf Electroscope . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.2.2 Demonstrating Charge Separation with a Leaf Electroscope . . . . . . . . . . . . . . . . . 31
3.2.3 Charging a Leaf Electroscope by Induction . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.2.4 Demonstration of Capacity Using a Leaf Electroscope . . . . . . . . . . . . . . . . . . . . 33
3.2.5 Demonstrating Positive and Negative Charge with a Leaf Electroscope . . . . . . . . . . . 34
4 Electric Force II 38
4.1 The Strength of the Electric Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.2 Review of Basic Vector Concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.2.1 Vector Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.2.2 Manipulating Vectors, Unit Vectors, and Magnitudes . . . . . . . . . . . . . . . . . . . . 40
4.2.3 Position and Displacement Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.3 Coulomb’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.3.1 Qualitative Features of Coulomb’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5 Electric Field 48
5.1 Definition of Electric Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.2 Coulomb’s Law for Electric Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5.2.1 Coulomb’s Law for the Electric Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5.2.2 Fields of More than One Point Charge . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
5.3 Arrow Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.3.1 Representing Electric Fields: Arrow Diagrams . . . . . . . . . . . . . . . . . . . . . . . . 52
5.3.2 Field and Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.4 Electric Fields of Lines and Planes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.5 General Vector Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.6 Electric Field of Many Objects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
6 Mechanics 63
6.1 Mechanics Problems Involving the Electric Force . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.1.1 Review of Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.1.2 Mechanics of Systems Involving Tension Forces . . . . . . . . . . . . . . . . . . . . . . . 64
6.2 Mechanics in an Electric Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
8 Electric Dipoles 81
8.1 Behavior of Electric Dipoles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
8.2 Drawing Dipole Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
8.3 Qualitative Dipole Behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
8.4 Dipole Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
8.4.1 Potential Energy of a Dipole in an Electric Field . . . . . . . . . . . . . . . . . . . . . . . 89
8.4.2 Torque on an Electric Dipole . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
8.4.3 Force on an Electric Dipole in a Non-Uniform Field . . . . . . . . . . . . . . . . . . . . . 90
9 Continuous Charges 93
9.1 Thinking About Continuous Systems of Charge . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
9.2 One Dimensional Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
10 Symmetry 98
10.1 High Symmetry Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
10.2 The Shape of a High Symmetry Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
10.3 Drawing Spherical and Cylindrical Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
10.4 Drawing Planar Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
19 Capacitance 187
19.1 Definition of Capacitance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
19.1.1 Definition of Capacitance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
19.1.2 Capacitance and Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
19.1.3 Effect of Dielectrics on Capacitance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
19.2 Capacitance from Electrostatics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
19.2.1 Computing the Capacitance of a Two Conductor System . . . . . . . . . . . . . . . . . . 191
19.2.2 Calculating the Capacitance of One Conductor and Ground . . . . . . . . . . . . . . . . . 199
21 DC Circuits 211
21.1 Ohm’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
21.1.1 Ohm’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
21.1.2 Physical Batteries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
21.2 Resistor Networks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
21.2.1 Identifying Series and Parallel Combinations of Resistors . . . . . . . . . . . . . . . . . . 213
21.2.2 Reducing Resistor Networks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
21.3 Kirchhoff’s Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
21.3.1 Kirchhoff’s Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
21.3.2 Navigating Multi-Loop Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
21.3.3 Analyzing Kirchhoff’s Law Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
23 RC Circuits 238
23.1 Reasoning About RC Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
23.1.1 General Behavior of RC Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
23.1.2 Short and Long Time Behavior of RC Circuits . . . . . . . . . . . . . . . . . . . . . . . . 238
23.2 RC Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
23.2.1 General Behavior of RC Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
23.2.2 Exponential Time Dependence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
23.2.3 Discharging RC Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
23.2.4 Charging RC Circuit Behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
23.3 Graphing RC Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
32 Inductance 349
32.1 Inductance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349
32.2 Inductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
32.3 Mutual Inductance and Transformers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 354
34 RL Circuits 360
34.1 Limiting Behavior of RL Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
34.2 RL Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
38 Light 395
38.1 Light and Matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
38.1.1 Index of Refraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
38.1.2 Wave Properties of Light in Matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 396
38.2 Reflection at a Plane Interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 397
38.2.1 Light Rays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 397
38.2.2 Reflection at Plane Interfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398
38.3 Snell’s Law - Refraction at Plane Interfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398
38.3.1 Refraction at Planes Interfaces - Snell’s Law . . . . . . . . . . . . . . . . . . . . . . . . . 398
38.3.2 Critical Angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
38.4 Transmission and Reflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
38.4.1 Intensity of Transmission and Reflection . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
38.4.2 Polarization by Reflection - Brewster’s Law . . . . . . . . . . . . . . . . . . . . . . . . . 404
43 Appendix 474
43.1 Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 474
43.2 Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475
43.3 Presenting your Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475
44 Index 477
Electric Charge
We are all familiar with the basic feature of matter called mass, which determines how strongly the gravitational
force pulls on an object. A second fundamental feature of all matter is electric charge. This course deals with
the effects of stationary and moving electric charges.
9
1.2. CONSERVATION OF CHARGE CHAPTER 1. ELECTRIC CHARGE
Symbols for Small Quantities of Charge: We often give you charges in milli-
Coulombs (mC = 1 × 10−3 C), micro-Coulombs (µC = 1 × 10−6 C), or nano-Coulombs
(nC = 1 × 10−9 C).
Law of Conservation
P of Charge: In all physical processes, the total
P charge before the
process, Qinitial , is equal to the total charge after the process, Qf inal .
X X
Qinitial = Qf inal
This law does not mean that the number of protons and electrons in the universe remains the same. Nuclear
processes change the number of protons, neutrons, and electrons. For example, your smoke detector turns a
neutron (charge 0) into an electron (charge -e), a proton (charge +e), and an antineutrino (charge 0). So the
number of charged particles changes but the total charge of the universe is the same, 0 = −e + e + 0 = 0.
In lab, and in your own personal experience, charge sometimes SEEMS to disappear. In these cases, it has
actually escaped into the atmosphere, the earth, or your body. We will call everywhere that charge escapes to
the environment.
Exchange of Charge with the Environment: If we are moving charge from place
to place or just letting a charged object sit around, some charge may be lost to the
environment.
Later on, we will be able to use a battery or a charged rod to draw charge out of the environment. When
analyzing a charge conservation process, missing charge or extra charge comes from the environment: the Earth,
water vapor in the air, etc.
Solution
As the containers are symmetric, the charge is equally shared between the two containers.
7C -5C 1C 1C
Solution
If the charge Q is made up of an integer number of charged particles whose charge is ±e, then Q = ±N e where
N is the number of electrons, Q is the total charge, and e = 1.602 × 10−19 C.
1.0 × 10−6 C
Q
N = = = 6.212
e 1.602 × 10−19 C
The law of quantization of charge requires that N is an integer, but that only affects the 12th significant figure
in the number above. Electrons have a negative charge. An object is deficient in electrons if the object’s charge
is positive.
N = 6.2 × 1012 deficient electrons
Definition of Element: An element, like Hydrogen or Helium, is all atoms with the
same number of protons in their nucleus.
The number of neutrons in the nucleus can vary somewhat from atom to atom for different atoms of the same
element. An atom with fixed number of protons and neutrons is called an nuclide. The collection of nuclides
belonging to the same element are the isotopes of the element. A nuclide is characterized by the atomic number
Z and the mass number, A. The mass number is the total number of protons and neutrons. For example, a
nuclide of carbon, one of the isotopes of carbon, with 6 protons (like all carbon atoms) and 7 neutrons would
have atomic number 6 and mass number 13 = 6 + 7. This nuclide is represented by the symbol 13 6 C. This nuclide
is called Carbon-13 as opposed to its more common relative with 6 neutrons, 126 C, Carbon-12.
Atomic Number: The atomic number, Z, of an atom is the number of protons in the
atom. Since atoms are neutral it is also the number of electrons. The atomic number
of Carbon is 6, so there are 6 protons and 6 electrons in an atom of carbon.
Mass Number: The mass number, A, is the total number of protons and neutrons in
an atom.
Any macroscopic object contains an enormous number of atoms; a number that is too large to conveniently
work with. Instead of working with the number of atoms, a arbitrary characteristic number of atoms in a
macroscopic object is defined, the mole.
Definition of Mole: A mole is a number of objects. One mole is defined as the number
of carbon atoms with mass number 12 (6 protons and 6 neutrons) required to make 12
grams of carbon.
Avogadro’s Number: The number of atoms (or anything) in one mole is called Avo-
gadro’s Number, NA , and equals
NA = 6.022 × 1023
The different isotopes of an element occur naturally with different abundances. For example, there is a lot
more Carbon-12 around than Carbon-13. The periodic table lists the average mass of one mole of an element.
This mass is given the name the atomic mass.
Atomic Mass: The atomic mass of an element is the mass of one mole of the element
in grams. So if a periodic table gives the mass of Helium as 4.0026 amu, then a mole
has a mass 4.0026g.
The periodic table reports the chemical symbol, the atomic number, Z, and the atomic mass as
Z 6
Symbol ⇒ C
M ass 12.01
where the periodic table entry for carbon is given as an example. If you have already sold back your chemistry
book, the web site www.webelements.com is an excellent source of chemical information.
(a) −4.6 × 107 C (b-Answer) −4.6 × 1010 C (c) −9.2 × 109 C (d) 4.6 × 1010 C (e)
−4.6 × 1013 C
Solution to Part(a)
(a) Compute Moles of Aluminum: The atomic mass of aluminum is 26.98154u where u is an atomic mass
unit. By the definition of atomic weight, one mole of aluminum has a mass of 26.98154g. Therefore, our block
of aluminum contains
10kg
N= g
26.98154 mole
N = 370 moles of aluminum atoms,
where I have used 1000g = 1kg.
(b) Compute the Number of Atoms: The number of atoms of aluminum is the number of moles multiplied
by Avogadro’s number
atoms
N = (370 moles) 6.022 × 1023 = 2.23 × 1026 atoms of Aluminum
mole
(c) Compute the Number of Electrons: Aluminum is number 13 in the periodic table, and therefore has 13
electrons per atom. The total number of electrons in the block of aluminum is then
Solution to Part(b)
Q = −4.6 × 1010 C
From the above, a very very small percentage of the atomic charge is involved in even the largest macroscopic
charge.
1.6.2 Geometry
To calculate the total charge of an object whose charge is well described by a uniform charge density, the
charge density is multiplied by the appropriate total length, area, or volume. This means we need to recall come
basic geometry.
Solution
By definition of surface area, the total charge of the shell is Q = σS where S is the surface area. For a sphere,
the surface area is S = 4πr2 , so the total charge is
C
Q = 4πr2 σ = 4π(0.1m)2 (4 × 10−6 )
m2
Q = 5 × 10−7 C
Electric Force I
Science is, at its core, a personal experience. It allows each person the power to find out what is true in the
universe. You most believe those parts of science that you have experienced, the things you have touched and the
things you have figured out for yourself. In this chapter, we examine features of electric force and charge which
we can understand without knowing how to calculate the force.
17
2.1. BASICS OF ELECTRIC FORCE CHAPTER 2. ELECTRIC FORCE I
The next step was to get some positive charge, which I (and
you) did by rubbing a clear plastic rod with felt. I then found
Support
that the golf tube in the hanger was attracted to the charged
clear plastic rod, so the clear rod could not have the same
kind of charge as the golf tube. This in itself meant nothing
since we know that charged things attract uncharged things. I + +
and you prepared a second charged clear rod and that rod was
repelled by the first rod, so by personal observation: There _ _ _ _ _+ _+
are At Least Two Kinds of Charge and Opposite Charges _ _ _ _ _ _
Attract. This experiment has been done with many combi- + +
nations of materials and a charged object which repelled both
kinds of charge has never been found. Therefore, there are
+ +
only two kinds of charge.
rods had to be close together for anything to happen, we know The Electric Force Decreases with Increasing
Distance.
Q Q
Q1
F21 = - F12
Q1
F21 = - F12
We will shorten the second and third points to Opposites Attract, Likes Repel and use it until you’re sick to
death of it. So back to the proton in Betelgeuse. The first property, that force falls off with distance, implies that
the electric force of a charged object will exert a very small force on a distant object. Betelgeuse is 600 light years
away, so the force from one proton is pretty small. Further, Betelgeuse, like everything else in the universe, is
nearly electrically neutral, so the force of the proton is cancelled by an equal and opposite force from an electron.
Insulator Conductor
t=0 t=0
t t
8
8
This is why it is so hard to use the golf tube, which is an insulator, to directly transfer charge to anything; but
it is easy to transfer charge using an electrophorus (Course Guide 3), since the plate is a conductor. Net charge
spreads over a conductor very quickly, so using the instruments of this class we will not be able to observe the
spreading process. Instead, we will always observe the conductor in its final state. Often an insulator is used to
cover the surface of a conductor, this is called insulation. In your house wiring, the insulation takes the form of
a plastic coating on the wires. In lab, a varnish will insulate the wires. Air is an insulator in most circumstances,
but will allow charge motion if a lot of electric force is applied. A spark is electric charge moving through the air.
t=0 t>0
Since the charges forming the net charge are all of the same sign, the charges, in this case electrons, will repel
each other, pushing each other as far apart as they can get. In a very short time, the electrons spread out on the
outside of the bucket as far apart as they can get. What would happen if the same experiment was done with
an insulator, spraying charge on the interior of a plastic bucket? The charges would stay where they were placed.
The behavior of the electrons on the conducting buckets leads to our first two principles of how electric charge
behaves in a material.
Net Charge Spreads out on the Outer Surface of Conductors: Because like charges
repel, charges try to get as far apart as possible, so charge spreads out on the outer
surface of a conductor.
-5C 2C
Valve
Conductor 1 Conductor 2
What happens when the valve is opened allowing the charge to mix? Since positive charge attracts negative
charge, the positive and negative charge will get as close together as possible—excluding the charges that cannot
find mates. X X
Qinitial = −5C + 2C = −3C = Qf inal
Therefore, once the valve is opened we will have the following
-1 1/2 C -1 1/2 C
Conductor 1 Conductor 2
Q
If the objects are identical, we would have f 2inal = − 32 C on each sphere. The other + and − charges were not
destroyed, but are paired up with a + charge very close to a − charge. In most processes, the total number of
electrons and protons remains the same. In ALL processes, the total charge is conserved.
Charge Sharing: If two conductors are placed in electrical contact, any net charge on
either conductor will spread out on both conductors. This will be called charge sharing.
Fnet
rod
This causes a surface charge density in a neutral insulator when immersed in an external electric field. This effect
is called polarization.
Polarization of a Dielectric: If a dielectric
is brought into the field of a fixed charge, +
atomic charge separates slightly producing a
surface charge density as drawn to the right. + Dielectric
Unlike charge separation in a conductor, no +
charged particle has moved more than the
distance across one atom.
There is still an attractive force between a charged object and an uncharged insulator. The induced surface charge
density on the insulator is equal and opposite, but the electric force falls off with distance causing a net attractive
force as shown below.
neutral
tube plastic
Ftube,+ Ftube,-
bottle
Fnet
Solution
Paper
neutral
tube
cut wire
First consider the system before the valve is opened. Because S2 is larger than S1 , the electrons carrying the
net charge of sphere S2 are farther apart than those in S1 . Like charged objects push on one another (repel) and
that force decreases with distance. Therefore, since the charges on S1 are closer together than the charges on
S2 , they feel a larger electric force pushing each other apart. That is, they feel a larger electric pressure. What
happens when we open the valve?
The charges in S1 and S2 both feel a force pushing each other apart and into the wire connecting the spheres,
but the charges in S1 feel a much greater force than the charges in S2 . So some of those charges will be pushed
from S1 to S2 until this force equalizes, leaving Q1 < Q2 . S2 holds more charge than S1 if they are connected
by a wire. We will call the amount of electric charge an object can hold its capacity and learn to compute it in a
few weeks.
Q1 Q2
2.3.2 Grounding
In most applications, a net static charge is a very undesirable thing. So we would like to have a way to remove
a net charge from a system. For an insulator, we’re out of luck. You just have to try different things (like washing
the system in water) to remove the charge. To remove charge from a conductor, we can use the fact that charge
is shared unevenly between conductors of different sizes. If a conductor is connected to a much larger conductor,
almost all of the net charge will move to the larger conductor. The largest conductor available is the earth, with
a radius of Rearth = 3, 963miles. If a conductor with a net charge is connected to the earth, the charge will be
shared in the ratio
Qconductor Rconductor
≈ ≈ 0 ⇒ Qconductor ≈ 0
Qearth Rearth
so the earth absorbs most of the charge. In lab, we have been removing charge from electroscopes and electrophorii
by touching them. When we do this, our body becomes the large conductor, and we make the approximation
Rconductor
≈ 0.
Rus
The objects in lab involve relatively small amounts of charge. For larger charges or larger objects, it is best to
remove the charge by electrically connecting them to the earth. This is called grounding.
The earth is filled with water, if you dig a bit, but it is a rather poor conductor in general, so you have to
work to form a good ground. If a wire is touched to earth, the place where contact is made may be insulating
because it’s dry or a rock and a poor connection will be formed. To form a good connection, a long conductor is
used. In a house, an 8ft steel pole is driven into the earth.
conductor
ground
Solution
(a) As she jumps on the trampoline, her feet rub the fabric of the
trampoline and charge (assumed to be negative for this discussion) is SD
transferred to her. The charge distributes on the surface of her body, -
- -
even to the surfaces of her hair. Since the strands of hair now have
- -
like charge, they repel one another so her hair stands up.
- -- -
- -
-- - -- -
- -
Earth
(b) When she touches the metal, the excess charge is transferred through the frame to Earth — the spark is the
visible evidence of this charge transfer. Since she is not charged, her hair returns to normal.
2.4 Shielding
2.4.1 Shielding Effect of a Conductor
When a charged object is brought near a conductor, the charge separates slightly
on the conductor. The amount of charge which separates is very small compared
to the amount of charge available.
conductor
The separated charge will not usually exceed a 1µC, whereas the available atomic charge is on the order of 1×105 C.
Why doesn’t all the available atomic charge separate? Let’s build up the separated charge one electron at a time
and examine the force felt by the next electron to be added to the separated charge. The net force on the first
electron is the full electric force, F~Q , of the external charge. The net force on the second electron which moves
to the surface is the sum of the external electric force F~Q to the left AND the electric force F~e1 to the right from
the first electron separated AND a force, F~p1 , to the right from the positively charged region left behind by the
first electron. Therefore the force from the first charge added to the separated surface charge partially cancels the
external electric force. The total shielding force due the first electron is represented as F~s1 in the middle figure.
The size of F~s1 is greatly exaggerated. As more charge separates, the electrons (and protons) in the interior of
the conductor feel progressively less NET force, because the force from separated charges partially cancels the
externally applied force. The charge separation continues until the external force is completely cancelled by the
electric force, F~sN , of the separated charge as shown in the third figure. The electrons in the interior of the
conductor then feel zero net electric force.
FQ FQ Fs1 FQ FsN
e1 e1 e2 eN
Q Q Q
First electron feels Net force on second Zero net force on Nth
full electric force electron reduced electron
The Inside of a Conductor is Shielded from the Electric Force: The mobile charge
inside a conductor feels no net electric force because the separated surface charge cancels
the external electric force. This effect is called shielding.
From this analysis, we can estimate that the amount of charge separated is not more than the external charge,
and probably much less.
Amount of Shielding Charge: Since the separated surface charge is closer to the
conductor than any external charge, it can be assumed that the amount of separated
charge is not greater than the external charge. If the external charge causing charge
separation is very near the conductor, the separated charge will be on the order of the
external charge.
This effect is very useful. It means the surface charge den-
sity on a conductor due to charge separation produces zero No net force on q
electric force inside a conductor, thus shielding the interior
of the conductor from the electric force. Many instruments
and devices are adversely affected by electric forces, so this q
shielding effect is very useful. Therefore, if we place a charge
inside a hollow space in a conductor, it will feel zero electric cavity
force from the outside world. conductor
Solution to Part(a)
Solution to Part(b)
The pith ball feels zero total electric force. The pith ball still feels the electric force from the golf tube, but this
is cancelled by an equal and opposite electric force from the separated charge on the bucket.
Net Charge
In the picture at the left above, charge separates until the interior of the conductor feels zero net electric
force. The separated negative charge is held in place by the fixed charge. The separated positive charge is mostly
shielded from the force of the fixed charge and is not held in place. When a ground is connected, the positive
charge is free to escape to the ground and greatly increases its separation. Note: for you atomists, what happens
atomically is that negative electrons are drawn from the ground to neutralize the region of net positive charge. If
the ground is disconnected while the fixed charged object is still near, the conductor will have a net charge.
Definition of Charging by Induction: An object is said to be charged by induction if
a fixed charge is brought near a neutral conductor causing charge separation and the
conductor is then grounded, removing separated charge not held in place by the fixed
charge, thus leaving a net charge on the conductor.
In charging by induction, the charge that escapes to the ground is not pushed away by the fixed charge, it is
simply not held in place and so is free to move farther away by moving to the ground. This means the amount
of charge we produce on an object is approximately the amount of charge needed to shield the interior of the
conductor from the electric force, so by our argument before, it should be of the same order of magnitude as the
fixed charge if it is close to the conductor.
The following example illustrates the escape of the untrapped charge a little more clearly. Remember there is
plenty of free atomic charge available at all points in a conductor, it is not just the small number of + and − we
draw on the figures that are in play.
System Neutral
Grounded
Charging by Friction: When two materials are rubbed against one another, there is
sometimes a charge transfer between the two materials. For example, you can charge a
balloon by rubbing it in your hair. My daughter is charged by jumping on a trampoline.
My car is charged when I drive it through dry air and I can charge my clothing by
rubbing it against the dryer walls.
Charging by Adhesion: Sometimes when two objects are stuck together, like two
pieces of tape or two sheets of foam insulation, pulling them apart will transfer charge
from one to the other.
Charging by Spraying Charge or Sparking: An object can be charged by spraying it
with charged particles. The earth is bombarded with charged cosmic rays. A pool can
be charged by a bolt of lightning. The screen of your TV is charged by the electrons
sprayed on it by the cathode ray tube.
Charging by Induction: A conductor can be charged by holding some of its charge in
place with another charged object and allowing the opposite charge to flow to ground.
Charging By Pumping Charge: The easiest and most common way to move charge is
to pump it between two conductors, or between the earth and one conductor. A battery
is a charge pump. We use the more exotic charging mechanisms in lab because it is
hard to develop a net charge whose force is observable using safe voltages.
Charging by Separation: A pair of objects can be given equal and opposite charges by
placing them in contact and causing their charge to spread by using a charged object,
and then separating the objects.
2.7 Summary
2.7.1 Summary of the Basic Principles of Electrostatics
Before continuing let’s gather together the main principles used in analyzing the behavior of electrostatic
systems.
Charge is Conserved:
1 1
Q Q
2 2
In the same way, if you want 31 Q you can take three uncharged spheres and
put them in a triangle. So if you wanted charges 2q and 3q you could do the uncharged uncharged
following: 1. Charge up a single ball with a charged rod or by touching it to
Q
the Van de Graaff generator. 2. Split that charge into two equal charges Q
by touching the charged ball with an uncharged ball. 3. Divide Q in half and
in thirds as in the previous examples, which gives you charges 21 Q and 13 Q.
Define q = 61 Q and you have 21 Q = 3q and 13 Q = 2q.
1
Q
3
1 1
Q Q
3 3
The symmetry is crucial because in the three spheres placed in a line at the
right, to get as far apart as possible most of the charge will be on the two All Charge Same Due to Symmetry
outer spheres leaving less charge on the center sphere. We could conclude the 1
Q
3
spheres on the ends have the same charge by symmetry. 1 1
Q Q
3 3
We use the same size spheres because the charge is free to move in a conductor,
so there is no reason for the charge on two identical conducting spheres to be
Generally Larger Objects Hold More
different. If, however, we use two different size spheres, the charge will be Charge at the Same Energy
greater on the larger sphere.
Q1 Q2
Q1 < Q2
Great! So what if I need +q and −q. Take two uncharged conducting spheres
and put them in contact near the charged golf tube.
_ _
_ _
_ _ _
+ _
_ _ +
+ _
_ _ + _
_ _
Now move the spheres apart and the total charge of the two spheres must be
zero (since we didn’t touch the spheres with the golf tube, they still have the Separate spheres in the presence of the charged
object.
same TOTAL charge they had at the beginning). _ _
_ _
_ _
_ _ +Q −Q
_ _
_ _
(i) Bring the golf tube near marble A and ground the marble. Remove the ground before removing the tube.
(ii) Without loss of charge, bring marbles A and B into contact, then separate them.
(iii) Then without loss of charge, bring marble C into contact with marble B.
The final charges are, if marble A was given a charge +Q in step (i):
Solution
In step i, the marble is charged by induction to a charge of +Q. In step ii, the identical spheres are brought in
contact and they share charge equally, QA = QB = Q/2 In step iii, the two spheres share the charge equally
leaving QA = Q/2, QB = QC = Q/4, so the answer is (e).
Electrostatic Devices
At the end of Course Guide 2, we illustrated some of the basic features of the electric force and its action on
a conductor with some experiments with identical conducting spheres. Unfortunately, these will have to remain
thought experiments since the charge developed on the marble is too small to detect. In lab, we will construct
two classic electrostatic devices used by early experimenters to produce large net charges, the electrophorus, and
to detect small net charges, the leaf electroscope.
3.1 Electrophorus
We would like to build something which places a large net electric charge on a conductor, so that charge could
easily be transferred to whatever we want. We know that we can place a large net charge on an insulator. Using
the insulator we can charge a conductor by induction. We have reasoned that since the electric force decreases
with increasing distance, the amount of charge produced using charging by induction depends on how close the
fixed charge is to the conductor to be charged. It also seems reasonable that the total charge produced should
depend on the total area of the conductor brought near the insulator. Therefore, to produce a large net charge
using charging by induction we need to bring a large surface of a conductor very near a fixed charge. We can do
this by charging a flat insulating (plastic) plate and placing a flat piece of metal on it. The charge in the metal
separates. If the metal plate is grounded while sitting on the plastic, it will obtain a large net charge. This device
is called an electrophorus.
_ _ _ _ _
Metal
_ +
_ _+ _+ _+ _+ _
Plastic
35
3.1. ELECTROPHORUS CHAPTER 3. ELECTROSTATIC DEVICES
The electrophorus is charged by induction using our body as the ground. To charge the electrophorus, I rub the
cutting board with the oven bag. I put the pie pan on the charged plastic board and grounded it with my finger.
I felt a spark when I grounded it. I picked up the plate (by the handle) and tried to ground it again, and I felt a
spark. Here’s a picture of what happened:
_ _
+ _ _ _ _ +
_ _ _ _ _ _ _ _ + _ + _+ +_ _
Since the metal plate is so close to the fixed charge on the insulator, it is a good approximation that the charge
density on the electrophorus is equal and opposite where the electrophorus makes contact with the insulating board.
The net charge produced on the electrophorus by charging by induction is then the charge density of the insulating
plate multiplied by the area of the electrophorus that makes contact with the plate. The sign of the charge of
the electrophorus is opposite that of the fixed charge on the insulator. You will understand this quantitatively in
Course Guide 9.
Solution
+ + Finger
+ + +
_ _ _ +_ _ _
The behavior of the leaf electroscope charged with the electrophorus when a negatively charged golf tube is
brought near that you observed in lab indicates the electrophorus charges positive and the wallboard negative.
_ _
_ _ ++ ++ + +
__ + + + + ++ + +
+ + + + ++ + + ++
+ +
++ ++ +
+ + + + +
+ +
+ + +
+ + +
+ +
+
+ + +
Detecting Charge with a Leaf Electroscope: The leaf electroscope detects charge
in two ways. If a charged object is brought near the steel bolt, without transferring
charge, then the electric force causes charge to separate in the electroscope, causing
a deflection of the foils. The foils also deflect if a net charge is transferred to the
electroscope.
+ +
+ +
_ _
_ _
++
+
+
_
_
_ _
Charge separation happens when a charged object is brought near a conductor. Opposite charges are drawn
closer to the charged object and like charges move farther away. This can be observed visually using a leaf
electroscope. When a charged object, like the golf tube, is brought near the electroscope, but not so near that
charge transfers, the leaves of the electroscope deflect indicating a net charge on the leaves. However, when
the tube is removed the foils hang straight down indicating that no charge was transferred and the electroscope
remained neutral throughout the process. What happened? The negatively charged tube attracted positive charge
to the bolt of the electroscope, leaving a net negative charge behind in the leaves. This is charge separation. It
turns out that since the tube is an insulator, there is no charge transfer even when the tube touches since charge
cannot move through an insulator.
__ _ __ _ __ _ _ _
_ _ _ _ _ _
_ _ _ _
_
_ _
Now, let’s return to the materials we worked with in lab and charge our leaf electroscope by induction. Charge
up a golf tube producing a negatively charged object and bring the golf tube near the bolt of the electroscope.
Charge separates in the electroscope and the leaves deflect. Touch the electroscope with your finger, thus
grounding the scope. You are a much larger conductor than the scope and are a good ground for it. Remove
your finger before the golf tube is removed and the electroscope will have a net positive charge.
Finger
Touch
_ _
_ _ the electroscope negative charge escapes into
your body leaving net
positive charge in the electroscope.
The order you do things in charging by induction is very important; if we touch the electroscope after the
charged object is removed, the electroscope is grounded and has zero net charge.
+ ++
+ +
++ +
+
Finger
Charged Electroscope
Charge escapes into
+ +
body, electroscope
+ + grounded
+ +
+ +
Problem: In , the instructor charged a leaf electroscope by induction using a negatively charged golf tube.
This gives the electroscope a net positive charge. A negatively charged golf tube was then brought close and the
behavior observed. Draw the charged electroscope with the golf tube near and far away (2 diagrams). Physically,
explain the behavior of the system in each drawing.
Solution
(a) Golf Tube Far Away: If the golf tube is far from the electroscope, the
repulsion between the like charges causes the positive charges to spread out over
the electroscope. The repulsive forces of the opposite charges on the leaves of
the electroscope cause the leave to deflect.
(b) Golf Tube Near: When the charged golf tube is brought near the charged
electroscope, the net positive charges are brought into the bolt, leaving the
leaves neutral, because opposite charges attract. The leaves are then uncharged
and do not deflect.
Step I Step II
Charge a Leaf ++ + + + +
+ +
Electroscope + + + +
+ + + + Touch it with a
+ + 100ft spool of
+ + + +
+ + insulated wire
+ + Still Slightly + +
Charged not connected to
anything
+ + Electroscope is once
After disconnecting
the wire, ground + + again charged, the
the electroscope charge was in the wire.
_ _
Negatively
Positively + + +
+ _
_ Charged
_
charged + +
Switch _ Electroscope
electroscope
+ + _ _
+ + _ _
+ + _ _
II A charged rod was brought near the bolt of one of the electroscopes, while still far from the other
electroscope. The leaves deflect as shown.
III The wire between the electroscopes is cut, without loss of charge.
No charge was transferred from the charged rod to the electroscope at any time. No charge is lost to the
environment at any time. The wire allows the flow of charge but is sufficiently fine to contain none of the net
charge. The electroscope far from the golf tube is sufficiently distant not to feel a force from the golf tube.
_ __
rod _ _ wire
wire _
II
I _
_ _
rod _ _ cut wire
_
III IV
reconnect wire
The charge is drawn in the figure below. The system of two electroscopes remains neutral so equal amounts
of opposite and negative charge is drawn. Opposites attract so the electroscope nearer the tube is positive.
No charge is in the leaves in figure III because they hang straight down. Charge spreads out throughout the
electroscope in figure IV.
_ __ + + _
rod _ _ wire
wire _ + +
_ _
II
I _
_ _ + _ _
rod _ _ + cut wire +
_ + +
_ + _
+
_ _ + _ _
III IV
The leaves are filled with a net charge of the same sign. Like charges repel, causing the leaves to push apart.
Since no charge was transferred to the system and no charge was lost to the environment the total charge of the
two scopes must be zero by conservation of charge.
The electroscopes have opposite charges so they attract each other with equal and opposite forces.
Electric Force II
In the previous chapter, quite a bit of time was spent learning to reason about the effects of the electric force
and the behavior of charge in materials. Now its time to crunch the numbers.
Coulomb’s Law (Version 0): The force an object with charge q1 exerts on an object
with charge q2 has the following properties:
• The direction of the electric force is along the line shared by the centers of the
two charged objects.
2
where d is the distance between the centers of the two objects and k = 8.99 × 109 Nm
C2 .
Solution
45
4.2. REVIEW OF BASIC VECTOR CONCEPTS CHAPTER 4. ELECTRIC FORCE II
F
1cm
qB
d
qA 1cm 2cm
Solution
(a) Compute the Magnitude of Force: The direction of the electric
force
is outward along the line connecting
kq q
the charges (Likes Repel). The magnitude of the force is Fe = dA2 B where d is the distance between the
charges. Using the Pythagorean theorem for the charges drawn, d2 = (2cm)2 + (1cm)2 = 5cm2 = 5 × 10−4 m2 ,
so 2
(8.99 × 109 Nm −6
C)(−2 × 10−6 C)
C2 )(3 × 10
Fe = = 108N
5 × 10−4 m2
= 100N with significant figures.
(b) Write the Force as a Vector: Force is a vector, therefore both a magnitude and a direction must be
reported.
F~e = 100N directed outward along the line connecting A & B
The above is a perfectly valid expression of the force, but it would be more useful if we could write it as:
F~e = (Fx , Fy , Fz ) or
F~e = Fx x̂ + Fy ŷ + Fz ẑ or
F~e = Fx î + Fy ĵ + Fz k̂
I called the above form of Coulomb’s law “Version 0” because the vector form of the law which follows is the
formulation that I think of as Coulomb’s Law.
Writing a Vector in its Coordinate Form: A vector may be written by giving its length
along each of the axes of a coordinate system. Each of the following are equivalent.
F~ = (Fx , Fy , Fz ) = Fx x̂ + Fy ŷ + Fz ẑ = Fx î + Fy ĵ + Fz k̂
Definition of Coordinate Unit Vectors: î = x̂, ĵ = ŷ, and k̂ = ẑ are vectors with
length one that point along the coordinate axis. You can use either î, ĵ, k̂ or x̂, ŷ, ẑ,
but don’t mix them in the same problem.
Vector Addition: The sum of two vectors is found by adding their components
F~R = F~1 +F~2 = (F1x , F1y , F1z )+(F2x , F2y , F2z ) = (F1x +F2x )î+(F1y +F2y )ĵ+(F1z +F2z )k̂
Vectors are graphically represented as arrows and addition and subtraction of vectors can be done graphically.
−r1
r12
Definition of Unit Vector: A unit vector is a vector of length one. Given a vector ~r,
whose modulus is |~r|, the unit vector is
~r rx ry rz
r̂ = = x̂ + ŷ + ẑ.
|~r| |~r| |~r| |~r|
The unit vector is a vector of length 1 with the same direction as the vector.
Writing Vector as Magnitude
q and Unit Vector: A vector ~v can be written as its
magnitude |~v | = v = vx + vy + vz2 multiplied by a unit vector, v̂ = ~v /|~v |, in its
2 2
direction,
~v = |~v |v̂.
Solution to Part(a)
The displacement vector, ~r12 , is a vector which points from point 1 to point 2,
~r12 = ~r2 − ~r1 = (3cm − 1cm, 3cm − 3cm, 0 − 5cm)
= (2cm)x̂ + 0ŷ − (5cm)ẑ
Solution to Part(b)
Solution to Part(c)
The vector r̂12 is a unit vector (a vector of length one) in the direction of the vector ~r12 . The unit vector is found
by dividing ~r12 by its length r12 .
~r12 2 5
r̂12 = = √ x̂ − √ ẑ
r12 29 29
A key step to visualizing and simplifying the calculation of the electric force of a system of point charges is
correctly drawing a diagram.
Use Opposites Attract/Likes Repel to Get Direction: Use opposites attract/like
repel and the fact that force acts along the line between the charges to get the direction
of forces correct. In the diagram below, q and q1 have the same sign so the force on
q points away from q1 (Likes Repel). q and q2 have different signs, so the force on q
points toward q2 (Opposites Attract).
y q2
q1 , q > 0 FT
q2 < 0
F2q
q1 = q2
F1q
q x
q1
where ~rAB is a vector which points from the location of object A to the location of
2
object B and k = 8.99 × 109 Nm
C2 .
To successfully use the vector form of Coulomb’s Law, we need to be able to manipulate the related quantities
~rAB , rAB , and r̂AB . I will restate their definitions for Coulomb’s law.
The Position Vector: The position vector, ~rA , for point A located at the coordinates
(Ax , Ay , Az ) is a vector from the origin to the point A, ~rA = Ax x̂ + Ay ŷ + Az ẑ.
The components of the position vector will be denoted by rAx = Ax , rAy = Ay , and
rAz = Az .
The Displacement Vector: The vector ~rAB points FROM point A TO point B and
has the length of the distance between the points. The displacement vector can be
computed as the difference of the position vectors
Modulus of the Displacement Vector: The length of the vector ~rAB will be denoted
by rAB and is the distance from point A to point B. It can be computed in the same
way as the length of any vector
q
2
rAB = |~rAB | = rABx 2
+ rABy 2
+ rABz ,
where the symbol |V ~ | represents the mathematical operation of taking the length of a
vector, called the vector modulus.
Unit Vector for the Displacement Vector: The vector r̂AB is a vector with length 1
(no units) pointing from point A to point B. It can be computed using
~rAB rABx rABy rABz
r̂AB = = , ,
rAB rAB rAB rAB
Let’s return to the calculation of the force charge A exerts on charge B and do it using vectors.
1cm FAB
B
rAB
A 1cm 2cm
x
Solution
(a) Use Coulomb’s Law: The electric force object A with charge qA exerts on object B with charge qB is
kqA qB
F~AB = 2 r̂AB
rAB
(b) Compute the Displacement Vector: The displacement vector points from the location of A to the location
of B,
~rAB = (2cm, 1cm, 0) = r̃B − r̃A .
which is consistent with the drawing since we have to move +2cm in the x direction and +1cm in the y direction
to move from A to B.
(c) Compute the Length of the Displacement Vector: The length of the displacement vector is by definition
of vector modulus
q p √
rAB = |~rAB | = rAB2 2
+ rAB 2
+ rAB = (2cm)2 + (1cm)2 + 02 = 5cm.
x y z
This had better equal the distance between points calculated earlier.
(d) Compute the Unit Vector: The unit vector for the displacement vector is by definition
~rAB
r̂AB =
rAB
~rAB 1
r̂AB = =√ (2cm, 1cm, 0)
rAB 5cm
2 1
= ( √ , √ , 0)
5 5
The unit vector r̂AB should end up dimensionless, which it did. It also had better have length 1
q
2
|r̂AB | = r̂AB 2
+ r̂AB 2
+ r̂AB
x y z
s 2 2
2 1
= √ + √ +0
5 5
r
5
= =1
5
(e) Substitute into Coulomb’s Law: Substitute √ the vectors computed above into Coulomb’s Law and turn
the crank, being careful to convert cm properly, ( 5cm)2 = 5cm2 = 5 × 10−4 m2 ,
kqA qB
F~AB = 2 r̂AB .
rAB
2
(8.99 × 109 Nm
C2 )(3 × 10
−6
C)(2 × 10−6 C)
2 1
F~AB = √ , √ , 0
5 × 10−4 m2 5 5
2 1
F~AB = 100N √ , √ , 0
5 5
F~AB = (89N, 45N, 0)
The previous example placed one of the charges at the origin. The next example has both charges away from
the origin.
Solution
Definitions
y(cm)
qB = 1.0µC ≡ Charge of qB
1.5 qB
FAB qA = −1.0µC ≡ Charge of qA
~rB = (0.50cm, 1.45cm, 0) ≡ Position of qB
1 ~rA = (1.4cm, 0.70cm, 0) ≡ Position of qA
rAB
rB ~rAB ≡ Vector from qA to qB
qA F~AB ≡ Force on qB due to qA
0.5 rA
x(cm)
0.5 1 1.5
Strategy: Calculate the force between two point charges using Coulomb’s Law; this requires calculating the
position vector in both magnitude and direction.
(a) Draw a Good Diagram: The charges are placed at the given locations. Since unlike charges attract, the
force exert by qA on qB is attractive.
(b) Use Coulomb’s Law: The force on qB from qA is given by: F~AB = kqrA2 qB r̂AB
AB
(c) Use Definition of Displacement Vector: The displacement vector is a vector that points from the location
of charge A to the location of charge B.
~rAB = ~rB −~rA = (rBx −rAx , rBy −rAy , rBz −rAz ) = (0.50cm−1.4cm, 1.45c−0.70cm, 0−0, ) = (−0.90cm, 0.75cm, 0)
= (−0.90cm, 0.75cm, 0)
(d) Use Definition of Vector Modulus: Calculate the length of the displacement vector,
q p
2
rAB = rABx 2
+ rABy 2
+ rABz = (−0.90cm)2 + (0.75cm)2 + (0)2 = 1.2cm
kqA qB
F~AB = 2 r̂AB
rAB
2
(8.99 × 109 Nm
C2 )(1.0 × 10
−6
C)(−1.0 × 10−6 C)
= · (−0.75, 0.63, 0)
(1.2 × 10 m)2
−2
Example 4.6 How Qualitative Features of the Electric Force are Represented in Coulomb’s
Law?
Problem: Coulomb’s Force Law relates the force object 1 exerts on object 2, to the charges of the objects and
the distance r12 between the objects:
kq1 q2
F~12 = 2 r̂12
r12
where k is a constant and r̂12 is a unit vector in the direction of the vector ~r12 which points from the center of
object 1 to the center of object 2. This equation contains a lot of information. Let’s explore.
(a)Draw two objects with charge q1 and q2 , the vector ~r12 and r̂12 . r̂12 is a vector of length 1 (but no
dimensions) in the same direction as ~r12 .
(b)What part of Coulomb’s Law represents the law“The magnitude of the electric force decreases with
the square of the distance between the charges?”
(c)What part of Coulomb’s Law represents the law,“The electric force between two objects is directly
proportional to the charge of either object”
(d)What part of Coulomb’s Law represents the law,“The electric force is directed along the line between
the charges”
(e)What part of Coulomb’s Law represents the law “Opposite charges attract, like charges repel”
The vector ~r12 points from the center of object 1 to the center
of object 2. The unit vector r̂12 is a vector of unit length, but r12
no dimensions, in the same direction. Since ~r12 is measured
in meters and r̂12 has no dimensions their lengths have no 1 2
r12
relation in the diagram.
Since F12 ∝ q1 q2 , the electric force is proportional to the magnitude of either charge.
The unit vector r̂12 points along a line from q1 to q2 . The direction of the force is either ±r̂12 .
The force is repulsive if F~12 points in the direction r̂12 and attractive if it points in the direction −r̂12 . Both
2
k > 0 and r12 > 0, so the direction of the force is given by q1 q2 r̂12 . If both charges have the same sign q1 q2 > 0
and q1 q2 r̂12 points in the direction of r̂12 so the force is repulsive. If q1 and q2 have different signs q1 q2 < 0, then
q1 q2 r̂12 points in the −r̂12 direction and the force is attractive.
People have a very had time accepting Newton’s Third Law; that for any pair of objects the force object 1
exerts on object 2 is equal and opposite to the force object 2 exerts on object 1. The next example explores
Newton’s Third Law and the electric force.
Example 4.7 Prove Newton’s Third Law for Electric Force
Problem: Consider two objects, 1 and 2, which have charges Q1 = Q and Q2 = 5Q where Q = +1µC. The
charges are at ~r1 = (0, 0, 0) and ~r2 = (10cm, 0, 0).
1 r21 2
(b) Compute F~12 : Using Coulomb’s Force Law and r̂12 = x̂, r12 = 10cm,
kQ1 Q2 kQ1 Q2
F~12 = 2 r̂12 = 2 x̂
r12 r12
2
(9 × 109 Nm
C2 )(1µC)(5µC)
F~12 = x̂ = 4.5 × 10−2 N(x̂)
(10cm)2
(c) Compute F~21 : Using Coulomb’s Law and r21 = 10cm and r̂21 = −x̂,
kQ2 Q1 kQ1 Q2
F~21 = 2 r̂21 = − 2 x̂ = −4.5 × 10−2 N(x̂)
r21 r21
F~21 = −F~12
Let ~rAB be the vector from the object with charge qA to the object with qB . The vector from B to A is then
~rBA = −~rAB . The distance from A to B equals the distance from B to A, so |~r| = ~rBA = ~rAB . The unit vectors
are related by
~rAB −~rBA
r̂AB = = = −r̂BA .
|~rAB | |~rAB |
So substituting into Coulomb’s Force Law,
kqA qb kqB qA
F~AB = 2 r̂AB = 2 − r̂BA = −F~BA .
rAB rBA
Electric Field
We directly compute the electric force for only a few point charges. We will find it much more powerful in
most cases to compute the electric field.
5
X
kqi
F~A = qA 2 r̂iP
i=1
riP
P5 kqi
where you only have to compute the quantity inside the parentheses, ( i=1 2 r̂ip )
riP
, once for each point P .
Simply multiplying the charge of any object placed at the point P by this quantity gives the force that the object
~
WOULD feel if it was placed at the point. Let’s divide by qA and separate this thing out, giving it the symbol E,
~ 5
~ P = FA = kqi
X
E r̂iP .
qA r2
i=1 iP
~ P and not E
Note, I wrote E ~ A because E
~ is a property of the charges qi and the point P. It has nothing to do
with the charge qA placed at point P. As my Texan father-in-law would say, ”Now we’re cooking with gas!”
~ at a point ~r is defined as the electric
Definition of Electric Field: The electric field E
force F~ a charge q0 would experience if placed at ~r, divided by q0 , E
~ = F~ /q0 .
55
5.2. COULOMB’S LAW FOR ELECTRIC FIELD CHAPTER 5. ELECTRIC FIELD
Units of Electric Field: The electric field is measured in Newtons per Coulomb, N/C.
Sizes of Electric Fields: Air sparks at an electric field of 3 × 106 N C . The golf tube
creates a field of 1 × 10 C at its surface, the pith ball creates a field of 1 × 103 N
6N
C at
2cm. The earth’s electric field, which none of us ever notice, is 150 NC .
There are two main things to take away from this section. First, if you have calculated the field, all you have
to do to calculate the force on an object with charge q is multiple by the field.
Calculate the Electric Force from the Electric Field: By definition of electric field,
~P ,
the electric force on an object with charge q placed at point P is the electric field, E
at point P multiplied by the charge
F~ = q E
~P
The second main point is that you can figure out the direction of the field, if you can figure out the direction
of the force a positive charge WOULD feel if placed in the field. This observation is extremely useful because you
should be quite good at using opposites attract/likes repel to figure out the direction of the electric force.
Direction of the Electric Field: The electric field at a point P points in the direction
of the electric force a positive point charge WOULD feel if placed at point P .
~ 12 = ~12
F
Applying the definition of electric field, E q2 , we derive Coulomb’s law for the electric field.
Coulomb’s Law for the Electric Field: A point charge produces an electric field that
~ 10 , produced
points radially outward from or inward to the charge. The electric field, E
by object 1 with charge q1 at point 0 is given by:
~ 10 = kq1 r̂10 ,
E 2
r10
where k = 8.99 × 109 Nm2 /C2 , and ~r10 is the vector which points from the location of
object 1 to the point 0 where the field is measured.
Definitions
y
Law of Linear Superposition: The electric field at a point P is the vector sum of the
~ iP , produced by each individual charge or charged object
electric field, E
~ total = ~ iP .
X
E E
i
Linear Superposition allows us to build more complicated fields by combining simpler fields. Coulomb’s law and
the law of linear superposition are really all we need to calculate any electric field since any charge distribution
can be subdivided into electrons and protons, which are for our purposes point charges. Their fields can be
calculated and added by linear superposition. The following example applies the Law of Linear Superposition to
point charges.
Solution
q2
r2P r1P
E2P
P
EP E1P
(b) Calculate Field of q1 : First find the field due to q1 . The displacement vector, modulus, and unit vectors
are
~r1P = ~rP − ~r1 = (1cm − 2cm)x̂ + (−3cm − 3cm)ŷ + 0 = −1cmx̂ + −6cmŷ
p √
r1P = (−1cm)2 + (−6cm)2 = 37cm
−1cm −6cm 1 6
r̂1P = √ ,√ = − √ , −√
37cm 37cm 37 37
So the field at P due to q1 is
2
9 Nm −9
~ 1P = kq1 r̂1P = (8.99 ×√10 C2 )(3 × 10 C)
1 6 N N
E 2 − √ , −√ = −1.20 × 103 x̂ − 7.19 × 103 ŷ
r1P ( 37 × 10−2 m)2 37 37 C C
(c) Calculate Field of q2 : Now find the field due to q2 . The displacement vector, modulus, and unit vectors
are
~r2P = ~rP − ~r2 = (1cm − (−2cm))x̂ + (−3cm − 1cm)ŷ + 0 = 3cmx̂ + −4cmŷ
p
r2P = (3cm)2 + (−4cm)2 = 5cm
3cm −4cm 3 4
r̂2P = , = ,−
5cm 5cm 5 5
So the field at P due to q2 is
2
(8.99 × 109 Nm
C2 )(−2 × 10
−9
C) 3 4
~ kq2 N N
E2P = 2 r̂2P = ,− = −4.32 × 103 x̂ + 5.76 × 103 ŷ
r2P (0.05m)2 5 5 C C
(d) Add Fields Using Linear Superposition: Now, by the principle of superposition of electric fields, we can
simply add them up.
~P = E
E ~ 2P = (−1.20 × 103 N x̂ − 7.19 × 103 N ŷ) + (−4.32 × 103 N x̂ + 5.76 × 103 N ŷ)
~ 1P + E
C C C C
Draw a Fixed Grid: To represent the vector field, select a uniform set of points and
draw the field vector at each point with appropriate direction and length.
The First Vector Sets the Length: The first vector you draw sets the scale for all
the other vectors.
Reason Using Ratios: Don’t compute the actual length of the vectors unless you have
to. In the diagram in an example that follows, I used the fact that twice as far from
the charge the field is 4 times weaker.
The Field is at the Tail of the Arrow: The vector represents the force or field at its
tail, not its point.
Solution
Solution
+Q
C B
The
√ electric field points radially outward and increases in strength
√ proportional to the distance. The point B is
~ B | = 2|E
2 ≈ 1.5 (1.41) as far from the origin as point C, so |E ~ C |. Using this observation, draw field arrows
at each of the dots.
y
E E
A
Fq
aq
x
E E
Fq Fq
C B
E E
Since q = −1µC < 0, the force vector is opposite the direction of the electric field. The force vectors must be
drawn with magnitude proportional to the field since we asked for the force on the same charge.
The acceleration ~aq is always in the same direction as the force, since F~ = m~a and m > 0.
~ x<0 = −σ x̂
E
2ε0
~ x>0 = σ x̂
E
2ε0
2
C
where ε0 = 8.85 × 10−12 Nm 2 . Notice, the field does not change with the distance from
the plane.
2
C
Permittivity of Free Space: The constant ε0 = 8.85 × 10−12 Nm 2 is called
the“permittivity of free space” and is one of the fundamental constants of the uni-
verse.
Relation of k and ε0 : The constant k found in Coulomb’s Law is related to ε0 by
1
k=
4πε0
Solution
Solution
~ = σ outward. Therefore, if the plane occupies the y −z plane
The electric field of an infinite plane of charge is E 2ǫ0
and the charge density is positive, then the electric field the +x side of the plane is in the positive x̂ direction
and the electric field on the −x side of the plane is in the −x̂ direction. The magnitude of the electric field is
σ 3 × 10−8 mC2 N
E= = C2
= 1695
2ǫ0 2(8.85 × 10−12
Nm2 )
C
The electric field must be reported as a vector, so the electric field of the plane is
~ = 1695 N x̂
E x>0
C
~ = −1695 N x̂
E x<0
C
For the golf tube, we will use the approximation of an infinite straight line with uniform linear charge density
λ, when the field point is near the tube, away from the ends, but not inside the tube.
Electric Field of an Infinite Line of Charge: The electric field of an infinite straight
line of charge is
~ r) = λ r̂
E(~
2πε0 r
where ~r points straight outward perpendicular to the line of charge.
Solution
x
d 2d
z
tube
λ<0
rtube, P
rtube, P
1cm
E
P x
1cm rtube, P
Solution
(a) Displacement Vector: The displacement vector can be read from the diagram. To move from the tube to
P , we move +2cm in the x direction and −5cm in the z direction. From tube to P,
By working with non-uniform charge distributions we will produce quite a variety of fields. In this section, we
~ r) = E(x,
work with E(~ ~ y, z) as a general vector function to gain some experience with functions that return a
vector.
Example 5.10 Computing the Electric Field at a Point Given the General Field
Problem: The electric field in a region of space is given by
~ r) = x2 ( N )x̂ + y 2 ( N )ẑ.
E(~
Cm2 Cm2
(a)Compute the electric field at the point ~r0 = (2m, 1m, 0)
(b)Compute the electric force a −10µC charge would feel at ~r0 .
(c)Compute the strength of the electric field at ~r0 .
Solution to Part(a)
Solution to Part(b)
By the definition of the electric field, the electric force is the field multiplied by the charge q,
F~0 = q E
~ = (−10µC)(Ẽ(r̃0 ))
N N
= (−10µC)(4 x̂ + 1 ẑ)
C C
= −4 × 10−5 Nx̂ − 1 × 10−5 Nẑ
Notice that, because the charge is negative, the force is in the opposite direction as the field.
Solution to Part(c)
The strength of the electric field is the length or magnitude of the electric field vector,
q
~ = E2 + E2 + E2
|E| x y z
r
N N √ N
~ r0 )| =
|E(~ (4 )2 + 02 + (1 )2 = 17 .
C C C
That’s a pretty small field.
There are three general “shapes” of the electric field that we will encounter over and over: the uniform field,
the radial field, and the cylindrically radial field. The simplest is the uniform field, which has the same magnitude
and direction at every point in space.
Definition Uniform Field: A uniform field is one that has the same value at all points
~ = (Cx , Cy , Cz ), where the C’s are constant.
in space, so it can be written as E
Solution
Because the field is uniform, if we know the field at one point we know the field at all points. The electric field
~ r) = 50 N x̂.
is E(~ C
The electric field of a single point charge is a radial field, which means that at every point in space it is
directed either inward or outward along a ray connecting that point in space and the single point charge. The
direction of a radial field is given in terms of the unit vector r̂ which points outward from the origin at every point.
Definition Radial Field: A field that has the form E ~ = f (r)r̂, where the vector
~r = (x, y, z) and f (r) is a function only of the distance from the origin is called a radial
field. The field points inward or outward along the radius of a sphere.
Solution
(a) Write the Electric Field: From the information given, we can write the electric field as
~ rA ) = γ √rA r̂A ,
E(~
where the r̂ directional dependence is deduced from the fact that the field is radial.
(b) Compute the Vectors: To compute the field we need the quantities r̂A and rA . By definition of vector
modulus, p √
rA = |~rA | = 02 + (5cm)2 + (3cm)2 = 34cm.
By definition of unit vector,
~rA (0, 5cm, 3cm) 5 3
r̂A = = √ = 0, √ , √
rA 34cm 34 34
(c) Compute the Electric Field: Substitute into the expression for the electric field.
√
q
~ rA ) = γ √rA r̂A = 100 √
E(~
N
34 × 10−2 m 0, √ , √
5 3
C m 34 34
~ rA ) = 24 N
E(~
5
0, √ , √
3
C 34 34
(d) Compute the Electric Force: By definition of electric force, the force on charge qA due to the electric
field is
~ ~ −6 N 5 3
FA = qA E(~rA ) = (5 × 10 C) 24 0, √ , √
C 34 34
5 3
= (1.2 × 10−4 N) 0, √ , √
34 34
We also encounter electric fields which point radially outward from the axis of a cylinder.
Definition Cylindrically Radial Field: A radial field points outward or inward in all
directions from a point in space. A cylindrically radial field points straight outward
~ r) = f (r)r̂ where
or inward from a line. A cylindrically radial field will be written E(~
~r = (x, y, 0) if the field is cylindrically radial about the z-axis.
~ tube,P = λ
E r̂tube,P .
2πǫ0 rtube,P
So let’s compute the electric field of this assemblage of junk.
Example 5.13 Many Different Fields
Problem: The x − y plane is occupied by a uniform sheet of charge with charge density σ = 1.6 × 10−7 C/m2 .
An infinite line of charge runs parallel to the y-axis through the point −5cmx̂ and has charge density λ =
0.5 × 10−6 C/m. A point charge with charge q = 10nC is placed at the point +5cmx̂. Consider the electric field
at a point P at +5cmẑ.
(a)Draw a good diagram including the direction of the electric field of the three objects at point P .
(b)Compute the electric field of the infinite plane of charge at point P .
(c)Compute the electric field of the infinite line of charge at point P .
(d)Compute the electric field of the point charge at point P .
(e)Compute the total electric field at point P .
Definitions
z
Strategy: Compute the electric field of each charge at the point P , then add using linear superposition.
Draw a Good Diagram: The electric field of the plane charge points straight upward in the +ẑ direction.
The electric field of the line charge is repulsive and the electric field of the point charge is repulsive. We cannot
determine the relative magnitudes of these fields without actually computing the field, so I do not draw an
approximation for the resultant on the diagram.
Compute the Electric Field of the Plane Charge: The electric field of an infinite plane of charge in the x − y
plane at a point above the plane is given by
~ plane,P = σ ẑ.
E
2ε0
−7 2
~ plane,P = (1.6 × 10 C/m2 ) ẑ = 9 × 103 N ẑ
E C C
2(8.85 × 10−12 Nm 2)
Compute the Electric Field from the Point Charge: The electric field of a point charge at point P is by
Coulomb’s Law,
~ point,P = kq
E 2 r̂point,P .
rpoint,P
The displacement vector from the point charge to the point P isp~rpoint,P = (−5cm, 0, 5cm), by
√ observation of
the diagram. The length of this vector is rpoint,P = |~rpoint,P | = (−5cm)2 + 02 + (5cm)2 = 5 2cm. The unit
vector of ~rpoint,P is by definition
~rpoint,P (−5cm, 0, 5cm) 1 1
r̂point,P = = √ = (− √ , 0, √ ).
rpoint,P 5 2cm 2 2
Now, substitute into Coulomb’s Law:
2
9 Nm −8
~ point,P = (8.99 × 10 √C2 )(1 × 10 C) (− √1 , 0, √1 ).
E
(5 2cm)2 2 2
E~ point,P = 1.8 × 104 N − √1 , 0, √1
2 2
~ point,P = (−1.3 × 10 N, 0, 1.3 × 10 N)
E 4 4
Compute the Electric Field of the Line Charge: The electric field of a line charge is given by
~ line,P = λ
E r̂line,P ,
2πε0 rline,P
where ~rline,P is the displacement vector from the line charge to the point P . By observation of our diagram,
√
~rpoint,P = (5cm, 0, 5cm). Using the results of the calculations of the previous step, this gives rline,P = 5 2cm
and r̂line,P = ( √12 , 0, √12 ). Substituting in the formula for the electric field,
Use Linear Superposition to Add the Fields: The total electric field at point P is the sum of the fields,
~P = E
E ~ point,P + E
~ line,P + E
~ plane,P .
~ P = (−1.3 × 104 N, 0, 1.3 × 104 N) + (8.99 × 104 N , 0, 8.99 × 104 N ) + (0, 0, 9 × 103 N )
E
C C C
~ P = (7.7 × 104 N , 0, 11.2 × 104 N )
E
C C
Mechanics
Newton’s Laws Apply to the Electric Force: Just like any other force, we can use
Newton’s three laws on the electric force.
Newton’s First Law: An object continues in its initial state of rest or uniform motion
unless it is acted on by an unbalanced or net force.
Deduction I from Newton’s First Law: If an object is at rest, the net force on the
object is zero.
Deduction II from Newton’s First Law: If an object is moving in a straight line with
a constant speed, the net force on the object is zero.
Deduction III from Newton’s First Law: If an object is turning or changing speed,
there is a net force on the object.
What is Net Force?: The net force is the vector sum of all the forces, F~i ,
F~net = F~i .
X
Newton’s Second Law: The acceleration, ~a, and net force, F~net on an object are
related by
F~net = F~i = m~a,
X
Newton’s Third Law: (Equal and Opposite Forces) If object X exerts a force F~ on
object Y, then object Y exerts a force −F~ on object X.
70
6.2. MECHANICS IN AN ELECTRIC FIELD CHAPTER 6. MECHANICS
Solution
Since the pith ball floats, it is not accelerating. This means that
the total force on the pith ball is zero. The forces acting on the
pith ball are: the force of gravity F~g and the electric force F~e .
z
Thus, by Newton’s First Law, F~e + F~g = 0.The force of gravity Fe
is F~g = −mgẑ with ẑ pointing upward. Therefore, F~e = −F~g = _ __
mgẑ. __
m
F~e = (0.0001kg)(9.8 2 ) = 9.8 × 10−4 N
s tube
Fg
_ _ _ __ _ _ _ _ _ _ _ _ _ __
_ _ _ _ __ __ _ _ __ _ ___
Velocity for Constant Acceleration: If we have a constant acceleration, ~a, then the
velocity is
~vt = ~v0 + ~at,
where ~vt is the velocity at time t and ~v0 is the velocity at time zero. For motion in one
dimension, along the y-axis for example, this can be simplified to
vt = v0 + ay t,
where vt is the velocity at time t in the ŷ direction and v0 is the velocity in the y
direction at t = 0. Be careful here, ay , vt and v0 are all “signed” numbers, not
magnitudes, positive if they are in the direction defined as positive in your coordinate
system, negative if they are in the opposite direction.
Position Equation for Constant Acceleration: The position after time t of a particle
moving with constant acceleration ay along the y-axis is
1
yt = y0 + v0 t + ay t2 ,
2
where yt is the position at time t, y0 is the position at time 0, and v0 is the velocity at
time 0. Again, ay and v0 are “signed” numbers, and so are yt and y0 .
Time of Flight Equation for Constant Acceleration: The time, ∆t, to travel a
distance, d, if v0 = 0 is, solving the above equation,
s
2d
∆t = .
|a|
We can take an absolute value here, because if you start out with v0 = 0 then the
displacement and the acceleration will have the same sign. If they have the same sign,
then you can just divide the magnitudes, and remember, distance is the magnitude of
the displacement, as long as the direction did not change during the motion.
Solution
The net force must be zero, since neither the magnitude of the velocity nor the direction of the velocity is changing,
by Newton’s First Law.
Definitions
y
λ
Use Newton’s First Law to Balance Forces: Since the charge is motionless, by Newton’s First Law
F~g + F~E = 0.
The force of gravity is F~g = −mg ŷ. Solving for the electric force gives
m
F~E = −F~g = mg ŷ = (1 × 10−3 kg)(9.81 2 ) = 9.81 × 10−3 Nŷ
s
(a) Electric Field of an Infinite Line of Charge: The electric field of the line charge is
~ = λ
E r̂
2πǫ0 r
(b) Find the Charge Density: By definition of electric field, the electric force is F~E = q E.
~ Substitute the
formula for the field,
qλ
F~E = +mg ŷ = q E
~ = ŷ
2πǫ0 d
where d = 10cm. Solving for λ and cancelling the vectors gives,
2πdǫ0 mg
λ= .
q
Solve 2
C
2π(5 × 10−2 m)(8.85 × 10−12 Nm 2 )(9.81 × 10
−3
N)
λ= = 1.36 × 10−8 C/m.
2 × 10−6 C
~ = |F~E /q|
|E|
Substitute the force we found,
which is an order of magnitude bigger than the Earth’s electric field of 150N/C.
Solution to (a)
Solution to (b)
The acceleration of a proton by Newton’s Second Law is ~a = F~p /mp where F~p is the net force on the proton. If
~ then F~ = eE
the force is provided by an electric field E, ~ by definition of the electric field, therefore
e ~ C N m
~a = E = 9.58 × 107 (100 î) = 9.58 × 109 2 î
mp kg C s
Solution to (c)
Since the electric field is constant, and therefore the acceleration is constant we can use v = |~a|t if the proton
starts from rest, where v is the velocity at time t. The time, t0.01c to reach v = 0.01c is
v (0.01)(3 × 108 ms )
t0.01c = = = 3.1 × 10−4 s
|~a| 9.58 × 109 sm2
Solution to (d)
An electron is far lighter than a proton, so from the equation in part (b), the acceleration would be far greater,
and therefore the time to reach 0.01c far less.
For the rest of this class, the field map will be used as the primary representation of both electric and magnetic
fields. This chapter develops the skills needed to draw these very useful diagrams.
Gauss’ Law (Version 0): The net number of electric field lines exiting any closed
surface is proportional to the charge enclosed in that surface.
Q ∝ lines.
Definition of Open and Closed Surfaces: A closed surface will, metaphorically, hold
water. An open surface will not. A filled balloon is a closed surface, but a popped
balloon is an open surface. A shoe box with the lid on is a closed surface. A shoe box
with the lid off is an open surface.
Definition of Field Line: A field line is a graphical object, a representation of the
electric field. IT IS NOT A REAL PHYSICAL THING. A field line is a line drawn such
that the electric field is tangent to the line (pointing in the direction of the line) at all
points along the line. Therefore, if you are given a field line you should be able to draw
the direction of the electric field vector at each point along the line.
75
7.1. GAUSS’ LAW, FIELD LINES, AND FIELD MAPS CHAPTER 7. ELECTRIC FIELD MAPS
(a)At three points on each line, draw the electric field vector.
(b)For three of the electric field vectors drawn, draw the direction
of the electric force on an electron. Draw the acceleration of an
electron at one of the electric force vectors.
Solution to Part(a)
The electric field vectors are at every point tangent (point in the same direction
as) the electric field lines. Some of the field vectors are drawn slightly off the
field lines so they can be seen.
Solution to Part(b)
E
F
F E
F a
Gauss’ Law states that the net number of field lines leaving a closed surface is proportional to the net charge
in the surface, but it looks like I can draw field lines anywhere. To draw a representation of the electric field to
which Gauss’ law can be applied an additional constraint on the drawing of field lines must be imposed.
Definition Electric Field Map: An electric field map is a representation of the electric
field using field lines, where
• The number of field lines leaving or entering any charged object is consistent with
Gauss’ Law.
• The distance between field lines is inversely proportional to the strength of the
electric field.
Electric fields occupy three dimensional space, but our field maps are two dimensional, which is hard enough
to draw. Therefore, we are projecting a three dimensional image onto a two dimensional drawing. We introduce
errors by doing this. Our field maps should correctly be thought of as extending uniformly into and out of the
paper. We ignore this approximation because the field maps we draw have the correct shape and all reasoning
proceeds correctly using the two dimensional image. When a field map of a point charge or spherical system is
presented, it is actually the field map of a cylindrical system extending infinitely into the page. If this is unclear,
ignore it, because everything will work out beautifully. (The upshot is, you cannot use field maps to determine the
exact ratio of point charge field strengths at different distances by measuring the separation of the lines, because
you drew 3 dimensions worth of lines in only 2 dimensions.)
The electric field map we have been working with is drawn for a region of space
containing no net charge. How do I know? Gauss’ Law states that the charge
enclosed in any closed surface is proportional to the electric field lines exiting
the surface. A closed surface (dashed line) in drawn on the figure to the right. I
count four field lines entering the surface and four field lines exiting the surface,
therefore zero net field lines exit the surface and by Gauss’ Law the net charge
enclosed is zero. Anywhere I draw a surface on the diagram to the right, I
will find zero net field lines exiting the surface, therefore there is no net charge
anywhere in the region of space where this field is drawn. Cool! We can do
physics just by counting lines.
I hope as you considered moving the Gaussian surface around the field map,
you started to wonder what a field map containing net charge would look like.
Simple, to produce a closed surface with net lines entering or exiting the surface,
field lines must start or end within the surface. The surface to the right shows
net lines entering the surface. Gauss’ Law states that the net field lines EXITING
a surface is proportional to the net charge enclosed in the surface. If net lines
exit, the charge enclosed is positive; if net lines enter, then the charge enclosed
is negative. In the figure at the right, net field lines enter the surface (the lines
point inward), therefore the charge enclosed is negative. In this way, we can
use Gauss’ law to locate charge in a field map.
Lines Begin on Positive Charge and End on Negative Charge: Electric field lines
begin on positive charges and end on negative charges or infinity.
Field Lines Do Not Cross: If two field lines cross, then the electric field would have
two different values at the same point in space. That can’t happen.
Field maps are drawn so that the distance between field lines is inversely proportional to the magnitude of the
electric field. This means that the relative magnitudes of the electric field at different places in space can be read
off the field map by comparing the separation of field lines. We can’t get the exact ratio, though, since we are
drawing 3 dimensions worth of lines in 2 dimensions.
A
Q EA
B
EB
Solution to Part(a)
Far from a charge distribution, the field lines become radial (point straight outward), and behave as if coming
from a point charge with the total charge of the distribution. Since field lines enter C, it must contain a net
negative charge. We are given the magnitude of the total charge as |Q|, so the total charge must be −|Q|. Let
this be written as −Q.
Solution to Part(b)
Four field lines enter C and eight field lines enter the surface A. Since field lines enter the surface, the charge
enclosed in A is negative. The number of field lines entering or leaving a surface is proportional to the charge.
Since twice as many lines enter A as enter C, the charge enclosed by A is twice the charge enclosed by C.
Therefore, the charge enclosed by surface A is −2Q.
Solution to Part(c)
Four field lines enter C and four field lines exit the surface B. Since field lines exit the surface, the charge enclosed
in B is positive. Since the same number of lines leave surface B as enter surface C, the charge enclosed by B is
negative the charge enclosed by C. Therefore, the charge enclosed by surface B is Q.
Solution to Part(d)
The electric field points in the direction of the field lines. For D the field lines point to the left, so the electric
~ D , is directed to the left. For E the field lines point to the right, so the electric field, E
field, E ~ E , points to the
right. Notice I had to interpolate between field lines to get the direction at these points. The magnitude of the
electric field at point D is larger than the magnitude of the electric field at point E because the field lines are
closer together at D, |E ~ D | > |E
~ E |.
Correct Incorrect
Everything is Radial Far From a Net Charge: At large distances from any non-zero
charge, the lines are equally spaced and radial just as they would be for a point charge
with Q = total charge of the system.
Select a Number of Lines Per Charge: The number of field lines entering or leaving
a charged object is proportional to the charge of the object. If we have point charges
q1 = 5µC, q2 = −10µC, I might randomly select four lines to represent q1 , therefore
eight lines represent q2 .
Draw Stubs of Field Lines: Draw little arrows on the charges for the number of lines
selected. Arrows should point out for positive charge and in for negative. Near a point
charge the field lines are radial, since E → ∞ as r → 0. This means that the field line
stubs should be evenly spaced around the charge.
Draw the Long Range Field: Far from a charge distribution, the electric field will
have a characteristic shape. For distributions with a non-zero net charge, the electric
field far from the charges will be that of a point charge with a charge equal to the total
charge of the distribution. If we continue with q1 and q2 above, far from the charge we
will see a radial field equal to that of a point charge with charge qt = q1 + q2 = −5µC.
Draw a dashed circle far from the charges, which is called the circle at infinity. Draw
the appropriate number of field lines leaving or entering this circle. For q1 and q2 , if
four lines leave q1 = 5µC, then four other lines must enter the circle at infinity since
qt = −5µC.
Connect the Stubs Without Crossing the Field Lines: Field lines do not cross, since
the electric field has a single direction at every point in space. (As Egon said, ”Do not
cross the streams. . . It would be bad.”) To finish the map, simply connect the field lines
on the stubs and the field lines at infinity, without crossing the lines. A line may not
begin and end with stubs pointing in different directions.
Respect Symmetry: The symmetry of your field map is affected by the initial choice
of stub directions, your choices for the field at infinity, and how you connect the stubs.
The field you end up with should have the same symmetry as the charges you started
with.
Let’s apply this process to the drawing of the electric field of one and two point charges.
Example 7.3 Drawing the Electric Field Map of One Point Charge
Problem: Draw the electric field map of one point charge : +Q.
Solution
(a) Draw and clearly label location of all charge: Compute the total charge of each object. Draw each
object and label it with its total charge. In this case we have only one charge which we will label Q.
(b) Select a number of lines for a certain amount of charge: Select a number so that a minimum of 2 lines
(preferably at least 4) are associated with an object. Use this to associate a number of lines proportional to the
charge with each object. For this problem use Q = 8 lines.
(c) Draw Stubs in the Direction of the Field Lines: Draw stubs for the number
of field lines. Q = 8 field lines evenly spaced around the charge. The electric field is
the electric force divided by the charge. So if I put another +q near the +Q charge it
will be pushed outward. [Like charges repel] So for the +Q charge, the field lines point
away from the positive charge. +Q
(d) Compute total charge of system and draw the field at ∞: Draw a circle at infinity. Compute the
number of lines for the total charge and draw stubs with direction arrows on the circle at infinity. For this case,
the total charge is Q so there are 8 stubs pointing outward.
+Q
(e) Connect the lines: Connect the lines coming off of the charge to the circle at infinity. Bend things smoothly
so that each positive line emerging from a positive stub ends on a negative stub or on a positive stub on the circle
at infinity. Do not cross the lines, remember that electric field lines do not cross because this would mean that
the electric field had two values at one point in space.
+Q
(f) Shake it up: Either the drawing will look great or weird. If it looks great, you’re finished. Congratulations.
If it looks weird, you need to rotate the stubs you initially drew and redraw until it doesn’t look weird.
Solution
(a) Draw Charges and Field Stubs: Draw the locations of the charges and
label their strengths. Select a number of field lines per charge. I pick Q ∝ 8
field lines. Field lines exit a positive charge. Draw the field line stubs exiting
the charge.
+Q +Q
(b) Draw the field far from the charges: We have a total
charge of 2Q, so we need 16 field lines at ∞.
Q Q
Q Q
(d) Respect the Symmetry: The charges are the same, there-
fore the field lines come out from each charge in the same way.
The field stubs I chose did not create an appropriately symmetric
picture, so the stubs had to be rotated. After rotating the field
stubs, the correct figure is drawn to the right.
+Q +Q
(a) Draw the system: The system of charge is drawn to the right.
+ +
+ +
_ _
_ _
_ _
_ _
(c) Connect Everything Up: I had to shorten some of my stubs to get this one to look right. Notice, since
the net charge is zero, no lines exit to infinity. This is a dipole system which will be covered in more detail next
chapter.
+ +
+ +
_ _
_ _
Reason About Field Strength: As we draw more complicated field maps, it becomes
more difficult to decide how to connect the field lines. This means we have to begin to
reason about where the field is strong and weak to get a good field map. The best way
to do this is to draw the field vectors at a few points.
Solution
+ +
+ +
(d) Connect the Lines Choosing Different Stubs for the Cen-
tral Charge: Reconnect the lines with a small change in what
stub goes where, giving a field map that has a weaker field in-
side the half-circle. Note, the field spacing is consistent with the
vectors we drew.
+ +
A
+ +
Electric Dipoles
It requires a lot of energy to produce a net charge, so most objects do not have a net charge. All objects
however contain charge and often the centers of positive and negative charge in an object are at different locations.
The behavior and shape of the electric field of these systems is determined by their electric dipole moment.
88
8.1. BEHAVIOR OF ELECTRIC DIPOLES CHAPTER 8. ELECTRIC DIPOLES
Moments of the Electric Field: Any electric field can be expressed as a series of
characteristic fields whose strength is determined by their “moment”. The net charge
of a system is the system’s monopole moment. The next moment is the dipole moment,
defined below. Higher order moments exit: quadrapole, octopole, etc. The long range
shape of the field is determined by the lowest order non-zero moment. The long range
shape of a system with net charge is radial, determined by its monopole moment. If
monopole moment is zero and the dipole moment non-zero, the long range shape is
dipole.
Definition Dipole Moment Vector: The dipole moment vector for a system with zero
net charge, p~, can be calculated for a collection of charges qi located at the points ~ri
using X
p~ = qi~ri
i
Solution
The dipole moment vector is by definition
X
p~ = qi~ri = (2nC)(0, 0, 0) + (−1nC)(1cm, 0, 0) + (−1nC)(1cm, 1cm, 0)
i
Dipole Moment for Equal and Opposite Charges: For a dipole formed of two equal
and opposite point charges, the dipole moment points from the negative charge to the
positive charge and has magnitude p = qd where d is the separation between the charges
and q is the charge of the positive charge.
Direction of the Dipole Moment Vector: The dipole moment vector points from
the center of the negative charge to the center of the positive charge of the charge
distribution.
The strength of the dipole, the size of |p|, increases with the amount of charge separated, q, and the amount of
separation, d, as illustrated below.
p p
p
Small Amount of
Charge Separation - More Charge Separated, Less Separation,
Dipole Moment Larger Dipole Moment Smaller
Small Dipole Moment
Far from the charges, all charge distributions with zero total charge but non-zero dipole moment have the
characteristic dipole electric field. If you see a dipole field, you should be able to draw the dipole moment and
tell me that the total charge is zero.
p
p
The mathematical form of the electric field for a dipole, far from the dipole, is somewhat complicated. We
state it for your reference,
Electric Dipole Field: The electric field of a point dipole at the origin with dipole
moment p~ is
~ r) = k 3r̂(~
E(~
p · r̂) − p~
.
r3
This is the field of a point dipole or the field far from a system with zero charge but
non-zero dipole moment.
Simplified Electric Dipole Field: The expression above for the dipole field can be
simplified if a direction for the dipole moment is chosen and only the strength of the
field along the axes is computed. If p~ = pŷ, then along the ŷ axis,
~ y, 0) = 2kpŷ .
E(0,
|y|3
Notice that the strength of the dipole field falls off as 1/r3 whereas the field of a distribution with net charge
falls off as 1/r2 . This is why, far from a distribution with net charge, we see only the radial field of a point charge
with the total charge of the distribution.
Solution
(a) Calculate the dipole moment: The dipole moment of a simple two charge dipole is p = qd where
d = 0.4cm is the separation of the charges p = qd = (1 × 10−9 C)(4 × 10−3 m) = 4 × 10−12 Cm.
(b) Calculate the field along the x-axis: The field point at 5m along the x-axis is far from the charges, so
the formula for the long range dipole field can be used
9 Nm2
~ = − kpŷ = − (8.99 × 10 C2 )(4 × 10−12 Cm)ŷ N
E = −2.88 × 10−4 ŷ
|x|3 |5m|3 C
(c) Calculate the field in the y direction: The field point at 5m along the y-axis is far from the charges, so
the formula for the long range dipole field can be used
2
9 Nm −12
~ = 2kpŷ = 2(8.99 × 10 C2 )(4 × 10 Cm)ŷ = 5.75 × 10−4 N
E
|y|3 |5m|3 C
Determine Direction of Dipole Moment: The dipole moment is directed from the
center of negative charge to the center of positive charge. Draw it on your figure.
Draw a Dipole Long Range Field: Draw the circle at infinity and draw a dipole field
matching your dipole moment.
Solution
(a) Draw the Charges: Draw the electric charges at the given locations to
scale. Since we are given an electric dipole along the y-axis, draw equal and
opposite charges along the y-axis.
y
x
(b) Draw the Dipole Moment: The dipole moment vector is drawn from
the center of negative charge to the center of positive charge. For two point
charges, the electric dipole is drawn from the negative to the positive charge.
y
x
(c) Draw the Long Range Dipole Field: For a charge distribution that has zero net charge and a non-zero
dipole moment, the electric field far from the charge has the characteristic shape of an electric dipole.
(d) Draw Stubs of Field Lines: I chose eight lines per charge. The field lines exit at the positive charge and
enter at the negative.
_ p
Our barbell dipole is placed in a number of electric fields below. The force on each charge, F~+ and F~− , and
the net force, F~net = F~+ + F~− , on the dipole is drawn in each case.
Figure (a) Equilibrium Figure (b) Away from Equilibrium Figure (c) Non-Uniform Field
F+
+ F+ F+
+ +
p p p
Fnet
_ _ _
F− F− F−
In figure (a) the field is uniform and the dipole moment p~ aligns with the field. The net force is zero and the
forces on the dipole do not tend to rotate the dipole. This is the equilibrium position of the dipole. In figure (b),
the dipole is rotated away from equilibrium. The net force is still zero, but the force on each charge tends to
rotate the dipole toward equilibrium.
Dipoles Rotate to Align with Field: A dipole placed in an electric field is at equilib-
rium when the dipole moment points in the same direction as the field line. A dipole
that is not at equilibrium will tend to rotate toward alignment with the field line.
Dipoles In a Uniform Field Feel Zero Net Force: If a dipole is placed in a uniform
electric field, constant through space, then the total force (but not the torque) is zero
since the forces on the plus and minus charge are equal and opposite. So the dipole
rotates but its center of mass stays in the same place.
In figure (c), the field is not uniform. The positive and negative charges forming the dipole experience forces
of different magnitudes and directions and therefore there is a net force on dipole.
Net Force on Dipoles In a Non-Uniform Field: If a dipole is placed in a non-uniform
field, the two charges experience difference forces, and the direction of the net force
must be determined by adding these forces.
The field lines are evenly spaced since we are told the field is uniform. The dipole moment, in the +ŷ direction
here, always points from the negative to the positive charge. See figure.
Solution to Part (b)
The force on the positive charge will point the same direction as the field; the force on the negative charge will
point in the opposite direction. The forces have the same magnitude since the field is uniform. See figure.
Solution to Part (c)
F+
+
p Rotation
F−
−
+ θ ∆h
+
p
d p
_
_ ∆h
The difference in potential energy, ∆U , from figure (a) to figure (b) is the work, W , an external agent would
have to do to rotate the dipole. Work is force times the distance in the direction of the force. Both the positive and
negative charge moved a distance ∆h against the force of field. The work done is W = F+ ∆h+F− ∆h = 2qE∆h,
where F+ is the force on the positive charge, q is the magnitude of the positive charge, and E is the electric field.
If θ is the angle between the dipole moment vector, p~, and the field E, ~ and d is the length of the dipole, then
∆h = d/2 − d/2 cos θ. Substituting gives the change in potential energy to rotate from figure (a) to figure (b).
d d
∆U = 2qE − cos θ = −pE cos θ − pE
2 2
U = −pE cos θ = −~ ~
p·E
where θ is the angle between the dipole moment and the field. The second expression
uses the vector dot product, which we will review in Course Guide 11.
F+
θ +
p
Moment Arm
_
F−
From UPI, the torque τ exerted on the object is the force multiplied by the moment arm, the perpendicular
distance to the line of action of the force. The moment arm and lines of action are drawn above. The torque
will be calculated about the center of the dipole. The torque on the dipole in the figure causes it to rotate in the
counterclockwise direction. Both forces, F~+ and F~− , exert a torque on the object. The total torque is the sum
of the torques of the two forces, τ = τ+ + τ− . The length of the moment arm is (d/2) sin θ for both forces if the
separation of the charges is d and the angle θ is measured from the dipole moment to the field. The angle θ is
positive above. The force on each charge is qE. The total torque is then τ = qEd sin θ or τ = pE sin θ where I
have used |~p| = qd.
Torque on an Electric Dipole: The torque, τ , on an electric dipole with dipole moment
~ is
vector p~ in an electric field E
τ = pE sin θ
~ A positive torque causes a counterclockwise angular
where θ is measured from p~ to E.
acceleration.
When we reach magnetic dipoles and have some experience with the vector cross product, the above expression
~
will be re-written as ~τ = p~ × E.
x F−
We would like to estimate the force on the dipole in the limit the length of the dipole, d, is small. If the field points
generally in the y direction at the location of the dipole then at the dipole we can write the field E~ = E(y)ŷ.
The net force will point generally in the y direction and have magnitude, Fnet = qE(y+ ) − qE(y− ) where y+ is
the location of the + charge and y− is the location of the − charge. If the separation of charges d is small, then
this is approximately
dE dE dE
Fnet = q (y+ − y− ) = q (d cos θ) = p (cos θ)
dy dy dy
where θ is the angle between the dipole moment and the field and once again I have use |~
p| = dq.
where θ = 135◦ is the angle between the dipole and the field. α
θ
This drawing is way out of scale, a molecule is tiny, so we can +
p
pretend E is in the same direction at either end of the dipole.
The positive sign indicates the force is outward from the golf tube.
Continuous Charges
With Coulomb’s Law, E ~ = kqr̂/r2 , and the Law of Linear Superposition, we can calculate the electric field of
any charge distribution. In this chapter we will use the two laws to calculate the electric field of some continuous
charge distributions.
EjP
riP
si P EiP
rjP
sj
If calculus did not exist, what has to be done to calculate the field is pretty obvious, cut the curve into small
bits of length ∆s, calculate the field of each bit, and add using linear superposition. Let the ith bit be at the
location si along the curve and contribute a field E ~ iP at the point P . The contributions of the ith and jth
segment are drawn above. The charge of each piece is the charge density multiplied by the length of the piece,
qi = λ(si )∆s. The total electric field is the sum of the fields of each piece,
X kqi X kλ(si )∆s
~P =
E r̂iP = r̂iP
2
riP 2
riP
i i
This expression could be summed directly on a computer to any accuracy or converted to an integral using
100
9.2. ONE DIMENSIONAL PROBLEMS CHAPTER 9. CONTINUOUS CHARGES
where now both the distance and the unit vector are functions of the length along the curve.
kλ(s)ds
Z
E~P = r̂P (s)
rP (s)2
To me, the sum looks easy and the integral looks horrible, even though they are exactly the same thing. This is
why I stress thinking in terms of the sum first.
The same process can be carried out for charge spread over a surface. Suppose a surface occupies part of the
x-y plane. The surface is covered with a surface charge density σ(x, y). We wish to calculate the electric field
at a point P at the point ~rP . Imagine dividing the surface into small squares of width ∆x and height ∆y. Let
the center of each square be ~rij = (xi , yj , 0). The displacement vector from one of the squares to point P is
~rijP = ~rP − ~rij . The charge of each square is qij = σ(xi , yj )∆x∆y. The electric field of the surface at point P
is
X X kqij X X kσ(xi , yj )∆x∆y
~P =
E r̂ijP = r̂ijP
2
rijP 2
rijP
i j i j
Those of you who have had Cal III should see the two dimensional integral. A similar expression can be written
for the volume charge distribution.
Alternate Form of Coulomb’s Law: The electric field at point P due to a point
charge q at location ~ri is
~ P = kq r̂iP = kq ~riP
E 2
riP 3
riP
where ~riP = ~rP − ~ri is the displacement vector from the point i to the point P .
(a) Divide the Line Charge into Segments: Cut the line
charge up into segments of length ∆x and center xi as drawn to
the right. The field points in the x̂ direction at R, so we just have
to calculate the magnitude of the electric field of each segment,
Ei = kqi /d2i , where di is the distance from the center of the
segment to the point R. Observing the diagram to the right, -L ∆x L
di = R − xi . The charge of the segment is λ∆x. So using linear xi R x
superposition, di
X kqi X kλ∆x y
E(R) = =
i
d2i i
(R − xi )2
(b) Convert Sum into Integral and Do the Integral: Let the segments become infinitely small, so ∆x → dx,
P RL
xi → x, and i → −L
Z L
dx
E(R) = kλ
−L (R − x)2
Use the u - substitution, u = R − x, so du = −dx,
R−L
du
Z
E(R) = −kλ
R+L u2
r n+1
rn dr = r−2 dr = − 1r + C.
R R
This is an integral of the form, n+1 + C with n = −2, so
R−L
1 1 1
E(R) = kλ = kλ −
u R+L R−L R+L
If you got here, you would get full credit on homework or an exam. I’m going to simplify a bit,
R+L R−L R + L − (R − L) 2kLλ
E(R) = kλ − = kλ =
(R − L)(R + L) (R + L)(R − L) (R2 − L2 ) (R2 − L2 )
Note, we get the field of a point charge if R → ∞, E(R) → 2kλL/R2 = kQ/R2 . Be sure to put the vector back
~ = E(R)x̂.
in at the end, E
Solution
R riP
xi
0 ∆x L x
(b) Calculate the charge of the ith segment: The charge, qi , is the length of the segment multiplied by the
charge density, qi = λ(xi )∆x = γxi ∆x.
(c) Calculate the Displacement Vector: The displacement vector is by definition, ~riP = ~rP −~ri . The position
of the point P is ~rP = (0, R, 0). The position of the segment i is ~ri = (xi , 0, 0), so the displacement vector is
~riP = (−xi , R, 0), which is consistent with our drawing. p
(d) Calculate the length of the displacement vector: The length of ~riP is riP = x2i + R2 .
(e) Use Linear Superposition: The total electric field at point P is the sum of the fields of all the segments
X kqi X kγxi ∆x
~P =
E ~riP = 3 (−xi , R, 0)
3
riP
i i
(x2i + R2 ) 2
P RL
(f) Convert the Sum to an Integral: Let ∆x ⇒ dx, xi ⇒ x, and i ⇒ 0
,
~P =
X kγxi ∆x
E 3 (−xi , R, 0)
i
(x2i + R2 ) 2
L
kγxdx
Z
= 3 (−x, R, 0)
0 (x2 + R2 ) 2
This is actually three integrals, one each for the x, y, and z component.
Z L Z L
kγx2 dx
~P = − kγxRdx
E , , 0
2 2 3 2 2 3
0 (x + R ) 2 0 (x + R ) 2
and
L
xdx
Z
EP y = kγR 3
0 (x2 + R2 ) 2
(h) Try Calculus: I typed x/(xˆ2 + Rˆ2)ˆ(3/2) into integrals.com and it told me
xdx 1
Z
3 = −√ +C
2 2
(x + R ) 2 x + R2
2
x2 dx x
Z p
3 = −√ + ln(x + x2 + R2 ) + C
(x2 + R2 ) 2 x2 + R2
where C is the constant of integration. The Log reported at the web site is actually the natural logarithm, which
we will denote as ln.
(i) Use the integrals: Substituting the limits 0 and L,
L L
x2 dx
x
Z p
EP x = −kγ 3 = −kγ −√ + ln(x + x2 + R2 )
(x2 + R2 ) 2 2
x +R 2
0 0
L p
EP x = −kγ − √ + ln(L + L2 + R2 ) − ln(R)
L + R2
2
Z L L
xdx 1
EP y = kγR 3 = kγR − √
2
0 (x + R ) 2
2 x2 + R2 0
1 1
EP y = kγR − √ +
L2 + R2 R
Symmetry
Plane Slab
• Cylindrical Symmetry—A system with cylindrical symmetry is unchanged when it is rotated about its axis
or translated down its axis. We will build systems of cylindrical symmetry out of infinitely long straight lines
of charge; thin, infinitely long cylindrical shells with uniform surface charge; infinitely long tubes of charge
with uniform volume charge density; and infinitely long thick cylindrical shells with uniform volume charge
density.
105
10.2. THE SHAPE OF A HIGH SYMMETRY FIELD CHAPTER 10. SYMMETRY
Cylindrical Symmetry
shell
line tube
thick shell
• Spherical Symmetry—A system with spherical symmetry is left unchanged by any rotation about its center.
We will build spherically symmetric systems of charge out of concentric point charges, thin spherical shells
of charge with uniform surface charge density, and spherical volumes of charge as well as thick spherical
shells with uniform volume charge density.
Spherical Symmetry
point sphere
charge
thin shell
thick shell
Problem: Draw the electric field map of a −1µC point charge surrounded by a spherical shell of charge with
surface density σ = 130 µC
m2 and radius 5cm concentric with the point charge.
Solution
(a) Compute the Total Charge of Each Object: The total charge of the point charge is given as qpoint =
−1µC. The total charge of the shell is
2
qshell = 4πrshell σ = 4π(0.05m)2 (130µC/m2 ) = 4µC.
(b) Select a Number of Lines per Charge: Gauss’ Law states that the number of field lines exiting a region
is proportional to the total charge inside the region. The first step in drawing a continuous field map is the same
as the first step of drawing the field map for a system of point charges, select a number of lines per charge. I
select 4 lines per 1µC.
Air
Air
Gaussian Surface
Gaussian Surface
(f) Draw the charges: Draw positive charges where lines begin and negative charge where lines end. If all went
well you drew a number of charges proportional to the total charge of the object. The charges are drawn above.
Note there are 16 + charges drawn which is correct for 4µC.
Drawing a field map for cylindrically symmetric distributions of charge is the same as for spherical distributions
except that, rather than allocating field lines based on total charge, field lines are allocated based on charge per
unit length.
Gauss’ Law
In this chapter, we learn how to calculate the electric field of the infinite line and infinite plane of charge using
Gauss’ Law. Course Guide 7 and Course Guide 10 applied Gauss’ law qualitatively, now its time to do the math.
~·B
A ~ = Ax Bx + Ay By + Az Bz .
~·B
Angle Form of Dot Product: The dot product A ~ can also be computed as:
~·B
A ~ = |A||B| cos(θ),
~ to B.
where θ is the angle from A ~
The angle form of the dot product allows easy computation of two important special cases, parallel vectors and
perpendicular vectors. If two vectors are perpendicular, the angle between the vectors is 90◦ , cos 90◦ = 0, so the
dot product is zero. If two vectors are parallel, that is if they point in the same direction, the angle between the
~ ·B
vectors is zero, cos 0 = 1, giving A ~ = AB. If one of the vectors is a unit vector, which by definition has length
~ · B̂ = A. If the vectors are anti-parallel, point in opposite directions, then A
1, A ~·B ~ = −AB
110
11.1. ELECTRIC FLUX CHAPTER 11. GAUSS’ LAW
φe = EA cos θ
where θ is the angle between the electric field and the normal to the surface.
Definition Surface Normal:
A surface normal is a vector
which points straight out from
the surface; a vector that is n
perpendicular to the surface.
A surface has two sides and
a normal vector for each side n
n
at every point on the surface.
The symbol n̂ is used to de-
n
note the normal to the surface.
Normals of Simple Surfaces: A sphere, centered at the origin, has outward surface
normal r̂ where the radius vector ~r = (x, y, z). A cylinder centered on the z axis has
outward surface normal r̂ where ~r = (x, y, 0) A surface in the x − y plane has two
surface normals, ẑ and −ẑ.
The electric flux through a surface is proportional to the number of field lines crossing the surface. Let’s examine
a flat surface in a uniform field. A side view of a square loop is shown below. As you can see, the number of
field lines passing through the surface depends on the area of the surface and the angle the surface makes with
the field.
n n θ
Solution
~
Since the field is parallel to the normal of the hoop, the flux is φe = |E|A. The area of the hoop is A = πr2 ,
~ 2 N N
φe = |E|πr = 150 π(0.5m)2 = 117.8 m2 .
C C
The electric flux can be defined for surfaces that aren’t flat and fields that aren’t uniform.
where n̂ is the outward normal to the surface and dA is an element of area of the
surface.
If the electric field is uniform and the surface is flat, the electric flux can be re-written
Z Z
φe = (E ~ · n̂)dA = (E ~ · n̂) dA = (E ~ · n̂)A
S S
where A is the area of the surface. Using the properties of the dot product, we can re-write this as
~
φe = |E||n̂|A cos θ = EA cos θ
since |n̂| = 1. The angle θ is the angle between the surface normal and the electric field. Therefore, we recover
our definition of electric flux for a flat loop.
P
θ
2
Top View Area= cos(θ)
2 Area=0
Area=
The rain drops per unit time falling into the bucket, the flux of drops through the surface bounded by the
top of the bucket, is proportional to the area of the top of the bucket projected on the ground. In figure (a), the
area projected on the ground is A = ℓ2 and the drops per unit time entering the bucked is F A. In figure (b), the
bucket is tipped and the projected area is ℓ2 cos θ = A cos θ, and the drops per unit time entering the bucked is
F A cos θ. In figure (c), the projected area is zero and no drops enter the bucket. If θ is the angle the normal to
the surface bounded by the top of the bucket makes with the direction of the falling drops then all three cases can
be written F A cos θ. If electric field replaces the flow of raindrops, then the flux of electric field into the bucket
would be EA cos θ, exactly the mathematical expression for flux in a uniform field.
Anytime you are dealing with an integral you can’t quite visualize, imaging doing it as a sum over small chunks.
~ · n̂)dA = Qenclosed
Z
φe = (E
S ǫ0
Now, consider how you would do this integral without Gauss’ Law. The second question is, why would you
want to do this integral at all, and I can’t think of a single reason. For actual uses of Gauss’ Law, we need a little
help from symmetry.
Spherical
Charge
Distribution
Gaussian Surface
The electric flux through the Gaussian surface is by definition of electric flux:
Z
φe = ~ · r̂)dA.
(E
S
The electric field has the same magnitude at all points on the surface S, by symmetry, and can be brought out
of the integral Z
φe = E(r) dA.
S
φe = 4πr2 E(r)
+
+
The surface area of a cylinder of length L and radius r excluding the ends is 2πrL. Apply Gauss’ Law,
Qenclosed
φe = 2πrLE(r) =
ε0
Gauss’ Law from Cylindrical Systems: Gauss’ law applied to a cylindrical surface of
radius r with length L for a system of charge with cylindrical symmetry reduces to
Qenclosed
φe = 2πrLE(r) =
ε0
+ nr
nl
+
x Plane Charge
Break the surface integral into an integral over the left end, the right end, and the sides.
Z Z
φe = ~
(El · n̂l )dA + (E~ r · n̂r )dA
lef t right
Z
+ ~ · n̂sides )dA
(E
sides
where E ~ l is the electric field at the left end of the cylinder and E ~ r is the electric field at the right end. The
integral over the sidesR is zero because the electric field is perpendicular to the surface normal at all times, so the
dot product is zero. sides (E ~ · n̂r )dA = 0. Let E
~ l = El x̂ and E ~ r = Er x̂. From the diagram, we can see n̂l = −x̂,
n̂r = x̂. Substitute this into Gauss’ Law,
Z Z Z Z
φe = El (x̂ · (−x̂))dA + Er (x̂ · x̂)dA = −El dA + Er dA
lef t right lef t right
since x̂ · x̂ = 1. By symmetry, the field is constant on the ends of the Gaussian cylinder and can be removed from
the integral, Z Z
φe = −El dA + Er dA
lef t right
The integrals are simply the areas of the ends of the cylinder.
φe = −El A + Er A
Definition of Gaussian Surface: A Gaussian surface is surface where the electric field
is constant and parallel to the surface normal or perpendicular to the surface at all
points on the surface.
If we let the area of the Gaussian surface with non-zero flux be Ag , then Gauss’ law assumes the much simpler
form φe = EAg = Qenc /ε0 . For example, in spherical systems we use a spherical Gaussian surface of radius r,
the area of the surface is Ag = 4πr2 and the specialized form of Gauss’ law is φe = EAg = 4πr2 E = Qenc /ε0 .
The same expression we worked so hard to extract with vector methods. For cylindrical systems where the surface
has radius r and length L, the area is Ag = 2πrL, and Gauss’ law becomes φe = EAg = 2πrLE = Qenc /ε0 . For
planar systems, we have to take care of the fact the flux is inward at the left side of the cylinder and has a negative
sign, −El A, and outward at the right end of the cylinder, Er A, giving Gauss’ law as −El A + Er A = Qenc /ε0 .
Gauss’ Law Applied to a Gaussian Surface: If the area of the Gaussian surface with
non-zero flux is Ag , then Gauss’ law becomes
Qenc
φe = EAg = .
ε0
Qenclosed
4πr2 E(r) =
ε0
Solving for E and writing the field as a vector yields
~ r) = Qenclosed r̂
E(~
4πε0 r2
Notice, we had to add the direction of the field, r̂, back in to write the full electric field. We had to know the
direction of the field to choose our Gaussian surface to begin with.
We should start by applying Gauss’ law to a point charge of charge q. If we don’t recover Coulomb’s law we’re
dead. For this case, Qenc = q for any radius surface and the electric field is
~ r) = Qenc r̂ =
E(~
q kq
r̂ = 2 r̂
4πε0 r 2 4πε0 r 2 r
where I have used k = 1/4πε0 . Therefore, Coulomb’s law and Gauss’ law are equivalent.
Gauss’ law states that the electric flux out of a closed surface is related to the charge enclosed in that surface by
Qenc
φe =
ε0
~ = E(r)r̂, this can be converted to
The Gaussian surface is drawn below. For spherical symmetry, since E
Qenc
φe = 4πr2 E(r) =
ε0
where r is the radius of the Gaussian surface. Solving for the field gives
~ = Qenc r̂
E
4πr2 ε0
For a Gaussian surface with radius r < a so that the outer edge of the surface lies in region I, the surface encloses
the point charge and part of the volume charge: The charge enclosed in the surface is therefore
4
Qenc = +2Q + πr3 ρ
3
therefore the electric field in region I is
4 3
~ I = Qtotal r̂ = 2Q + 3 πr ρ r̂
E
4πε0 r2 4πr2 ε0
An example of a Gaussian surface in region II is drawn below. For a Gaussian surface with radius a < r < b in
region II the charge enclosed is the point charge and all of the volume charge therefore:
4
Qenc = +2Q + πa3 ρ
3
therefore the electric field in region II is
4 3
~ II = Qtotal r̂ = 2Q + 3 πa ρ r̂
E
4πε0 r2 4πr2 ε0
A spherical Gaussian surface with radius b < r always encloses all the volume charge, the point charge, and the
shell of charge, therefore for region III:
4
Qenc = +2Q + πa3 ρ + 4πb2 σ
3
therefore the electric field in region III is
4 3 2
~ III = Qtotal r̂ = 2Q + 3 πa ρ + 4πb σ r̂
E
4πε0 r 2 2
4πr ε0
_ I
++2Q+ _
+ +
volume
charge
_ II
b _
+ +
_ _
III _
The Gaussian surface is a sphere so its outward surface normal is the vector r̂.
Introduce an Arbitrary Area: Let the area of the top and the bottom of the cylinder
be A. This should cancel out of the final calculation.
Gauss’ law applied to the cylindrical surface becomes Eright A − Elef t A = Qenc /ε0 . If you try to apply this
blindly, you will find you are one equation short. You need the following interesting application of symmetry to
actually do a planar problem.
Outer Fields are Equal and Opposite in a Planar System: The electric field for a
planar system does not fall off with distance, therefore the magnitude of the field to
the far right is always equal and in opposite direction to the field to the far left.
(b) Select and Draw the Gaussian Surface: For planes, the appropriate Gaussian surface is a cylinder whose
left and right faces are parallel to the planes. Let the cylinder have end area A.
(c) Select the Appropriate Form of Gauss’ Law for Planar Symmetry: For a planar system, the fields have
the form E ~ I = EI x̂, E
~ II = EII x̂, and E
~ III = EIII x̂. The appropriate form of Gauss’ law is
Qenclosed
−El A + Er A =
ε0
or
Qenclosed
Er − El =
Aε0
where Er is the field at the right end of the Gaussian surface and El is the field at the left end of the Gaussian
surface.
(d) Compute the Outer Fields: The electric field for infinite planes does not change with distance, therefore
the electric field at the very left of the system is equal and opposite to the electric field at the very right of the
system.
EI = −EIII
Apply our general form for Gauss’ Law of parallel planes to a Gaussian surface with one end in Region I and one
end in Region III as drawn above,
Qenclosed
Er − El = EIII − EI =
Aε0
The total charge inside our Gaussian surface is (−3σ + σ)A = −2σA = Qenclosed
−2σA −2σ
EIII − EI = =
ε0 A ε0
Using EI = −EIII gives,
−2σ
2EIII =
ε0
~ III = −σ x̂
E
ε0
~ I = −E
E ~ III = σ x̂
ε0
E~ II = −2σ x̂ _ _ +
ε0
I x II III
Check the fields have the same magnitude and direction as
the field map.
Solution
(a) Draw the System: Field lines begin at various places within
the volume charge and point radially outward.
I
II
r
(b) Select Appropriate form of Gauss’ Law and a Gaussian Surface: The appropriate Gaussian surface for
~ = E(r)r̂
a cylindrical system is a cylinder of length L and radius r. The electric field is radial and has the form E
where now ~r = (x, y, 0) if the axis of the system is the z axis. For this surface, Gauss’ law can be reduced to
2πrLE(r) = Qenc /ε0 or writing the field as a vector
~ = Qenc r̂
E
2πε0 rL
(c) Calculate the Field in Region I: A Gaussian surface in region I encloses part of the volume charge. The
charge enclosed by a Gaussian surface with radius r < a is the volume of the surface, the area of the end πr2
multiplied by L, time the charge density ρ,
Qenc = πr2 Lρ
and therefore the electric field in region I is
2
~ I = Qenc r̂ = πr Lρ r̂ = rρ r̂
E
2πε0 rL 2πε0 rL 2ε0
(d) Calculate the Field in Region II : A Gaussian surface in region II encloses all of the volume charge. The
charge enclosed by a Gaussian surface with radius r > a is
Qenc = πa2 Lρ
This chapter covers the final few topics for systems where the location of all charge is known. This is in
contrast to next chapter where we have to figure out the location of the charge.
Field Far from a System of Net Charge: The electric field at a point far from a
system with net non-zero charge (compared to the extent of the system), is the field of
a point charge at the center of the system with charge, Qtotal , the total charge of the
system.
Long Range Field of a Dipole System: If a system has zero net charge, but non-zero
dipole moment p~, the electric field far from the charge will be the electric field of a
simple dipole with dipole moment p~.
Disk
Solution
124
12.2. GAUSS’ LAW INSIDE AND OUTSIDE ALL CHARGE CHAPTER 12. FINAL TOPICS TEST 1
Since all charge densities are positive, the system has a net charge and therefore the field far from the system is
that of a point charge with the total charge of the system, Qtotal . The total charge is the sum of the total charge
of both shells and the volume charge
4 4
Qtotal = πρa3 + πb2 σ + πρa3
3 3
Therefore, the electric field far from the system is
4 3 2 4 3
~ = kQtotal r̂ = k 3 πρa + πb σ + 3 πρa r̂
E
r2 r2
~ outside = QT kQT
E 2
r̂ = 2 r̂
4πε0 r r
Electric Field inside all Charge - Spherical Symmetry: For a spherically symmetric
system of charge, if we are inside all charge then the electric field is zero.
Electric Field Inside All Charge - Cylindrical Geometry: Inside all charge for a
cylindrically symmetric charge distribution, the electric field is zero.
Electric Field Outside All Charge - Cylindrical Symmetry: The electric field outside
all charge for a cylindrically symmetric distribution of charge is
~ λ
E(r) = r̂
2πε0 r
Planar systems are somewhat different because the concept of inside and outside does not make much sense.
There is no center for a planar system. For planar systems, we can consider the electric field to the far right and
far left of the system.
Electric Field Outside All Charge - Planar Symmetry: For a system with planar
symmetry, far from all charge, the electric field is the same as a single plane with charge
density being the total charge density of the system
~ = σT
E outward.
2ε0
The total charge density can be computed by taking a cylindrical Gaussian surface with
end area A and finding the total charge in it, QT , σT = QT /A.
neutral
tube plastic
Ftube, + bottle Ftube, −
Fnet
Some of our observations about the qualitative behavior of conductors will allow us to deduce the key math-
ematical relations governing conductors in an electric field. Recall the following:
• Charge can move in a conductor.
• The net electric force inside a conductor is zero. We argued that charge separation would occur until
the total force on the electrons (and protons) inside a conductor is zero. The total force is the sum of the
external force and the electric force of the separated charge.
• Charge will move in response to an electric force. If there is a net electric force on charges in a
conductor the charges will accelerate and move.
From these observations, we can deduce these laws for the electric field inside a conductor in electrostatic
equilibrium:
126
13.1. RESPONSE OF MATERIALS CHAPTER 13. CONDUCTORS AND DIELECTRICS
Electric Field Inside the Conductor is Zero: The mobile charge in a conductor
~ = 0, in the interior of the conductor.
rearranges itself to produce zero electric field, E
If there was a net field, charge would move.
It is somewhat unbelievable that a conductor can accomplish this at all points no matter what. I have
always found the following the most convincing argument for the need for zero electric field in a conductor.
Suppose the field was not zero. Since there is mobile charge in a conductor, there would be a flow of charge, an
electric current. Electric currents in conductors cause heating and therefore a loss of energy. This cannot happen
indefinitely. Eventually, the conductor must reach an equilibrium where the current is zero and therefore the field
is zero.
The same reasoning can be used to show the electric field must
be perpendicular to the surface of the conductor. Suppose the
electric field, E ~ 0 , was not normal to the surface at some point
as shown to the right. The field could be decomposed into a Eperp
~
field perpendicular to the surface E⊥ and a field parallel to the
surface E ~ par . The component of the field parallel to the surface
E0
will cause currents to flow along the surface. The surface charge
will rearrange until these currents stop. Epar
conductor
Solution
The electric field is just the electric force that a positive charge would feel. In Course Guide 2 Electric Force I, we
asked the question, “Why doesn’t all the charge separate out of a conductor?” We reached the conclusion that
each charge that separates partially cancels the electric force that the next charge would feel. The same reasoning
applies to the electric field in a conductor. Initially, the electric field causes electrons to move to one surface,
leaving a positive region on the other surface. The electric field for these separated charges partially cancels the
external electric field. More charge separates until the electric field is zero inside the conductor. If the field were
not zero, more charge would be pushed to the surface.
Alternately, if the electric field is not zero, charge must flow until the electric field becomes zero.
Dielectric Breakdown: Dielectrics stay insulating only for electric fields below a certain
magnitude, otherwise you obtain a spark and a current flows. The magnitude of the
electric field where the dielectric begins to spark is the dielectric breakdown voltage.
~ = 3 × 106 N/C.
For dry air, dielectric breakdown happens at |E|
Fixing the Conductors: First erase any field lines crossing a conductor because the
field inside a conductor is zero, then bend the field lines that intersect the conductor so
they are normal to the conductor surface.
Fixing the Dielectrics: First thin field lines crossing the dielectric by a factor of the
dielectric constant κ, so if κ = 2 erase half the lines, and if κ = 3 erase 23 of the lines,
then bend the field lines which intersect the dielectric so they are closer to the normal
of the dielectric surface.
Draw Induced and Bound Charge: Draw + charge where field lines begin and −
charge where field lines end.
Example 13.2 Drawing the Electric Field Map with Dielectrics Present
Problem: Draw the electric field map of one point charge, +Q, and an uncharged dielectric sphere with dielectric
constant κ = 3.
Solution
Strategy: Draw the fixed charge map, then thin the lines in the dielectric by κ and bend them toward the
surface normal.
(a) Draw the field of the fixed charges: Draw the field of
the fixed charge. Since the dielectric is uncharged, just draw
the physical location of the dielectric.
Dielectric
+Q
(b) Thin Field Lines in the Dielectric: The dielectric will thin the field lines which cross it in proportion to
its dielectric constant. So if κ = 3, the field lines are a third as dense. Erase an appropriate number of field lines
within the dielectric. There were 3 field lines inside the dielectric, and 3/κ = 1, so erase 2 field lines.
Example 13.3 Drawing the Electric Field Map with Conductors Present
Problem: Draw the electric field map of one point charge, +Q, and an uncharged conducting sphere.
Solution
(a) Draw the field of the fixed charges: Draw the field
of the fixed charge. Sketch the location of the conductor and
~ = 0 in a conductor.
erase the lines inside, because E
+Q Conductor
(c) Conserve the Induced Charge: The net number of lines leaving the conductor must be correct for its
charge. So if the conductor is uncharged, the lines entering must equal the lines leaving. Counting field lines will
show that no net lines are leaving, so the conductor is uncharged.
~ 2 · n̂2 A = σA
~ 1 · n̂1 A + E
φ1 + φ2 = E
ε0
or cancelling A and multiplying by ε0
~ 1 · n̂1 + E
ε0 (E ~ 2 · n̂2 ) = σ
Surface Charge Density in Gaussian Pillbox: The surface charge density in Gaussian
pillbox with normals n̂1 and n̂2 is
~ 1 · n̂1 + E
ε0 (E ~ 2 · n̂2 ) = σ
~ 1 and E
where E ~ 2 is the electric field on either side of the surface.
n− n+
Solution
(a) Draw the Gaussian Surface: Use the Gaussian pillbox drawn above. Gauss’ law applied to the pillbox is
~ + · n̂+ A = σA
~ − · n̂− A + E
E
ε0
~ + · n̂+ = σ
~ − · n̂− + E
E
ε0
(b) Work out the Dot Products: By observation, n̂− = −x̂ and n̂+ = x̂,
~ − · (−x̂) = −100 N
~ − · n̂− = E
E
C
~ + · (x̂) = 180 N
~ + · n̂+ = E
E
C
Apply Gauss’ law,
~ − · n̂− + E
~ + · n̂+ ) = (8.85 × 10−12 C2 N N C
σ = ε0 (E )(−100 + 180 ) = 7.1 × 10−10 2
Nm2 C C m
The expression above can be used at any surface. It can be simplified quite a bit if the surface and the field
have planar symmetry.
ε0 (E2 − E1 ) = σ
_ +
Decrease in Electric Field : The electric field is reduced inside the dielectric by the
field of the bound charge. The dielectric constant, κ, tells how much the electric field
is reduced inside the dielectric. So in symmetric geometries if the electric field would
have been E ~ 0 with no dielectric, then the electric field becomes E
~κ = E
~ 0 /κ inside the
dielectric. The figure below is drawn assuming κ = 2 so there are half the field lines in
the dielectric.
Dielectric
- +
- +
- +
E0 Eκ E0
I II III
The magnitude of the bound surface charge density can be found by applying Gauss’ law to a Gaussian pillbox
shown surrounding the left surface of the dielectric. Gauss’ law yields
E0 1
−σb = ε0 (EII − EI ) = ε0 ( − E0 ) = ε0 E0 ( − 1)
κ κ
1
σb = ε0 E0 (1 − )
κ
Magnitude of Bound Charge at Dielectric Surface: The bound charge density at
the surface of a planar dielectric with dielectric constant κ is
1
σb = ±ε0 E0 (1 − )
κ
Polarization Electric Field : The electric field in the dielectric is decreased because
the bound charge creates an electric field E~ b . The electric field in the dielectric, E
~ κ , is
~ ~
the vector sum of the field in free space, E0 , and Eb ,
~κ = E
E ~0 + E
~ b.
~ 0 and E
Since these fields E ~ b are always in opposite directions, the magnitude of E
~ κ is
always the difference of their magnitudes, and in the direction of the free space field.
For any case we consider in this class E~ 0 is always larger.
- +
- +
- +
Eb
Solution to Part(a)
- +
- +
E0 Ekappa E0
Solution to Part(b)
The bound charge produces an electric field which partially cancels E ~ 0 to produce E
~ κ . This is why the electric
~ ~ ~
field is reduced in the dielectric. Therefore, |Eb | = |E0 | − |Eκ | = 10000 C − 1785.7 N
N N
C = 8214.3 C . The bound
~ 0.
field points in the opposite direction to E
Solution to Part(c)
Using a Gaussian cylinder with end area A at the right surface of the dielectric, we find
Qenclosed
φe = AEright − AElef t =
ε0
where Eright = E0 is the magnitude and sign of the field on the right surface and Elef t = Eκ is the magnitude
and sign of the electric field on the left surface. Let the dielectric slab lie in the y − z plane and let +x be to the
right.
Qenclosed σb A
Eright A − Elef t A = =
ε0 ε0
E0 1
σb = ε0 (Eright − Elef t ) = ε0 (E0 − ) = ε0 E0 (1 − )
κ κ
Cancelling A yields, Therefore, the bound charge density is
2
1 −12 C N 1 nC
σb = ǫ0 E0 1 − = (8.85 × 10 )(10000 ) 1 − = 72.7 2
κ Nm2 C 5.6 m
Solution to Part(d)
The individual atoms or molecules polarize throughout the material, mostly cancelling in the middle, but leaving
a little “extra” charge on each surface.
13.5 Superposition
The response of both a planar conductor and a planar dielectric to an external electric field was to generate
equal and opposite planes of surface charge. The field is zero in a planar conductor because the field of the
induced surface charge exactly cancels. To use this observation, we will need to understand the field of equal and
opposite planar charges.
Consider the two fixed planes of charge shown below. The left plane has surface charge density +σ and the
right plane surface change density −σ. If a cylindrical Gaussian surface with end area A encloses all the charge
of the system, the charge enclosed is Qenc = (σ + (−σ))A = 0. Therefore the field in region I and III is zero.
A Gaussian surface that encloses only the left plane encloses a charge of Qenc = σA. Applying Gauss’ law to this
surface yields,
Qenc σA
EII A − EI A = =
ε0 ε0
Since EI = 0, this can be solved to yield EII = σ/ε0 .
The electric fields of the infinite parallel planes can be added to an external field. Since the electric field
outside of the planes is zero, the charges on the planes do not change the external field except in the region
between the planes. Therefore, to calculate the electric field of a planar conductor we add the applied field to the
field of the equal and opposite surface charges.
_ _
+ +
added to _ + equals _ +
_ + _ +
_ + _
+
Example 13.6 Charge Density Required For Zero Field Between Planes
Problem: An external charge density (not drawn in the problem) produces a uniform electric field E ~ 0 = E0 x̂
where E0 > 0. Two infinite parallel planes with equal and opposite charge densities |σ| are placed in the field as
shown to the right. Calculate σlef t and σright such that the electric field between the planes is zero.
Solution
(a) Draw the Field Map: We are given that the applied field is
uniform and that the equal and opposite parallel planes of charge
create a field that cancels the external field producing zero field E0 σleft σright E0
between the planes. Therefore, to draw the field map we draw a _ +
uniform field outside the planes and zero field between the planes,
as shown above. Draw + charges where lines begin and − charges _
+
where lines end.
Air _ Air Air
+
_ +
_ +
I II III
(b) Use the Field Map to Deduce the Charge Densities: Observing the field map, we see that the left
plane must have a negative surface charge density σlef t < 0 and the right plane a positive surface charge density
σright > 0. The planes are equal and opposite so σlef t = −σright .
(c) Reason About Field Produced by the Planes: The applied field is E ~ 0 = E0 x̂ and it points in the positive
x̂ direction. The total electric field between the planes is zero, therefore the electric field of the planes in the
region between the planes, E~ planes,II , must satisfy
~0 + E
E ~ planes,II = 0,
σlef t = ε0 (EII − EI ) = ε0 (0 − E0 ) = ε0 E0
so viewing the problems as a superposition of plane charges is consistent with using a Gaussian pillbox.
We can
do the External Field Field of Bound Charge Field of Dielectric
same a b a b
analysis _ _
+ +
for a
dielectric
slab.
added to _ + equals _ +
_ + _
+
Example 13.7 Charge Density Required for Reduced Field between Parallel Planes
Problem: An external charge density (not drawn in the problem) produces a uniform electric field E ~ 0 = E0 x̂
where E0 > 0. Two infinite parallel planes with equal and opposite charge densities |σ| are placed in the field.
Calculate σlef t and σright such that the electric field between the planes is reduced by a factor of κ = 2.
Solution
(a) Draw the Field Map: We are given that the applied field is
uniform and that the equal and opposite parallel planes of charge
create a field that reduces the external field by a factor of 2 be- σleft σright
E0 E0
tween the planes. Therefore, to draw the field map we draw a _ E +
uniform field outside the planes and a field with half the number E= 0
2
of lines between the planes, as shown above. Draw + charges
where lines begin and − charges where lines end.
Air _ Air Air
+
_ +
I II III
(b) Deduce the Sign of the Charge on Each Plane: Observing the field map, we see that the left plane must
have a negative surface charge density σlef t < 0 and the right plane a positive surface charge density σright > 0.
The planes are equal and opposite so σlef t = −σright .
(c) Reason About Field Produced by the Planes: The applied field is E ~ 0 = E0 x̂ and it points in the positive
~ ~
x̂ direction. The total electric field between the planes is E0 /κ = E0 /2, therefore the electric field between the
planes, E~ planes,II , must satisfy
~
E ~ planes,II = E0 .
~0 + E
κ
The electric field of the planes, between the planes, must point in the negative x̂ direction. Let E ~ II,planes =
EII,planes x̂.
E0
E0 + Eplanes,II =
κ
From this we conclude
E0 1
EII,planes = −E0 + = −E0 1 −
κ κ
.
(d) Use the Field of Parallel Planes: The magnitude of the electric field between TWO infinite parallel planes
of charge is
σright σlef t
|Eplanes,II | =
=
ε 0 ε0
1 σright σlef t
|Eplanes,II | = E0 1 − = =−
κ ε0 ε0
Therefore,
1
σright = −σlef t = ε0 E0 1 −
κ
or substituting κ = 2,
ε 0 E0
σright = −σlef t =
2
(e) Check Against Our Formula for Surface Charge Density: In the previous section, we derived a formula
for surface charge density at a plane surface
E0 1
σlef t = ε0 (EII − EI ) = ε0 ( − E0 ) = −ε0 E0 1 −
κ κ
so viewing the problems as a superposition of plane charges is consistent with using a Gaussian pillbox.
The electric field in a conductor is zero. To add an uncharged conductor to a field map of a system with
spherical, planar, or cylindrical symmetry, erase the lines in the conductor and draw + charges where lines end
and − charges where lines begin. To add a charged conductor to a field map, redraw from scratch using Gauss’
Law (Version 0). I am sorry that the stupid charge can move in a conductor, that just ruins everything.
Solution
Draw the field map ignoring the conductor based on a calculation of the field strength. Sketch in the location of
the conductor. Erase all field lines in the conductor. Draw charge where field lines begin and end.
141
14.1. FIELD MAPS CHAPTER 14. HIGH SYMMETRY SYSTEMS II
Solution
(a) Draw the Field Map Ignoring the Dielectric: This is the same fixed charge system as the previous
example. Draw the field map ignoring the dielectric.
σ1
σ1 σ2 = −
2
+ - + -
+
+ - +
+ -
+ - + -
+
+
+ -+ -
+ - + -
+
+
+ - + -
I II III IV V
(b) Fix the Field Map Interior to the Dielectric: Now add the dielectric and thin the lines by a factor of
κ = 3, so only three lines cross the dielectric. Draw charges where field lines begin and end.
Solution
The dielectric thins the field lines by a factor of κ = 2, so eliminate half the field lines in the dielectric. Since the
dielectric is uncharged, the same number of field lines must leave the dielectric as leave the central charge. Draw
+ charge where lines begin and − charge where lines end.
The System
Field of Fixed Charge
+ +
_ _
Q Q Q
Air
_ _
+ +
Dielectric Dielectric
Air
highly symmetric system where Gauss’ Law is useful for computing the field proceeds in the same way for both
objects. We work the problem as we did in the static system and then correct the electric fields for the presence
of the object, setting the field to zero inside a conductor and dividing the field by κ inside an insulator. As the
Gaussian surface is moved from region to region, Gauss’ Law allows the calculation of the induced charges on the
conductors and the bound charges on the dielectrics.
Adding an Uncharged Conductor: To work a problem with an uncharged conductor,
work the problem without the conductor, then draw the conductor in, erasing the field
lines inside. The electric field inside the conductor is zero, the electric field outside the
conductor is unchanged. The induced surface charge can be computed at each surface
by applying Gauss’ Law and requiring the field to be zero in the conductor.
Adding an Uncharged Dielectric: To work a problem with an uncharged dielectric,
we work the problem without the dielectric, then draw the dielectric in. The electric
~
field inside the dielectric is E/κ, ~ is the field we calculated with no dielectric
where E
and κ is the dielectric constant. The electric field outside the dielectric is unchanged.
The bound surface charge can be computed at each surface by using Gauss’ Law, with
the requirement that it produces the correct dielectric field.
Example 14.4 Gauss’ Law with Conductors and Dielectric but no Surface Charge
Problem: A point charge with charge Q is surrounded by an uncharged dielectric shell with dielectric constant,
κ, and an uncharged conductor as shown below.
(a)Compute the electric field in region I.
(b)Compute the electric field in region II.
(c)Compute the electric field in region III.
(d)Compute the electric field in region IV .
(e)Compute the electric field in region V .
Dielectric
Air
Conductor
Air
Q
I
II
III
IV
V
For a system of charge with spherical symmetry, Gauss’ Law can be simplified to
~ = Qenclosed r̂.
E
4πε0 r2
The only net charge in the system is the point charge, since neither the conductor nor the dielectric have net
charge. The electric field of the point charge, either using Gauss’ Law or Coulomb’s Law, is
~ = Q
E r̂.
4πε0 r2
The conductor and dielectric only modify this field in their interior, so in regions I, III, and V, the total charge
enclosed is Q and the electric field is
~ I = Q r̂.
E
4πε0 r2
The electric field in region III is outside of the conductor and dielectric and is unchanged from the point charge
field.
~ III = Q r̂
E
4πε0 r2
The electric field in region IV is the electric field in the region if no dielectric were present, divided by κ, the
dielectric constant. Therefore,
~ IV = Q
E r̂.
4πε0 κr2
A Gaussian surface in region V is outside of all charge and encloses a total charge of Q, therefore the electric
field in region V is
~ V = Q r̂.
E
4πε0 r2
The next step is to learn how to add a charged conductor to a highly symmetric system of charge. This
presents a big problem. In every other case, we have known where the net charge is. In this case, the net charge
can move from side to side in the conductor and we have to use Gauss’ law to figure out where it is. The first
example uses only conductors and the second example throws a dielectric into the mix.
Solution
(a) Draw a Good Diagram: Let the inner conductor be conductor 1 and the outer shell be conductor 2. Select
6 lines per Q. The total charge in a spherical Gaussian surface in region II as drawn is Qenc = Q, so by Gauss’
Law version 0, 6 lines cross region II. Four lines cross region IV because the total charge enclosed by a Gaussian
surface in region IV is Qenc = Q − Q/3 = 2Q/3. Draw − charge where lines end and + charge where lines
begin.
_
Conductor 2
_ _
+
+
+ Conductor 1 c
b
+ I a +
+ _
_
Air II
III
_ IV
Air
(b) Select Gauss’ Law Appropriate for Symmetry: For spherical symmetry, the electric flux out of a Gaussian
surface of radius r is 4πr2 E(r) and by Gauss’ law
Qenc
φe = 4πr2 E(r) =
ε0
where Qenc is the charge enclosed in a Gaussian surface of radius r.
(c) Compute the Electric Field in all Regions: The electric field in region I and region III is zero because
of the conductors.
E~I = E~ III = 0
The total charge enclosed by a Gaussian surface in region II is Q, therefore the electric field in region II is
~ II = Q
E r̂
4πε0 r2
The total charge enclosed by a Gaussian surface in region IV is Qenc = Q − Q/3 = 2Q/3. Apply Gauss’ Law
~ IV = 2Q/3 r̂ = Q r̂
E
4πε0 r2 6πε0 r2
(d) Compute the Surface Charge Density on Conductor 1: The net charge Q must be on the outer surface
of conductor 1. The surface charge density on the outer surface of the inner conductor is
Q
σ1 =
4πa2
(f) Compute the Surface Charge Density on the Outer Surface of Conductor 2: The total charge on
conductor 2 must be −Q/3 which must be the sum of the charge on the inner surface and the charge on the
outer surface,
−Q
= Q2,in + Q2,out = −Q + Q2,out
3
therefore
2Q
Q2,out =
3
The charge density on the outer surface of the conducting shell is
Q2,out Q
σ2,out = =
4πc2 6πc2
Definitions
Gaussian Surface
Strategy: Use result of Gauss’ Law for spherical symmetry in each region. Figure out bound charges from the
change in field across material boundaries.
(a) Draw a Good Diagram: All field lines are radial. I chose 8 lines going out for the central Q. All these lines
end on the conductor. If the dielectric was not present a Gaussian surface outside the conductor would contain a
total charge of −Q, so in region IV there are 8 lines going in. The dielectric thins the lines by a factor of κ, so
in the dielectric there are 8/3 ≈ 2 lines going in. Draw charge where lines begin and end.
(b) Compute E ~ I : By Gauss’ Law, the electric field of a spherically symmetric charge distribution is
~ = Qenclosed r̂
E
4πε0 r2
In region I, the total charge enclosed is Q, so
~I = Q
E r̂
4πε0 r2
Compute E ~ III : In region III, the electric field would be the same as region IV if no dielectric were present. The
dielectric thins the fixed charge field by a factor of κ, so
~
~ III = EIV = −Q r̂
E
κ 4πε0 κr2
~ IV = −Q
E r̂
4πε0 r2
Compute Induced Charge on Conductor: Since the electric field is zero in the conductor a Gaussian surface
in the conductor encloses zero charge. The charge enclosed by a Gaussian surface in the conductor is the central
charge, Q, plus the induced charge on the inner surface of the conductor Qc,inner , so Q + Qc,inner = 0, or
Qc,inner = −Q
The total charge of the conductor is −2Q, which must be the sum of the inner and outer charge densities on the
conductor, so −2Q = Qc,inner + Qc,outer and solving
Qc,outer = −Q
Compute Bound Charge on the Dielectric: The field in the dielectric is E ~ III = E~ IV = −Q 2 r̂, but in the
κ 4πε0 κr
dielectric this must be the field from the total charge enclosed which is Qenclosed = Q − 2Q + Qd,inner , which
by Gauss’ Law give a field of E~ = Qenclosed
4πε0 r 2 r̂. Comparing these two expressions we get,
Q Q
Qenclosed = − = − = Q − 2Q + Qd,inner .
κ 3
Solving gives
2
Qd,inner = Q
3
2
Qd,outer = − Q
3
Solution to Part(a)
a +
b
+ -
+
+ - + - +
+
Conductor Dielectric
x
(b) Compute the Field Outside the Conductor and Dielectric: The electric field of an isolated plane of
−σ
charge is 2ε0
x̂ if x < 0 and 2εσ0 x̂ if x > 0. Therefore, we can immediately write the field in regions I, III, IV ,
V I since they are outside of the conductor and dielectric.
~ III = − σ x̂
~I = E
E
2ε0
~ V I = σ x̂.
~ IV = E
E
2ε0
(c) Correct the Fields in the Conductor and the Dielectric: The electric field in the conductor is zero,
~ II = 0. The electric field in the dielectric is reduced by a factor of κ from the field that would exist if no
E
dielectric were present.
E~ V = σ x̂
2ε0 κ
Solution to Part(b)
(a) Select an Appropriate Gaussian Surface: Use the cylindrical Gaussian surfaces drawn which enclose the
left surface of the conductor, a, and the left surface of the dielectric, b.
(b) Apply Gauss’ Law to Surface a: If surface a has end area A, the charge enclosed is σc,ℓ A. By Gauss’ Law,
Qenc σc,ℓ A
EII A − EI A = =
ε0 ε0
Since EII = 0,
σc,ℓ
−EI =
ε0
σ
.
σc,ℓ = −ε0 EI =
2
(c) Conserve Charge on Conductor: Since the conductor is uncharged, σc,ℓ + σc,r = 0, therefore
−σ
σc,r = ε0 EI =
2
Notice the signs of the charge density match the signs of the charges you drew in the figure.
(d) Apply Gauss’ Law to Surface b: The total charge enclosed in surface b, if it has end area A, is Qenc = σd,ℓ A.
Apply Gauss’ Law,
Qenc σd,ℓ A
EV A − EIV A = =
ε0 ε0
σ σ σd,ℓ
− =
2ε0 κ 2ε0 ε0
σ 1
σd,ℓ = −1 .
2 κ
(e) Conserve Charge on the Dielectric: Since the dielectric is uncharged, σd,ℓ + σd,r = 0.
−σ 1 σ 1
σd,r = −1 = 1−
2 κ 2 κ
Definitions
End View of Cylinder
_
Conductor Qc ≡ Charge of Conductor
_ _ Qv (r) ≡ Charge Enclosed in Volume at radius r
Air L ≡ Arbitrary Length of Objects
+ ~
_ _ + Volume _ _ Ei ≡ Electric Field in Region i
Charge + ~r ≡ Radius Vector
+ I c ≡ Radius of charged cylinder
−λ ≡ Charge per unit length on conductor
_
II _ a ≡ Inner radius of conductor
_
_ III b ≡ Outer radius of conductor
Gaussian IV
Surface
(a) Derive a General Expression for Gauss’ Law in Cylindrical Symmetry: Select a Gaussian surface that
is a cylinder co-axial with the ẑ axis or radius r and length L. The electric flux out of the Gaussian surface is
φe = 2πrLE(r) and applying Gauss’ law
Qenc
φe = 2πrLE(r) =
ε0
(b) Compute Electric Field in Region I: Now we apply our general result to each region. For any radius
in region I, the Gaussian surface encloses charge Qv (r) = πr2 ρL. Notice the charge enclosed changes with the
radius. Substitute into the general expression for the field computed above
πr2 ρL
EI (r) =
2πǫ0 rL
~ I (r) = ρr r̂
E
2ǫ0
For any system where an arbitrary length L must be introduced, that L must cancel from the final answer, or else
you have done something very wrong.
(c) Compute Electric Field in Region II: In region II, the Gaussian surface always encloses all the volume
charge Qv (c) = πc2 ρL. Substitute:
πc2 ρL
EII (r) =
2πǫ0 rL
2
~ II (r) = c ρ r̂
E
2ǫ0 r
When the conductor is not present this is the field at all points outside of the volume charge.
(d) Compute Electric Field in Region III: The electric field inside the conductor is zero for static systems.
~ III (r) = 0.
E
(e) Compute Electric Field in Region IV : In region IV the total charge enclosed by the Gaussian surface is
Qenclosed = Qv (c) + Qc = πc2 ρL − λL. Substitute:
πc2 ρL − λL
EIV (r) =
2πǫ0 rL
2
~ IV (r) = πc ρ − λ r̂
E
2πǫ0 r
The electric field for the full system is computed above. In the conductor the field is zero, so by Gauss’ Law
the charge enclosed in a Gaussian surface in the conductor must be zero. The total charge enclosed is Qv (c) +
λc,inner L = 0 where λc,inner is the charge per unit length on the inner surface of the conductor. Therefore,
λc,inner = −πc2 ρ
To convert this to the surface charge density we divide the total charge on a length L of the inner surface,
λc,inner L by the surface area 2πaL, giving
λc,inner c2 ρ
σc,inner = =
2πaL 2a
The total charge per unit length on the conductor is −λ which is the sum of the charge on the inner and outer
surfaces, the only place charge can be: λc,outer + λc,inner = −λ , so
λc,outer = πc2 ρ − λ
λc,outer πc2 ρ − λ
σc,outer = =
2πbL 2πb
We are given πc2 ρ < |λ|, so |Qv | < |Qc |. This means that outside the conductor the net charge enclosed is
negative so the field lines go in. Inside the conductor only positive charge is enclosed so field lines point outward.
Since lines end on both the inside and outside of the conductor, there is negative charge on both surfaces. Using
a dashed line, draw the general Gaussian surface on the diagram. The surface is a cylinder with length L. Its
radius is r and its outward normal is r̂.
The formula above allows us to calculate the charge enclosed in a Gaussian surface for a non-uniform volume
charge.
Example 15.1 Gauss’ Law with a Non-Uniform Charge Density r
Problem: A system of charge has spherical symmetry. For points r < R, the charge density is ρ(r) = ρ0 e− a
and zero for points r > R. Calculate the electric field everywhere.
Solution
(a) Draw the Field Map: Separate the system into two regions,
one inside and one outside the charge. The field lines begin at
various points throughout the volume charge. A sample Gaussian
surface of radius r is drawn.
I
II
153
15.1. NON-UNIFORM CHARGE DENSITIES CHAPTER 15. NON-UNIFORM CHARGE DENSITIES
(b) Calculate the Field in Region I: A spherical surface of radius r encloses a charge of
Z r Z r Z r
r r
2 2 −a
Qenc (r) = 4πr ρ(r)dr = 4πr ρ0 e dr = 4πρ0 r2 e− a dr
0 0 0
(c) Calculate the Electric Field in Region II: The total charge enclosed by any Gaussian surface in region II is
2 2 2
~ II (r) = Qenc (r) r̂ = 4πρ0 a(a − exp(−R/a)(a + 2aR + R )) r̂
E
4πε0 r 2 4πε0 r 2
Note the expression we got was ugly, but the physics we did was easy and the math we just looked up.
The course up to this point has presented a “force” picture of electrostatics. Our primary descriptive tool
has been the electric field which is the force per unit charge. For many situations it is more powerful to analyze
systems in terms of energy. This chapter reviews the energy concepts from UPI that we will need to understand
electrostatic energy.
Definition of Work: The work done by A to move object B along the curve C is
Z
WAB = F~AB · d~r
C
where FAB is the force A exerts on B and d~r points in the direction of the curve.
The integral in the definition of work is taken along the curve C. As usual to see where it comes from imagine
cutting the curve below up into little pieces of length ∆s. The vector d~r points from one end to the other end of
the small segment, as drawn below. The vector d~r has length ∆s and points in a direction tangent to the curve.
155
16.2. TOTAL ENERGY CHAPTER 16. WORK AND ENERGY
The Curve C
FAB
2
∆s
θ
dr
The component of force in the direction of motion is FAB cos θ. The work done by A on B as it moves from
point 1 to point 2 along the curve is
F~AB · ∆~r
X X
WAB = FAB ∆s cos θi =
i i
This work is just the component of the force along the curve at each point multiplied the length of the curve.
System
If two particles i and j are both in the system, then the forces between the particles,F~ij and F~ji , are called internal
forces. If the particle i is in the environment and particle j is in the system, then we will call i an external agent
and the force F~ij , the force the external agent exerts on the system at point j or more briefly F~ij is an external
force.
Imagine building a system from scratch, starting from an empty box with all the particles that make up the
system infinitely far apart. (With the particles scattered to the ends of the universe?) The person or thing, the
external agent, who builds the system does so by exerting the force F~i on each particle in turn needed to bring
the particle in from infinity and place it where it goes in the system. If the particle is to have kinetic energy,
the external agent then exerts some more force to give the particle the velocity it needs. It requires no work to
move the first particle into the empty box, since there are no particles in the system to exert a force to resist the
motion. The external agent must do work against the force of the first particle to place the second particle. Any
system can be built up one particle at a time. Define the total energy, ET OT , of the system as the total work to
build it. Note I am using the symbol ET OT for total energy not electric field.
Definition of Total Energy: The total energy of a system, ET OT , is the total work
required to build the system, X
ET OT = Wi
i
where Wi is the work to place the ith particle in the presence of the particles j < i.
If the system is isolated, that is if there are no external forces acting on the system, then the total energy is
constant.
W = ∆ET OT
where U is the total potential energy of the system defined as the work an external
agent would do to build the system piece by piece, placing each particle at rest. The
total energy of an isolated system is constant.
This is more general than we need for this class. In most cases we will only allow one of the particles in
the system to move. Only that particle has kinetic energy and the potential energy of the system becomes the
potential energy of the particle. In this case, we choose the zero of potential to absorb the irrelevant potential
energy of the other particles.
Conservation of Energy for One Particle: If only the ith particle of the system can
move, then the total energy of the system can be written
ET OT = Ki + Ui = constant
where Ui is the potential energy of the ith particle and the energy is constant only if
the system is isolated.
This can be re-written in terms of the change in energy. For an isolated system the sum of the change in
kinetic energy, Ki , and the change in potential energy, Ui , is zero
∆Ki + ∆Ui = 0
If the particle only moves in the x-direction, then the potential energy has the form Ui (x). We can predict
many of the qualitative features of particle’s motion just from the shape of potential energy function. Consider
a particle released in the potential energy function below, with total potential energy ET OT .
U(x)
ETOT
B A
turns fast slow released
x
If a particle is released at point A with zero kinetic energy, the particle will travel to the left converting
potential energy into kinetic energy. Eventually, it reaches point B where the total energy of the particle is all
potential and the particle turns. Since there are no losses, the particle oscillates between A and B forever. If
we are given or can find the potential energy Ui (x) and the point where the particle is released, x0 , then the
kinetic energy at any point can be predicted. Since the particle is released at rest, the kinetic energy at x0 is
zero, Ki (x0 ) = 0. The change in kinetic energy plus the change in potential energy is zero
Re r
Solution
Since the hole opens up under my feet, my initial kinetic energy is zero, K(Re ) = 0, when I an a distance Re
from the center of the earth. Conservation of energy tells us
mgRe2
1
∆K + ∆U = 0 = (K(0) − K(Re )) + (U (0) − U (Re )) = (0 − mvc2 ) + (0 − )
2 2Re
Naturally, I fall completely through the Earth and only turn around as I reach the surface of the Earth on the
other side of the planet.
My wife felt you would be alarmed by the above expression after your experience with gravity in UPI. It is true
that g is only valid at the surface of the earth, but I can remember g. So to save looking up the gravitational
constant G and the mass of the earth M , I used GM = rRe2 . I also chose the zero of potential energy at the
center of the Earth rather than at the surface of the Earth.
where x0 is the point where the potential energy is zero and Fx is the force an external agent must exert to move
the particle with zero kinetic energy from x0 to x. By the fundamental theorem of calculus, this implies
dU
Fx =
dx
is the force the external agent must apply. The force F that the other particles exert on the particle must be
equal and opposite the force the external agent must exert.
Force is the Derivative of Potential Energy: The force, F , on the particle x from
the other particles of a system is
dU
F =−
dx
where U (x) is the potential energy of the particle.
mgr2
U=
2Re
where m is his mass, g is the acceleration of gravity, r is the distance from the center of the Earth, and
Re = 6.4 × 106 m is the radius of the Earth.
Calculate the force on Dr. Stewart at the Earth’s surface and at the center of the Earth.
Solution
dU d mgr2 mgr
F (r) = − =− =−
dr dr 2Re Re
where positive forces are directed upward. At the surface of the earth, F (Re ) = −mg, which is what we expected.
At the center of the earth, F (0) = 0. So there is no force on me as I pass through the center of the earth at
7800 ms .
Point A is a local maximum in the potential energy, the top of a mountain. If a mass, like a skier, was placed
at point A and released, they would feel a force toward lower potential energy and slide downhill. When the lines
of equal potential energy are close together, the vertical elevation is changing in a short distance and the terrain
is steeper. A skier placed at B would feel a large force down the mountain. Where the lines are farther apart,
like at C, the country is flatter and the force down the mountain less. Remember, a skier is accelerated by the
component of gravity along the ground.
Electric Potential
where F~ext
i
is the force the external agent exerts and d~ℓ points along the path. The integral is taken along the
path the particle moves.
If the external agent does not change the kinetic energy of the ith particle as it is moved along the path, then
i
the work done is the change in potential energy ∆UAB between the points ~rA and ~rB ,
i i
WAB = ∆UAB if ∆K i = 0.
Only changes in energy actually matter in physics, so we get to pick the point where the energy is zero. Let the
potential energy be zero at the point ~r0 which we will call the reference point. For systems of charged particles,
the reference point will usually be a point far from the system of charge, “a point at infinity”. With a choice of
reference point, we can define the potential energy, U i (~rA ) of the ith particle at the point ~rA as the potential
difference between the point ~r0 and the point ~rA ,
U i (~rA ) = ∆U0A
i i
= W0A
where the work must be done without changing the kinetic energy. So if we could get the work somehow, we
could get the electric potential energy, the part of the potential energy due to the internal electric force.
Imagine the external agent moves the particle i at constant velocity so that the change in kinetic energy is
zero. To do this, the external agent must exert a force F~ext
i
which exactly balances the total force exerted by the
other charges of the system on charge i,
162
17.1. INTRODUCTION TO ELECTRIC POTENTIAL CHAPTER 17. ELECTRIC POTENTIAL
where E ~ i is the total electric field from the other charges in the system at the location of charge i. Substituting
this into our definition of work gives
Z
i
WAB = ∆UAB i
= −qi ~ i · d~ℓ
E
A→B
Rather than working with electric force, we found working with electric force per unit charge, electric field, more
powerful. While working with energy, we will find working with the difference in electric potential energy per unit
charge more useful.
Definition Electric Potential Difference: The electric potential difference ∆VAB
between the point A and the point B along a path ~ℓ in an electric field E
~ is defined to
be Z
∆VAB = − ~ i · d ~ℓ
E
A→B
where E~ i is the total electric field from the particles j 6= i, that is the other particles
in the system. We will shorten the name of the electric potential difference to simply
potential difference for this class.
Definition of Electric Potential: The electric potential V (~rA ) at the point ~rA is the
potential difference between a reference point, ~r0 , defined to have zero potential energy
and the point ~rA ,
V (~rA ) = ∆V0A
We will shorten electric potential to potential for this class.
Units of Electric Potential and Electric Potential Difference: The SI unit for
electric potential V and electric potential difference ∆V is the volt V. The symbol
representing potential and the symbol representing its units are very similar. Be
sure to look at the context of the symbols to determine which meaning should be
given. Equivalent units for electric potential are:
J N·m
1V = 1 =1
C C
New Units for Electric Field: When working with electric field it is often more con-
venient to use the units V/m instead of N/C for the electric field. They are equivalent.
V N
1 =1
m C
Electric Potential is Work per Unit Charge: This electric potential difference be-
tween a point A and a point B, ∆VAB , is the work, WAB , per unit charge an external
agent must do to move a charge q from A to B along the path
WAB
∆VAB =
q
Difference in Electric Potential Energy: The difference in electric potential energy
of the charge q between the point ~rA and the point ~rB is the charge multiplied by the
potential difference
∆UAB = q∆VAB
Computing Potential Difference from Potential: The potential difference between
two points ~rA and ~rB can be computed from the electric potential
and from there conserve energy and get the change in kinetic energy and then the change in velocity. The electric
potential is the integral of the electric field. The electric field can change discontinuously, for example inside and
outside of a charged shell. An integral can not change discontinuously, therefore:
Additivity of Electric Potential: The total electric potential may be found by adding
the potential produced by individual charges as long as a consistent reference point is
used.
Solution
The work done by the battery W = qe ∆V = (−1.602 × 10−19 C)(−12V) = 1.9 × 10−18 J is equal to the energy
added to the electron. From this number you may deduce that many electrons participate in electric current.
(a) 3 × 108 ms (b) 1.05 × 1016 ms (c-Answer) 1 × 108 ms (d) 9.8 × 10−8 ms
Solution
(c) 1 × 108 ms : As the electron accelerates, its potential energy due to the electric field will be converted to
kinetic energy, and so by Conservation of Energy
|∆U | = |∆K| = Kf
The difference in potential energy is related to the difference in potential through ∆U = q∆V so
1
|q∆V | = Kf = mv 2
2 f
r
2q∆V m
vf = = 1.0 × 108
m s
if we choose to integrate outward from A to B along a radial path such that d~ℓ = r̂dr. If we choose A as the
reference point and let ~rA become ~r0 and ~rB become ~r the potential becomes
Z r
V (r) = − E(r)dr.
r0
Electric Potential for a Field that Points Along a Coordinate Direction: If the
electric field has the form E(x)x̂, E(y)ŷ, E(z)ẑ, or E(r)r̂ then the electric potential
~ = E(x)x̂,
for this field is given by the negative derivative of the field. If, for example, E
then Z
V (x) = − E(x)dx + C
Calculation of Electric Field from Potential for Potentials with One Variable: If
the electric potential depends on only one variable, then the electric field is the negative
integral of the potential with respect to that variable. For example, if the potential is
V (x) then the electric field is
~ = − dV (x) x̂
E
dx
where similar relations apply for V (y), V (z), and V (r).
We will explore the properties of electric potential qualitatively in the next section, then work a bunch of
examples for the rest of the chapter.
Independence of Path: Neither the work to move an electric charge nor the potential
difference between any two points in an electric field depends on the path taken through
the field. This means we can use the most convenient path to compute work and
potential difference.
An alternate and equivalent statement of independence of path is that the total work to move a charge around
a closed path in a static electric field is zero.
Work Around any Closed Path is Zero: If a charge is moved around a closed path
in an electric field so that it returns to its starting point, the total work done on the
charge is zero.
Independence of path sort of looks like an insignificant side feature, but it is actually a new fundamental law
with the same importance as Coulomb’s law. We will study this law when we get to magnetism, where is will
be called Faraday’s law. So independence of path is really an application of Faraday’s law, where there are no
changing magnetic fields. With that, I’ll leave Faraday’s law for magnetism. Independence of path places some
strong constraints on what kind of static electric field maps we can draw.
17.1.4 Batteries
A battery is a device which maintains a fixed potential difference between its two terminals. The only way to
maintain or create a potential difference is to move charge. A battery is actually a chemical engine for pumping
charge. The engine runs by using a chemical reaction to supply the energy to move the charge. Since the internal
chemical reaction of the battery supplies energy to the system, the battery’s internal energy must be included in
any energy conservation equation.
Since we have finally brought up the idea of an electric circuit, it is a good time for introducing our basic
instrument for measuring potential difference, the voltmeter (the lab multimeter set in the DCV region).
An electric field is drawn to the right along with a path along which an external
agent moves a positive charge. If a positive charge, q, is moved from A to B
the electric force, F~ , points opposite the direction of motion. An external agent A
would have to exert a force opposite to this force to get the charge to move
from A to B. Since the external agent exerts a force in the direction the particle F
moves the work done by the external agent is positive. Therefore the potential
energy increases as a positive charge moves from A to B and VB > VA . q
Potential Increases Opposite to the Direction of the Electric Field: The electric
potential increases along a path that moves opposite to the direction of the electric
field.
We can turn this around and use it figure out the direction of the field from the potential. If ∆VAB > 0 then
you have to do work to move a positive charge from point A to B, therefore the electric field must on average
point from B to A so that the electric force on the charge resists the motion.
Reasoning from Potential Difference to Electric Field: The electric field points to
lower potential.
Since the electric field for a simple system is given by E = − dV
dx , the faster the potential changes with position,
the stronger the field.
Reasoning About the Size of the Field Based on the Potential: The stronger the
electric field, the greater the rate the electric potential changes with distance. Therefore,
places where the electric potential changes quickly are places where the electric field is
stronger.
There is no Potential Difference Across a Conductor: Since the electric field in a
conductor is zero, there is no potential difference between different points in a conductor
(if all charge is stationary).
Solution to Part(a)
The electric field points from high potential to low potential, since you have to do work on a positive charge to
move it to higher potential. Therefore, the electric field points in the negative x direction.
Solution to Part(b)
The electric field is strongest where the potential changes the fastest; therefore, the electric field is strongest at
(1cm, 2cm).
Equipotentials Do Not Cross: If two equipotentials cross, then the must have the
same potential. If they have different electric potentials then the electric potential has
two values at the same point. Not allowed.
Conductors are equipotential surfaces: Since E ~ is perpendicular to the surface of a
~ = 0 inside a conductor, all
conductor, the surface is an equipotential surface. Since E
points in that conductor are at the same potential, because it takes no work to move
a charge in zero electric field. Make sure none of the equipotentials you add to a field
map cross a conductor.
From an electric field map with equipotentials, we can reason about the relative magnitude of the potential
along different equipotentials and the relative work to move from point to point on a field map.
No Work to Move along Equipotential: Since an equipotential is, by definition, all
at the same potential, it takes no work to move along an equipotential.
Electric Potential Increases Opposite to
Direction of Field Lines: If two equipoten-
tials are connected by an electric field line,
the equipotential the field lines point to will
have lower potential. Potential difference is V1 V2
work per unit charge, so if we would have to
work to move a positive charge from point
2 to point 1, as above, then the potential is
higher at point 1.
Solution
(b) Draw equipotentials: Starting and ending at each point you selected, draw a closed curve that intersects
each field line at a right angle. Some of the curves may close outside the area of your figure.
Equipotentials Equipotentials
far apart close together
where field where field is
is weak strong
D C B A +Q +Q
Solution
Solution
The field map is just that of a dipole. To draw the equipotentials, draw dashed lines that are perpendicular
everywhere to the field lines.
(a)Draw the field map in the region between the conductors using 8 lines per +Q. You may draw the
diagram on the figure below.
(b)Draw 8 equipotential surfaces.
+Q
Conductor
-Q
+Q
_ _ _ _ _ _ _ _
Conductor
+ + + + + + + +
-Q
A
C
Solution
Compute the Magnitude: Use the definition of potential to compute the magnitude of
the electric potential, and then reason about the sign. It is sometimes remarkably hard
to get the sign correct out of this path integral. So by definition of electric potential,
Z
~ · d~ℓ
|∆Vab | = − E
path(a→b)
The sign of the potential difference will be the sign of the work you would do to move
a positive charge from a to b.
Reason About Work: The electric potential is work per unit charge, so if you would
have to do work to move a positive charge from point A to point B, then the potential
is higher at point B than point A, and ∆Vab > 0.
Example 17.10 Find the Electric Potential Difference from the Electric Field
Problem: A uniform spherical volume of charge has charge density ρ and outer radius redge . Compute the
potential difference between the outer edge of the sphere at redge and the center, ∆Vec .
Solution
Definitions
e
~ ≡ Electric field
E
+ ∆Vec ≡ Potential Difference Between Outside and Center
+ redge ≡ Radius of Sphere
+ r
+ c
+
+ +
+
Path of Integration
Strategy: Find the electric field. Use definition of electric potential, integrate it to get the magnitude of the
potential difference, then use potential as work per unit charge to get the sign.
(a) Solve for the Electric Field: In the volume charge, a Gaussian surface of radius r encloses a total charge
Qenc = 34 πr3 ρ. Using the form of Gauss’ law for spherical symmetry, this yields
4 3
~ = Qenc r̂ = 3 πr ρ r̂
E
4πε0 r2 4πε0 r2
~ = ρr r̂
E
3ε0
(b) Use Definition of Electric Potential: The electric potential difference between a point on the edge of the
sphere and the center is Z
∆Vec = − E~ · d ~ℓ
path
Separate the calculation into two parts: (1)the calculation of the magnitude |∆Vec | by doing the integral, (2)
computing the sign by reasoning about the work.
(c) Select Path of Integration and Perform the Integral: Select a path of integration that makes the
integral as easy to do as possible. Any path between the two points will yield the correct result. Draw the path
of integration on your diagram. Use the chosen path to convert the path integral to a simple integral. Choose a
path of integration from the edge to the center along a radius. The vector d~ℓ = −r̂dr, so E
~ · d~ℓ = −E(r)dr,
Z r=0
|∆Vec | = E(r) dr
r=redge
Z r=0 2
ρredge
ρr
|∆Vec | = dr =
r=redge 3ε0 6ε0
(d) Reason About the Sign: Use the definition of potential difference as work done per unit charge to fix the
sign. If ρ > 0, then the field points outward and we would do work on a positive charge to move it from a point
on the edge to the center, so if ρ > 0 ∆V > 0, so
2
ρredge
∆Vec =
6ε0
The answer is also correct if ρ < 0, because the sign of ρ takes care of the sign of the potential.
∆VAB between point A at (0, 2cm, 0) and point B at (0, 1cm, 0)?
Solution
(a) Use Definition of Electric Potential: The electric potential difference between point A and point B is
Z
∆VAB = − ~ · d ~ℓ
E
A→B
Separate the calculation into two parts: (1)the calculation of the magnitude |∆VAB | by doing the integral, (2)
computing the sign by reasoning about the work.
(b) Select Path of Integration and Perform the Integral: Select a path of integration that makes the integral
as easy to do as possible. Choose a straight path along the y-axis from point A to point B, therefore d~ℓ = −ŷdy.
To simplify notation let ~rA = (0, a, 0) where a = 2cm and ~rB = (0, b, 0) where b = 1cm. The path of integration
is drawn below. Substitute d~ℓ and E(y)
~ into the expression for potential and perform the integral.
Z b
~
∆VAB = − E(y) · (−ŷdy)
a
Z b 3 b
2
y γ 3 3
|∆VAB | = γy dy = γ = b − a
a 3 a 3
N 3 3
(150 Cm2 )((0.01m) − (0.02m) )
|∆VAB | =
= 3.5 × 10−4 V
3
(c) Reason About the Sign: Use the definition of potential difference as
work done per unit charge to fix the sign. Since γ > 0, the electric field points a
in the +ŷ direction, so we have to do work to move a positive test charge from
y = 2cm to y = 1cm, therefore ∆VAB > 0. Therefore,
Example 17.12 Electric Potential Difference from Potential for a Radial Field
Problem: A uniform spherical volume of charge has charge density ρ and outer radius redge . Compute the
potential difference between the outer edge of the sphere at redge and the center, ∆Vec .
Solution
Definitions
e
~ ≡ Electric field
E
+ ∆Vec ≡ Potential Difference Between Outside and Center
+ redge ≡ Radius of Sphere
+
r
+ c
+
+ +
+
Path of Integration
(a) Solve for the Electric Field: In the volume charge, a Gaussian surface of radius r encloses a total charge
Qenc = 34 πr3 ρ. Using the form of Gauss’ law for spherical symmetry, this yields
4 3
~ = Qenc r̂ = 3 πr ρ r̂
E
4πε0 r2 4πε0 r2
~ = ρr r̂
E
3ε0
(b) Use Definition of Electric Potential: Since the electric field points only in the r̂ direction, the definition
of electric potential difference allows the calculation of the electric potential through
Z
V (r) = − E(r)dr
where the constant of integration can be used to make it so the potential is zero at the reference point.
(c) Perform the Integral: Substitute the value of E(r) calculated from Gauss’ law.
ρr ρr2
Z
V (r) = − dr = − +C
3ε0 6ε0
where C is the constant of integration.
(d) Calculate the Potential Difference: The electric potential difference between the edge, redge , and the
origin is
2 2
ρredge ρredge
Vec = V (0) − V (redge ) = C − − +C =
6ε0 6ε0
where the sign will be correct if we did the calculus correct. Note, the constant of integration, which contained
our arbitrary choice of the zero of potential cancelled when we calculated potential difference.
We didn’t need to set the constant C to calculate the potential difference in a single region. The normal
choice for the reference point in a spherical problem is V (∞) = 0, but since the field is different for r > redge
the potential function calculated is incorrect if r > redge . We could however choose to report the potential with
the center of the sphere as the reference point. In this case V (0) = 0 → C = 0.
∆VAB between point A at (0, 2cm, 0) and point B at (0, 1cm, 0)?
Solution
(a) Use Definition of Electric Potential: Since the electric field points only along the coordinate axis, the
definition of potential difference can be used to calculate the potential as
Z
V (y) = − E(y)dy
where the constant of integral will be used to fix the zero of potential.
(b) Perform the Integral: Substitute the function given and perform the integral
γ
Z
V (y) = − γy 2 dy = − y 3 + C
3
(c) Calculate Potential Difference: Once again let ~rA = (0, a, 0) and ~rB = (0, b, 0). The potential difference
between point A and point B is
γ γ γ
∆VAB = V (b) − V (a) = − b3 + C − − a3 + C = (a3 − b3 )
3 3 3
Substitute,
N
150 Cm 2
∆VAB = ((2cm)3 − (1cm)3 = 3.5 × 10−4 V.
3
where the sign should be correct if we’ve integrated properly.
To complete the calculation of the potential above, we would have to set the constant C. For planar systems,
we usually choose V (0) = 0. With this choice C = 0. You can check either of the above calculations by using
E(y) = −dV /dy for potentials of only one variable.
where ∆x is the distance moved in the x direction. Since potential decreases as x increases, ∆V = −E∆x.
Potential Difference in Uniform Field: If the electric field is uniform in the x direction,
~ = E x̂, then the potential difference between a point xa and the point xb is
E
∆Vab = −E(xb − xa )
or the potential difference between two points a distance d apart in the x direction is
|∆V | = Ed.
Solution
We can substitute this electric field into the definition of potential difference to find the potential difference
between point A and B for any system of point charges
Z X kqi
∆VAB = − ~ ~
E(~rP ) · dℓ = ~
2 r̂iP · dℓ
riP
A→B i
but the thing in parenthesis is just the potential difference due to a point charge. So we can calculate any potential
difference simply by summing the potential difference due to each point charge. This can be made even cleaner
if we focus on potential instead. Since a point charge has a field that only depends on r, we can calculate the
potential of the point charge simply by integrating
kq kq
Z Z
V (r) = − E(r)dr = − 2
dr = +C
r r
The reference point usually chosen from a point charge is V (∞) = 0, so
Electric Potential of Point Charge: The electric potential, V1 (~rA ), at point ~rA due
to a point charge q1 at point ~r1 is:
kq1
V1 (~rA ) = ,
r1A
where r1A is the distance from point ~r1 to the point A. The reference point for this
potential is a point infinitely far from the charge.
The potential of a point charge depends only on the distance, r, from the charge.
For one point charge, all we need to compute the potential is the distance between the charge and the point
where we want to know the potential. No vectors, now isn’t life getting better all the time? We can compute the
potential of any system of charge simply by adding the potentials.
Electric Potential of a System of Point Charges: Because potential is additive, the
electric potential of a system of point charges qi is the sum of the potentials of the
individual charges
X X kqi
V (~rP ) = Vi (~rP ) =
i i
riP
where riP is the distance from the charge qi to the point P .
Solution
Definitions
y(cm)
4
q1 = 3muC ≡ Charge producing the potential
q1 ~r1 = (4cm, 2cm, 0cm) ≡ Location of q1
2
P ≡ Location where potential is computed
~r2 = (−3cm, −3cm, 0cm) ≡ point P
V1P ≡ Potential at P due to q1
-4 -2 r1P
2 4
x(cm) V0 ≡ Arbitrary constant in point charge equation
-2 r1P ≡ Distance from q1 to P
P -4
(a) Use Electric Potential of Point Charge: The electric potential at P from the field of q1 is
kq1
V1P =
r1P
(b) Compute distance between q1 and P: Observing the diagram, the distance between q1 and P is found by
applying the PythagoreanTheorem,
p
r1P = (−7cm)2 + (−5cm)2 + 0 = 8.6cm
The total potential function for a set of point charges is just the sum of the potential functions of the individual
point charges.
Solution
Definitions
ycm
4
Vi0 ≡ Potential at origin from qi
q1 = q2 = 1µC ≡ Point charges
2
q3 = q4 = 1µ ≡ Point charges
-4 -2 r20 r30 2 4 ~r10 = (4cm, 0, 0) ≡ Vector from 1 to 0
xcm ~r20 = (1cm, 0, 0) ≡ Vector from 2 to 0
q1 r10 q2 q3 r40 q4 ~r30 = (−1cm, 0, 0) ≡ Vector from 3 to 0
-2 ~r40 = (−4cm, 0, 0) ≡ Vector from 4 to 0
V0 ≡ Electric Potential at Origin
-4
(a) Write Equation for Total Potential: By Principle of Additivity of Electric Potential, the total potential is
the sum of the potentials from the individual charges. The total potential at the origin is
X kqi
V0 =
i
ri0
Solution
(a) Cut the System into Pieces: The system is drawn to the
right. Divide the system into small segments of length ∆x and
center xi . The charge of each segment is qi = λ∆x. The distance y
of each segment to the point P a distance R along the axis is
q P
di = R2 + x2i
xi x
-L L
∆xi
(b) Convert the Sum into an Integral: Let the segments become infinitely short to convert the sum into an
P RL
integral, ∆x → dx, xi → x, and i −L
L
kλdx
Z
V (R) = √
−L R2 + x2
Since the integral is even
L
kλdx
Z
V (R) = 2 √
0 R2 + x2
(c) Look up the Integral: I looked this one up in my ancient integral table
dx x
Z
√ = sinh−1 +C
2
R +x 2 R
(d) Use the integral formula:
L L
kλdx x
Z
V (R) = 2 √ = 2kλ(sinh−1 )
0 R2 + x2 R 0
−1 L
V (R) = 2kλ sinh
R
since sinh−1 (0) = 0.
By symmetry we know the field is in the y direction. We can then calculate the field by taking the derivative
of the potential
dV (y)
V (y) = − ŷ
dy
Solution
Since the potential only depends on y, the electric field along the y axis is
~ dV d −1 L
E=− ŷ = −2kλ sinh ŷ
dy dy y
√
I looked up the derivative d sinh−1 (x)/dx = 1/ 1 + x2 , so taking the derivative and using the chain rule,
~ d −1 L 1 L
E = −2kλ sinh ŷ = −2kλ p − 2 ŷ
dy y 1 + (L/y)2 y
~ = p2kλL ŷ
E
y y 2 + L2
If we take the limit, L >> y, we get back the electric field of an infinite line charge, so we’re OK.
III
I chose four lines per total charge of the volume charge. These
lines originate somewhere within the charge, stop on the conduc-
tor, and restart on the outer surface of the conductor. path of integration
+
_
+ _ _ +
I
_
II
+
III
(c) Region III: The charge enclosed by a Gaussian surface of radius r in region III is the total charge of the
volume change, since the conductor is uncharged:
4 3
Qenc = πa ρ
3
The electric field is given by the symmetry specific form of Gauss’ Law
4 3
~ III = Qenc r̂ = 3 πa ρ r̂
E
4πε0 r2 4πε0 r2
A Gaussian surface in the conductor must enclose zero charge because the field is zero. Therefore, the charge on
the inner surface of the conductor must be equal and opposite the charge of the volume charge
4
Qc,in = − πa3 ρ
3
The path of integration points opposite the direction of the field, so all the potential differences are positive. The
total potential difference from infinity to zero is
∆V∞0 = |∆VI | + |∆VIII |
where I have used the potential difference across the conductor as zero. The potential difference across region
III is Z b Z b 4 3
~ ~ 3 πa ρ
|∆VIII | = −
EIII · dℓ = −
dr
∞ 4πε0 r2
∞
ρr2 0 ρa2
|∆VI | = − =
6ε0 a 6ε0
where once again the absolute value was removed to yield a potential with the correct sign. The total potential
difference is then
4
πa3 ρ ρa2
∆V∞0 = |∆VI | + |∆VIII | = 3 +
4πε0 b 6ε0
which can be simplified to
a3 ρ ρa2
∆V∞0 = +
3ε0 b 6ε0
Kinetic Energy of a Point Mass: For a non-rotating or point mass, the kinetic energy
is
1
K = mv 2
2
where m is the mass and ~v is the velocity.
The law of conservation of energy applies to all forms of energy. A system of charges brings a new type of energy
to conserve, the electric potential energy. Let me remind you of the connection between potential and potential
energy.
Electric Potential Energy: The electric potential energy, U , of a charge, q0 , at some
location ~r0 , is the electric potential at that point V (~r0 ) multiplied by the charge,
U = q0 V (~r0 )
Change in Electric Potential Energy: The change in electric potential energy ∆VAB
of a point charge q0 when moved from a point A to a point B is
∆VAB = q0 ∆VAB = UB − UA
Ki + Ui = Kf + Uf
where K is the kinetic energy and U is the potential energy. If the potential energy
is only electric potential energy and there is only one charge moving q, this can be
rewritten in terms of the electric potential,
Ki + qVi = Kf + qVf
where Vi and Vf are the electric potentials at the initial and final locations of the charge.
Conservation of Energy in Terms of Potential Difference: If a charge, q, moves
from point A with velocity vA to a point B with velocity vB , the conservation of energy,
if we knew the potential at point A, VA , and point B, VB , would be
1 1
mv 2 + qVA = mvB
2
+ qVB
2 A 2
We can rewrite this in terms of the potential difference, ∆VAB = VB − VA , as
1 2 1 2
mvA = mvB + q∆VAB
2 2
After A Long Time: Potential problems will often ask for the energy or velocity after
a long time or very far from the charge. You need to figure out where the particle will
be after an infinitely long time, and use this location as the final location of the particle.
If a particle is infinitely far away from a charge distribution of finite size whose reference
point for potential energy is infinity, the particle has zero potential energy.
Solution
Definitions
ycm
q1 ≡ Charge at origin
q2 ≡ Charge moving
ri = 5cm ≡ Initial separation
q2 rf ≡ Final Separation
xcm
q1 V ≡ Electric Potential
5 10 vi = 0 ≡ Initial Velocity
vf ≡ Final Velocity
Kx ≡ Kinetic Energy at i and f
Ux ≡ Potential Energy at i and f
m = 1g ≡ mass moving
Trajectory
Strategy: Write conservation of energy equation and solve for vf .
(a) Use Conservation of Energy: Because the electric force is conservative, the total energy when the particle
is released is the same as the total energy after a long time. The total energy is the sum of the electric potential
energy plus the kinetic energy.
Ki + Ui = Kf + Uf
(b) Understand Final State: At a very long time, the positive charge moves infinitely far away from the fixed
positive charge, so rf ⇒ ∞.
1
(c) Select proper form of KE and PE: The kinetic energy of q2 is K = mv 2 . By definition of electric
2
potential, U = q2 V .
1 1
mvi2 + q2 V (ri ) = mvf2 + q2 V (rf )
2 2
(d) Select Proper Form of Potential: The electric potential of a point charge at the origin with V (∞) = 0 is
kq1
V (r) =
r
(e) Substitute Initial and Final Conditions: In most problems, some terms of the conservation equation will
simplify because of the initial and final conditions. Because the particle is released at vi = 0, Ki = 0. Since
rf ⇒ ∞, Uf ⇒ 0, therefore
kq1 q2 1
Ui = = mvf2 = Kf
ri 2
(f) Solve for vf : s
2
2(8.99 × 109 Nm 2
r
2kq1 q2 C2 )(1µC) m
vf = = = 19
mxi i (0.05m)(0.001kg) s
Solution to Part(a)
The charge on the car is positive so the electric field created by the charged walls must point from left to right.
Therefore you must do work to move a positive test charge from wall 2 to wall 1, so wall 1 is at a higher potential.
Wall 1 = 1000V . Wall 2 = 0V . We could also reason that a positive charge spontaneously moves to lower
potential so Wall 1 must have higher potential than Wall 2.
Solution to Part(b)
The car’s electric potential energy when it is at Wall 1 is converted into kinetic energy as it travels to Wall 2.
This energy is released as heat and goes into deformation when the car crashes. By definition of electric potential,
|U | = |q∆V | = (1C)(1000V ) = 1000J.
Solution to Part(c)
When it reaches the wall, it has lost 1000J of potential energy, which has been converted into kinetic energy so
−∆U = 21 mv 2 , where v is the velocity. Solving for v gives
r s
2|∆U | 2(1000J) m
v= = = 1.41 .
m 1000kg s
Solution to Part(d)
Electrostatic Energy
ET OT = K + U
The total electric potential energy of a system will be called the total electrostatic energy.
Energy Is Total Work: The total electrostatic energy of a system of stationary charges
is the total work an external agent has to do to assemble it leaving each particle with
zero kinetic energy. To compute the total energy, add the work required to add each
charge.
Sign of the Energy: The energy is positive if you have to do work to put the system
together, if you have to squeeze on it. The energy is negative if the system does work
on you, if you have to hold onto it to keep it from going together.
Visualize Building It: To compute the total energy of a system of charges, imagine
building it up charge by charge by hauling each charge, one at a time, in from infinity
and gluing them in place.
Solution
You can build this system by fixing the + charge at the origin, and then bringing the negative charge in from
infinity. As you bring the − charge in, it is pulled toward the positive charge (Opposites Attract) doing work on
you. Therefore you do negative work on the charge and the energy of the system is negative.
190
18.1. ELECTROSTATIC ENERGY CHAPTER 18. ELECTROSTATIC ENERGY
Solution
Definitions
y(cm)
-10 r23 10
x(cm)
Strategy: Bring one charge in from infinity at a time. Compute the work done (change in potential energy) as
each charge is brought in. Add it all up.
(a) Write Total Work: We’re going to assemble the system piece by piece and add up the work to place each
new charge. The total electrostatic energy of the system, U , is the total work done to assemble the system which,
is the sum of the work to place the individual charges. It doesn’t matter in what order we place the charges, the
total work will be the same. The total work to assemble the system is the work to place the first charge, W1 , plus
the work to place the second charge, W2 , in the field of the first charge, plus the work to place the third charge,
W3 , in the field of q1 and q2 .
U = W T = W1 + W2 + W3
(b) Place First Charge: The work to place the first charge q1 at location ~r1 is zero, W1 = 0, since none of
the other charges are present.
(c) Place Second Charge: Compute the work to place the second charge using the electric potential energy.
The work to place q2 in the electric field of q1 (the only charge placed so far) is the change in potential energy
to move q2 in from infinity with q1 fixed,
kq1
W2 = q2 ∆V1 (~r2 ) = q2
r12
where I have used the potential difference between infinity and a point r from a point charge. The distance is
r12 = 10cm by observation. Substitute everything:
2
kq2 q1 (8.99 × 109 Nm
C2 )(1µC)(1µC)
W2 = = = 0.09J
r12 0.1m
(d) Add the Third Charge: The work to place q3 in the electric field of q1 and q2 is the change in potential
energy to move q3 in from infinity with q1 and q2 fixed,
kq1 kq2
W3 = q 3 + q3
r13 r23
U = W T = W1 + W2 + W3
= 0 + 0.09J + 0.135J
U = 0.225J
This energy is stored in the electric field. The field is uniform, so the energy density must be uniform. The energy
density is the energy divided by the volume V = Aℓ between the planes. Let ue be the energy density of the
electric field
Q2 ℓ
ET OT 2ε0 A Q2 σ2 ε0 E 2
ue = = = = =
V Aℓ 2ε0 A2 2ε0 2
where I have used E = σ/ε0 .
Electrostatic Energy Density: The energy per unit volume, ue , stored in the electric
field is
1
ue = ǫ|E0 |2 ,
2
where ǫ is the permittivity of the material (ǫ0 for free space, ǫ = κǫ0 for a material
with dielectric constant κ.) and |E0 |2 is the modulus of the electric field squared,
~0 · E
|E0 |2 = E ~ 0.
This is rather profound. When we rubbed the golf tube with the oven bag, an electric field was produced
which propagated out into the universe. That field carried energy, so when we rubbed the golf tube we sent a
blast of energy travelling across the universe.
To calculate the total energy in some volume for a uniform field, multiply the energy density by the volume
(if the field is uniform, the energy density is the same at all points in the volume).
Solution
A 1m cube is so small compared to the Earth, that the field is approximately uniform in the cube. The energy of
~ 2
a region of the electric field is the energy density of the field ue = |E|2 ǫ0 , multiplied by the volume of the space,
so
~ 2 ǫ0 V −12 C2
|E| (150 N 2
C ) (8.85 × 10 Nm2 )(1m)
3
U= = = 9.96 × 10−8 J
2 2
Capacitance
Definition Capacitance: The feature of electrostatic systems that enters into electric
circuits is the capacitance C; if equal and opposite charge ±Q are placed on two
conductors a potential difference ∆V is produced. The capacitance of the system of
conductors is
C = Q/∆V.
Capacitance is defined as the ratio of the charge on one of the conductors to the
potential difference between two conductors carrying equal and opposite charge.
Capacitance Does Not Depend on Charge or Potential Difference: If the restric-
tions below are met, C does not depend on either Q or ∆V , but only on the geometry
of the conductors and dielectrics, their size and shape, dielectric constants, and ε0 . Ca-
pacitance is independent of Q and ∆V , if when all excess charge is removed from the
conductors, there is no remaining charge or field in the system. That is, the conductors
are not in the presence of fixed charge or external fields.
Units of Capacitance: The SI unit for capacitance is the farad F . The farad is related
to other units by
C J
1F = 1 = 1 2 = 1C2 /Nm
V V
New Units for Permittivity: In capacitance problems it is often convenient to use
ε0 = 8.85 × 10−12 F/m.
194
19.1. DEFINITION OF CAPACITANCE CHAPTER 19. CAPACITANCE
Solution
The capacitance of a parallel plate capacitor is given by
ε0 A
C=
d
Solve for the plate area
dC (1 × 10−4 m)(1000 × 10−6 F)
A= = C2
ε0 8.85 × 10−12 Nm 2
A = 1.13 × 104 m2
+ _
+ + + + + +
_ _ _ _ _ _
No charge actually passes between the plates. In the simplest capacitors, the area between the plates is filled
with air and the only way charge can pass between the plates is through a spark. Once a capacitor sparks, it
is usually shot. Commercial capacitors are rated both by their capacitance and the maximum voltage they can
support, ∆Vmax . The maximum voltage is crucial to the amount of energy a capacitor can store.
A capacitor, when charged, has a positively charged plate separated from a
negatively charged plate. To produce this separated charge an external agent
had to do work, so the capacitor contains energy. We can imagine building
the system of charge by moving chunks of charge from one plate to the other.
This isn’t what happens in practice and I’m not sure how you would actually ∆V
do it, but for a thought experiment it is only important that it could be done ∆q
in principle. The work to move the ith chunk ∆q is ∆q∆Vi where ∆Vi is the
potential difference the ith chunk is moved through.
The total work to build up a charge of +Q on one plate and −Q on the other plate is
Q
qi qdq 1 Q2
X X Z
W = ∆q∆Vi = ∆q = =
i i
C 0 C 2 C
where I used the definition of capacitance C = Q/∆V . This is one of my favorite formulas.
1 1 1 Q2
U= Q∆V = C∆V 2 =
2 2 2 C
where each version of the formula is just a result of applying the definition of capacitance
C = Q/∆V .
(a)If the capacitance is increased, does the power supply deliver or remove positive charge from the
positive capacitor plate?
(b)Is the energy that is stored in the capacitor increased or decreased?
Solution to Part(a)
Capacitance is the charge stored per potential difference, therefore capacitance multiplied by potential difference
is the charge stored on one capacitor plate. If the capacitance increases, the charge stored on the positive plate
increases, Q = C∆V .
Solution to Part(b)
Solution
The energy stored in a capacitor is given by U = 21 C∆V 2 , so the potential difference needed to store a given
amount of energy is s
r
2U 2(3.6mJ)
∆V = = = 12V
C 50µF
I love capacitance because it allows us to calculate the energy of complicated systems without having to
integrate the energy density over all space. A capacitor has charge densities and fields. If you want to calculate
the energy of a system of charge and the charge and field of that system happens to match the charge and field of
a capacitor you can use the energy of the capacitor to calculate the total energy of the system. After all, there is
only charge and field. If an isolated spherical capacitor with respect the ground is charged to a charge Q, all that
Q is at the surface of the sphere forming a surface charge density. The field in the sphere is zero and the field
outside the sphere is Q/4πε0 r2 . This is exactly the system of charge and field I calculated the total energy of at
the end of last chapter by integrating the energy density. The capacitance of an isolated sphere is C = 4πε0 R,
where R is the radius. If I use the energy of a capacitor, I get a total energy of
1 Q2 1 Q2 Q2
U= = =
2 C 2 4πε0 R 8πε0 R
exactly what I calculated last chapter with a ton less work.
Solution to Part(a)
The electric field due to the charge is reduced by a factor of κ when the dielectric is inserted, therefore the
potential difference is reduced by κ, ∆V = ∆V0 /κ.
Solution to Part(b)
The charge remains the same, so the new capacitance is C = Q/∆V = κQ/∆V0 = κC0 . The capacitance is
increased by a factor of κ.
Solution to Part(c)
Solution to Part(d)
Solution to Part(a)
Solution to Part(b)
The dielectric increases the capacitance by a factor of κ, so the new capacitance is Cκ = κC = (4)(67µF) =
268µF
Solution to Part(c)
Again by definition of capacitance, ∆Vκ = Q/C = (100µC)/268µF = 0.38V. We could have also used ∆V0 /κ.
• Add a charge +Q to one of the conductors and −Q to the other. Capacitance does not depend on the
total charge of the system, so this arbitrary charge must cancel out of the capacitance.
Conductor
Strategy: Place ±Q charge on the inner and outer conductor. Compute the field using Gauss’ Law, correcting
for the presence of the region II conductor. Integrate the field to get the potential difference. Apply the definition
of capacitance.
Solution to Part (a)
(a) Draw a Good Diagram: Select four field lines going out for the charge Q on the central conductor, so
region I has four field lines. All these lines end on the conductor in region II and begin again on the outer surface
of the conductor. Region III has four field lines since it encloses Q net charge.
(b) Compute the Electric Field: The charge enclosed in region I is Q, so using Gauss’ Law in spherical
coordinates
~I = Q
E r̂.
4πε0 r2
A Gaussian surface in region III encloses a charge Q. The electric field in region III has the same functional form
as that in region I,
E~ III = Q r̂.
4πε0 r2
Compute the Potential Difference Between the Inner Conductor and Middle Conductor: By definition
of potential, Z r2 Z r2
~
Q
|∆VI | = −
EI · r̂dr = −
2
dr
r1 r1 4πε 0 r
r2
Q
= Q
1 1
|∆VI | = −
4πε0 r r1 4πε0 r2 r1
Use a positive potential to yield a positive capacitance,
Q 1 1
∆VI = −
4πε0 r1 r2
Compute the Potential Difference Between the Middle Conductor and the Outer Conductor: By
definition of potential, Z r4 Z r4
~
Q
|∆VIII | = −
EIII · r̂dr = −
2
dr
r3 4πε0 r
r3
r
Q 4 Q
1 1
|∆VIII | = = −
4πε0 r r3 4πε0 r4 r3
Select a sign for the potential consistent with the choice for region I.
Q 1 1
∆VIII = −
4πε0 r3 r4
(a) Compute ∆VII : The potential difference across the conductor is zero, ∆VII = 0
(b) Compute Total Potential Difference: Since the field points in the same direction in all regions the
magnitudes of the potential difference add, so ∆V = ∆VI + ∆VII + ∆VIII .
Q 1 1 Q 1 1
∆V = − +0+ −
4πε0 r1 r2 4πε0 r3 r4
(c) Apply Definition of Capacitance: The capacitance is defined as
Q Q
C= =
∆V Q 1 1 Q 1 1
4πε0 r1 − r2 + 4πε0 r3 − r4
4πε
C= 0
1
r1 − r12 + r13 − 1
r4
4πε
C= 0
1 1 1 1
5cm − 7.5cm + 17.5cm − 20cm
C = 15.1pF
Example 19.7 Parallel Plate Capacitor with Dielectric Partially Filling Airspace
Problem: The conducting plates of a parallel plate capacitor each have an area of 2.0m2 . They are separated
by 4.0mm of air. A dielectric slab, with constant κ = 2.5 and width 2mm, is placed halfway between the plates.
What is the capacitance of this arrangement?
Solution
Definitions
x
I II III
A = 2m2 ≡ Area of Conducting Plate
σ ≡ Charge Density on Conducting Plate
_ + ~ i ≡ Electric Field in Region i
E
+Q ∆V ≡ Potential Difference between Conductors
_ -Q
+ C ≡ Capacitance
κ = 2.5 ≡ Dielectric Constant
_ + di ≡ Distance across region i
Strategy: Compute the electric field with the dielectric removed, then correct the field for the presence of the
dielectric. Compute the potential difference using the formula for a uniform field, then apply the definition of
capacitance.
(a) Place Charge on the Conductors: Place +Q on the left conductor and −Q on the right conductor. The
area of the conductor is A = 2m2 , therefore the charge density on the left conductor is σ = +Q/A and the charge
density on the right conductor is −σ = −Q/A.
(b) Compute the Electric Field: Under the assumption that the plate area is infinite, the electric field of two
equal and oppositely charged planes is
~ 0 = σ x̂ = E
E ~I = E ~ III .
ε0
This can be computed either from Gauss’ Law or by superimposing planes. This field is reduced in the dielectric
by a factor of κ = 2.5,
~ II = σ x̂.
E
κε0
(c) Compute the Potential Difference between the Conductors: The potential difference across a region
~ is just |∆V | = |E∆x|, where ∆x is the distance across the region. Since all of the fields
with uniform field, E,
point in the same direction, the potential difference between the conductors is
Using σ = Q/A,
Q d2
|∆V | = (d1 + + d3 ).
Aε0 κ
(d) Apply Definition of Capacitance: By definition,
Q Q
C= = Q
.
∆V Aε0 (d1 + dκ2 + d3 )
2
C
Aε0 (2m2 )(8.85 × 10−12 Nm 2)
C= d2
= 0.002m
d1 + κ + d3 0.001m + 2.5 + 0.001m
C = 6.32 × 10−9 F = 6.32nF
Definitions
Gaussian
Surface
shell
a = 0.4mm ≡ Radius of inner wire
foamed Teflon
b = 2.2mm ≡ Inner radius of outer conducting shell
EII ≡ Electric Field in region II
L = 3.05m ≡ Length of Cable
r a
b ℓ ≡ Length of Gaussian Surface
wire
Q ≡ Arbitrary Charge placed on Conductor
C ≡ Capacitance
r ≡ Radius from axis of system
Path of Integration
Strategy: Introduce an arbitrary charge. Use Gauss’ Law to compute the electric field, then apply the definition
of electric potential to a path of integration between the conductors. Apply definition of capacitance.
(a) Place Charge on the inner and outer conductor: We will assume the cable is infinitely long to compute
the electric field and potential difference. To compute the capacitance, we first place +Q on the inner wire and
−Q on the outer shell. This creates a charge per unit length on the inner wire of λ = Q/L.
(b) Compute the Electric Field: For cylindrical geometry, Gauss’ law reduces to
Qenclosed
E=
2πε0 ℓr
where ℓ is the length of the Gaussian cylinder. In region II, which is the only region where we need the field, the
total charge enclosed in a Gaussian cylinder of length ℓ is λℓ = Qℓ/L. So the electric field in region II if it were
filled with air is
Qℓ/L Q
EII = =
2πε0 ℓr 2πε0 Lr
(c) Compute the Potential Difference: The potential difference from the outer surface of the wire at radius
a to the inner surface of the outer conductor at radius b is
Z b
~
|∆Vab | = −
EII · r̂dr
a
Z b
Q
= − dr
a 2πε0 Lr
Z b
Q dr Q
|∆Vab | = − =− (ln(b) − ln(a))
2πε0 L a r 2πε0 L
(d) Apply Definition of Capacitance: By definition, the capacitance of the cable is
Q Q
C= = Q
|∆Vab | 2πε0 L (ln(b) − ln(a))
2πε0 L 2πε0 L
C= =
ln(b) − ln(a) ln(b/a)
Compute the Capacitance of the Cable in Lab: Substitute the values measured for the cable in lab into the
general expression derived above,
2
C
2πε0 L 2π(8.85 × 10−12 Nm 2 )(3.05m)
C= =
ln(b/a) ln(2.2mm/0.4mm)
Compute the Dielectric Constant of Foamed Teflon: If the dielectric completely fills the region where the
electric field exists, then the dielectric increases the capacitance from C to Cκ , where Cκ is the capacitance with
the dielectric. Therefore the dielectric constant of foamed Teflon is
Cκ 170pf
κ= = = 1.71
C 99pF
(a)If the field is established by connecting a power supply between the plates, what potential difference
must the power supply provide?
(b)What is the charge density on the left plate?
(c)Draw the electric field and the location and sign of any bound charge in the system.
A dielectric slab of thickness d/4 with dielectric constant κ = 3 is inserted between the plates without
disturbing the surface charge densities on either plate as shown in figure (b).
(d)Compute the bound charge densities on the left and right surface of the dielectric, σb,l and σb,r .
(e)If the left and right plates are conductors, compute the capacitance of the system. Leave the answer
symbolic, do not compute a number.
d d/4
Solution to Part(a)
~ by
For a uniform field, the potential difference, ∆V , is related to the electric field, E,
N
|∆V | = |Ed| = (10000 )(0.02m) = 200V
C
where d is the separation of the plates.
Solution to Part(b)
The magnitude of the electric field of two equal and opposite planes of charge is
σ
E=
ε0
Therefore the charge density on the left plate is
C2 N C
σ = ε0 E = (8.85 × 10−12 2
)(10000 ) = 8.85 × 10−8 2
Nm C m
Solution to Part(c)
_ +
_ +
_ +
I II III
Solution to Part(d)
The system is divided into regions in the previous step and a cylindrical Gaussian surface with end area A is
drawn. The electric field for the three regions are
~ III = σ x̂
~I = E
E ~ III = σ x̂
E
ε0 κε0
~ I = EI x̂ and E
Apply Gauss’ law to the Gaussian surface drawn. Let E ~ II = EII x̂.
Qenclosed
EII A − EI A =
ε0
The charge enclosed in the surface is the bound charge at the left side of the dielectric,
Qenclosed = σb,l A
Solution to Part(e)
If the dielectric has thickness d/4 then the distance across region I and III is d−d/4
2 = 3d
8 . The potential difference
between the two plates is the sum of the potential differences across the regions. The potential difference across
the single region is, because of the uniform field,
3d 3dσ
∆VI = EI = = ∆VIII
8 8ε0
d dσ
∆VII = EII =
4 4κε0
Therefore the total potential difference is
3dσ dσ 3dσ
∆V = ∆VI + ∆VII + ∆VIII = + +
8ε0 4κε0 8ε0
The total charge, Q, on the plates is the plate area Ap multiplied by the charge density, Q = σAp . Apply the
definition of capacitance,
σAp
C = 3dσ dσ 3dσ
8ε0 + 4κε0 + 8ε0
ε0 Ap 1
C= 3 1
d 4 + 4κ
Solution
Definitions
R = 1m ≡ Radius of Sphere
Q ≡ Charge on Sphere
+
+ + ~r ≡ Radius Vector
∆V ≡ Potential difference between sphere and ∞
+ + ~l ≡ Path of integration
I
+Q C ≡ Capacitance of Sphere
+ +
II +
Path of Integration
Strategy: Place Q charge on the conductor, compute the electric field and integrate from infinity to the
conductor to get the potential difference, and apply the definition of capacitance.
(a) Place Arbitrary Charge +Q on the Conductor: Put a charge, Q, on the conductor.
(b) Compute the Field, E: ~ The electric field outside the conducting sphere is the same as that of a point
charge, Q, located at the center.
~ = Q r̂
E
4πε0 r2
(c) Integrate the Field to get ∆V : Integrate the electric field between the conductor and the ground to
get the potential difference between the surface of the conductor and infinity. Draw the path on the diagram.
Integrate the electric field from the zero potential at infinity to the surface of the conductor, r, to obtain the
electric potential. By definition of electric potential
Z
∆V = − E~ · d~l
path
Compute the magnitude, then select the sign to give a positive capacitance. For the path drawn on the diagram
d~l = r̂dr, the limits are R and ∞,
Z ∞
~
|∆V | = −
E · (r̂dr)
ZR∞ Z ∞
Q Q 1
=−
r̂ · (r̂dr) =
dr
R 4πǫ0 r2 4πǫ0 R r2
∞
Q 1
= −
4πǫ0 r R
Q
=
4πǫ0 R
Q Q
C= = = 4πε0 r
∆V Q/4πε0 r
The charge Q you introduced had better cancel out or you have made a mistake.
(e) Substitute and Calculate:
C2
C = 4π 8.85 × 10−12 (1.0m) = 1.1 × 10−10 F = 0.11nF
Nm2
20.1 Current
20.1.1 Current
Since the first day of this class, we have been moving electric charge from place to place. Previously, we have
focused on the effects of moving charge so as to produce a net static charge. Now, we focus on the effects of
moving charge through an electric field, where no net electric charge is produced. Suppose we place a conductor
in an electric field, charge flows until the electric field in the conductor is zero as shown in figure (a) below. The
field is zero because of the induced surface charge. Now, let’s attach a pump and pump the induced charge away
as fast as it is created. Since induced charge is equal and opposite, we can do this by pumping the + induced
charge off the right side of the conductor, through the pump, and onto the left side of the conductor. Since there
is no induced charge, there is an electric field in the conductor and a potential difference across the conductor.
We will call a flow of charge an electric current and a system of conductors and other elements that potentially
allow charge to flow, an electric circuit. A circuit is closed when charge flows and open when there is a barrier to
charge flow.
Suppose we have someone, an observer, measure how much charge ∆Q flows past some point in the circuit
above in a time ∆t. We define the electric current I = ∆Q/∆t.
208
20.1. CURRENT CHAPTER 20. CURRENT AND RESISTANCE
∆Q
I=
∆t
where a charge ∆Q moves through a cross-section in the time ∆t.
Units for Current: The SI unit for current is the ampere, A. The ampere is related to
other units by
C
1A = 1
s
What does this mean? Figure (a) shows charged particles moving
through space. In this case, the cross-section, labelled A, is a (a) Charged Particles Traveling through Space
closed hoop in space. If a lightning bolt went through the center
of a hula-hoop, it is a current through the hoop. Figure (b) shows A
charges flowing in a wire. In this case, the natural closed hoop to
use for the current is the cross-section of the wire.
I
(b) Charges Moving Through a Wire
Positive Current is the Direction of Positive Charge Flow : Positive charge moving
in a certain direction constitutes a positive current in that direction. Negative charge
moving in the opposite direction also constitutes a positive current in that direction.
_
+
_
+ _
+
I I
A current may be formed of charge particles moving through space or from charges moving in a wire. If the
same amount of charge moves through a cross-section at the same time, the current is the same. There is one
big difference, the charged particles moving through space have a net charge, and therefore an electric field. The
current in the wire moves against the background of the positive atomic cores and the wire has no net charge,
and no electric field.
Current Density: Currents flow in wires or in extended regions of space, we can define
a current density j by dividing the current, I, by the area, A, its flowing through,
I
j=
A
If you have the current density, the current is found by multiplying by the cross-sectional
area.
The behavior of a electric circuit changes dramatically if the magnitude or direction of the current is changing
with time.
DC Circuits: Direct Current (DC) circuits are circuits where the current always flows
in the same direction.
The simple circuits we build with batteries and light bulbs in lab are direct current circuits.
Alternating Current Circuits: Alternating Current (AC) changes direction with time.
The current delivered from a wall plug is alternating current, a sine wave with frequency 60s−1 .
+ A _
A B A B
I I
A _ B
V +
A B
∆VAB
In Activity 12 Series and Parallel Circuits, we will build a simple circuit to light a light bulb with a battery.
The figure below shows how the physical circuit is represented by a circuit diagram.
The figure to the right shows the same current flowing at all points
in the bulb/battery circuit. The corners of the circuit have been
labelled. Because the potential difference around any closed loop I
is zero, a basic law of the universe, we can write the sum of the B C
potential differences around the circuit.
∆VAB = −∆VCD A D
I
We have already said that the effect of a battery is to establish, ∆Vbatt , voltage between the terminals of the
battery. Therefore, from A to B, the potential difference is ∆VAB = ∆Vbatt > 0. The potential difference is
positive, so the potential increases across the battery. This is called a potential rise. The reference point for the
potential is any convenient point in the circuit, usually the negative side of the battery. If ∆VAB is positive then
∆VCD is negative. Therefore, the potential decreases from point C to point D. This is called a potential drop.
∆W ∆Q∆VAB
= = I∆VAB
∆t ∆t
Work per unit time is power, P . The analysis above depended only on charge moving through potential difference
and is completely general.
Power Dissipated by a Current: The rate at which energy is dissipated or stored in
an object in a circuit is the product of the current flowing through the object and the
potential different across the object
P = I∆V.
This is also the expression for the power provided by the batteries and capacitors in a
circuit.
Definition of Power: Power, denoted by the symbol P , is the rate of doing work or
providing energy.
Units of Power: The units of power are Watts and are related to other units by
J
1W = 1
s
Energy and Power: Power is energy per unit time. If a device provides or consumes
power, P , at a constant rate, then the energy provided or consumed by the device in
time t is
U = Pt
If the power changes with time, P (t), the energy provided becomes
Z
U = P (t)dt
A battery or capacitor provides energy to a circuit when the current flows from lower to higher potential. A
battery removes energy from the circuit when the current flows from higher to lower potential. When a battery
or capacitor removes energy from the circuit, the energy is stored. We call this charging the battery or capacitor.
I I
The sign of the potential difference, ∆VCD , as the current goes through the light bulb is negative, the light
bulb does negative work on the current, or equivalently, the current does positive work on the light bulb. The
current provides energy to the light bulb that the bulb turns into heat and light. This energy is lost to the
environment and the light bulb is said to dissipate energy. The rate at which the energy is lost is the power. We’ll
be sloppy and talk about power dissipated, but what that will mean is, the rate at which the energy is dissipated.
Conservation of Energy for Circuits: Energy is conserved. The power provided, Pin ,
to a circuit must equal the power dissipated by a circuit, Pout , added to the power
stored in the circuit, Pstored ,
Solution
We can use the definition of resistance in the general expression for power.
P = I∆V = I 2 R
Solution to Part(a)
The power dissipated by any device is always P = I∆V , so I = P/∆V = 100W/110V = 0.91A.
Solution to Part(b)
A light bulb’s resistance changes as it heats up, so the resistance you measure when it’s cold is different than its
operating resistance, therefore the light bulb does not obey Ohm’s Law. The definition of resistance still applies
though. This light bulb’s resistance at its operating temperature is, by Definition of Resistance, R = ∆V /I =
110V/0.91A = 121Ω
Eventually, the charges must smash into something or the velocity would become infinite. Let the time τ be the
average time the charge has been travelling since its last collision. The average velocity of the charges is
qEτ q∆V τ
vd = aτ = =
m mℓ
and is called the drift velocity. If the density of mobile charges is n, then the current density in the material is
j = qnvd and the total current in the conductor
nq 2 ∆V τ A
I = jA = qnvd A =
mℓ
Rearranging gives,
∆V m ℓ
=R=
I nq 2 τ A
where I have used the definition of resistance. The stuff in parenthesis does not depend on the shape of the
conductor. It represents the conductive material’s intrinsic resistance, and is called the resistivity, ρ,
m
ρ=
nq 2 τ
So as the number of charge carriers goes up the resistance goes down. As the time between collisions goes up
the resistance goes down. As temperature increases, the atoms of the metal shake around more, and the moving
charges collide more often. Therefore, resistance increases with temperature.
Calculation of Resistance from Resistivity: The intrinsic feature of the material
which resists current flow is called its resistivity, ρ. The cross-sectional area, A, of
the material and the distance, L, through which the current flows also play a role in
determining the resistance of a material. The resistance of a material whose resistivity
is independent of current is related to its geometry by
L
R=ρ
A
Solution
Let LA be the length of wire A and LB = 2LA the length of wire B. Let rA be the radius of wire A and
rB = rA /2. The resistance of a wire is given in terms of the resistivity, ρ, by R = ρL
A , so the resistance of wire
A is
ρLA ρ LA
RA = = 2
AA π rA
ρLB ρ LB ρ 2LA ρ LA
RB = = 2 = 2
=8 2
AB π rB π (rA /2) π rA
1
RA = RB
8
Solution
The resistance is given by
ρL (10 × 10−8 Ωm)(3m)
R= = = 3 × 10−3 Ω
A (0.01m)2
where A is the area of the end and L is the length.
Conduction and resistance are intrinsically quantum mechanical phenomena. The mobile electrons travel as
waves through the material. To a first approximation the mobile electrons completely ignore the atoms and
other mobile electrons. They have the energies associated with a single free particle trapped in a box, where the
conductor is the box. Just like filling up atomic orbitals in chemistry, two conduction electrons cannot occupy the
same state, by the Pauli exclusion principle. So as conduction electrons are added to the system, each must be
added to a different state, with ever increasing energy. So what we actually have is a quantum mechanical gas of
Avogadro’s number of electrons pretending each other aren’t there.
The crystal structure of the metal further complicates things. The waves become grouped into energy bands.
Filled bands are insulating. Partially filled bands allow conduction. Since bands are filled in order of energy, only
the highest energy band conducts and is called the conduction band. Sometimes the conduction band is mostly
empty, populated only by charges that are thermally excited from a lower band. Since the conduction band is
mostly empty, the density of charge carriers in low, and the material is a conductor, but a poor conductor or
semi-conductor. Semi-conductors have been very important technically because their conduction properties can
be adjusted by adding impurities. This process is called doping.
DC Circuits
Voltage vs. Current for Light Bulb Voltage vs. Current for Resistor
V V
I I
Ohm’s Law: In some cases, the resistance of a material remains constant as the poten-
tial difference or current is changed. In such cases, the material obeys an experimental
relationship called Ohm’s Law
∆V = IR
Resistance Abbreviations: Resistances are often large numbers. We will use the
following abbreviations for large resistances:
218
21.1. OHM’S LAW CHAPTER 21. DC CIRCUITS
Recognize when Ohm’s Law is valid : Ohm’s Law is an experimental law and is
valid only for materials which have a constant resistance for varying potential differences
across them. If the ratio ∆V /I is not constant for the material in question, the material
is not Ohmic and does not obey Ohm’s Law.
Solution to Part(a)
Convert the units on resistance, 4.0MΩ = 4.0 × 106 Ω. Apply Ohm’s law,
∆V 12V
I= = = 3 × 10−6 A = 3µA
R 4.0 × 106 Ω
Solution to Part(b)
Solution to Part(c)
Since the power is constant, the energy dissipated, U , is Power multiplied by time
Solution
Note this is not your biggest problem. The voltage drop across the wrench is
A car battery has potential difference 12V, so 12V − 0.153V = 11.85V drops across the internal resistance within
the battery. Therefore the heat energy deposited in the battery per second is
∆V I = (11.85V)(600A) = 7110W
All this energy is heating the sulfuric acid in the battery, which will hurt when it hits you in the face.
∆Vbatt
If no current is drawn from the battery, the voltage across the battery is equal the that of the perfect
battery, ∆Vbatt = ∆Vterm . A perfect voltmeter draws no current, so connecting a voltmeter across a battery
measures ∆Vbatt . This does not give you an idea of how what the terminal voltage will be at the operating
current. As current is drawn from the battery, pumped through the battery, the terminal voltage decreases,
∆Vterm = ∆Vbatt − Ir. Once the terminal voltage drops to zero, you are drawing the maximum current the
battery can provide.
Solution
The potential difference is the same. Since I = ∆V /R and P = I∆V , the current and the power dissipated
cannot be the same.
R3
R1
Solution
R1 has no simple series/parallel relationship to R2 . R1 does not carry the same current as R2 , so they are not in
series. R1 and R2 are not connected so that each end of each device is at the same potential, so they aren’t in
parallel.
Problem: The circuit shown at the right is composed of identical light bulbs.
Reason, without calculating, the relation between the currents through B1 and B1
B2 .
B3 B2
Solution
B1 is in series with the combination B2 and B3 . The current through B1 is the same as the TOTAL current
through B2 and B3 . Since the bulbs are identical, half the current flows through each bulb, so B2 carries half as
much current as B1 .
Rs = R1 + R2
I2
The sum of the currents through each resistor is the same as that
coming into the parallel network due to conservation of charge. R2 Rp
R1 ∆Vp
I1 + I2 = Ip . ∆V1 I2 ∆V2
I1 Ip
In some sense, a parallel network increases the area through which
the current travels, which decreases the resistance. We have Ip =
∆Vp /Rp which means we have common numerators. To add the
currents, we need a common denominator, so we must invert the
expression. Thus we can write the equivalent resistance, Rp ,
1 1 1
= +
Rp R1 R2
Solution
Unscrewing either B2 , B3 , or B4 will cause B1 to dim, because the resistance of the combination B2 , B3 , B4
will be increased.
Solution to Part(a)
When many resistors with the same resistance are in parallel, the equivalent resistance is the resistance of one
divided by the number of resistors, using the equivalent resistance of a parallel combination
1 1 1 1 1 1 1 6
= + + + + + =
Req R R R R R R R
R 100Ω
Req = = = 16.7Ω
6 6
Solution to Part(b)
Solution to Part(c)
∆V (∆V )2 (6V )2
P = ∆V = = = 2.2W
Req Req 16.7Ω
Finally, let’s look at a circuit which has both series and parallel combinations of resistors.
Solution
Strategy: Working from the smallest elements, use series and parallel equations to replace them with equivalents
and keep working outward.
R1 R2
R4
1 1 1 1 1 1
= + = + = Rp Rs
Rp R1 R2 60Ω 120Ω 40Ω
Rp = 40Ω
Use the series formula to reduce R3 and R4 to their equivalent Rs :
1 1 1 1 1
= + = +
Req Rp Rs 40Ω 20Ω
40
Req = Ω Req
3
Don’t get carried away and reduce a part of the circuit which
is not a simple combination. Don’t do more than one step at
a time, or you will make mistakes.
Strategy: Progressively reduce individual series and parallel combinations, and then redraw. Use Ohm’s Law
and the power dissipated by an Ohmic device to compute the properties of the individual elements.
the current through and ∆Vi the voltage across resistor Ri . When a resistor
I0
is an equivalent, i will be the numbers of the resistors that went into the
R12 R3
equivalent. The value of R12 is given by
∆V0
1 1 1 1 1
= + = +
R12 R1 R2 100Ω 100Ω R5 R4
R12 = 50Ω
(b) Reduce the Two Series Combinations: Resistors R12 and R5 are in
series and resistors add in series, so their equivalent resistance is a
(c) Reduce the Final Parallel Combination: Resistors R125 and R34 are
in parallel and resistors in parallel add, using the formula a
1 1 1 1 1 7 I0
= + = + =
Req R125 R34 150Ω 200Ω 600Ω
600Ω
Req = = 85.7Ω
7
(a) Compute the Current in the Parallel Branches: Both R125 and R34 have the full ∆V0 = 10V across
them, so by Ohm’s Law
∆V0 10V
I125 = = = 0.067A = I5 = I12
R125 150Ω
(b) Compute the Voltage Drops in the Series Circuits: Since we have the currents in the series circuits, we
can use Ohm’s Law to compute the voltage drops,
As computed above
∆V0 10V
I34 = = = 0.05A = I3 = I4
R34 200Ω
As computed above,
∆V5 = I5 R5 = (0.067A)(100Ω) = 6.7V
As is often the case, these circuit reductions have their own pattern of solution which should be follow regardless
of the order in which the circuit properties are asked in the problem.
Kirchhoff’s Loop Equation: If there are no significant changing magnetic fields the
integral of the electric potential around any closed loop is zero. Therefore, if the wires
in a system have negligible resistance (support negligible potential differences), the sum
of the potential drops or gains across the devices in a circuit must be zero for any loop:
X
∆V = 0
loop
Later on, we will find that changing magnetic fields are responsible for the spark you
sometimes see when unplugging a device.
f e d
f e d
e d
f
We can reason about the sign of these potential differences physically. We already know the magnitudes. The
magnitude of the potential difference across a battery is the battery voltage. The magnitude of the potential
difference across a resistor is given by Ohm’s law, IR. Batteries are easy, if the potential difference goes from
negative to positive, it is positive. If this is reversed, it is negative. For resistors, we know that the current loses
energy as it goes through the resistor, so the potential goes down in the direction the current flows.
With these observations, we can make a table of sign conventions for the sign of the potential differences.
Sign Convention for Potential Changes in a Kirchhoff Loop: When applying Kirch-
hoff’s Loop rule, there is a distinction between a potential drop and a potential in-
crease(rise). When traversing the loop in the direction chosen for the loop, the sign of
the potential change will either be negative (a potential drop) or positive (a potential
increase) according to the following conventions:
Because of conservation of charge and the because surface junctions store minimal charge the current into a
junction is equal to the current out of a junction, so
I1 + I2 = I3
∆V1 ∆V2
I3 I2
a e d
I1
(b) Write Loop Equation 1: The potential difference around any closed path is zero, so
To move from a to b, one must move from the + to − end of battery 1, so ∆Vab = −∆V1 . To move from b
to j, we move through a resistor in the direction of current, so the potential decreases, ∆Vbj = −I1 R. To move
from j to e, one must move from the + to − end of battery 2, so ∆Vje = −∆V2 . Finally, to move from e to a,
we move through a resistor in the direction of current, so the potential decreases, ∆Vea = −I1 R. Substitute to
produce the first loop equation,
−∆V1 − I1 R − ∆V2 − I1 R = 0
∆V1 + ∆V2 + 2I1 R = 0 Loop 1
To move from e to j, one must move from the − to + end of battery 2 increasing the potential, so ∆Vej = +∆V2 .
To move from j to c, we move through a wire , ∆Vjc = 0. To move from c to d, one must move through two
resistors opposite the direction of current, so potential increases in each resistor, therefore ∆Vcd = +I2 R + I2 R.
To move from d to e, we again move through a wire so the potential does not change, ∆Vde = 0. Substitute to
produce the second loop equation,
∆V2 + 0 + 2I2 R + 0 = 0
∆V2 + 2I2 R = 0 Loop 2
(a) Look for Simplifications: Before punching the three equations with three unknowns into your calculator,
see if there are simplifications. In this problem, loop 1 only depends on current 1 and loop 2 only depends on
current 2.
(b) Solve for I1 : Solve loop equation 1 for I1 ,
∆V1 + ∆V2 18V + 6V
I1 = − =− = −2.4A
2R 2(5Ω)
The power dissipated by a resistor is P = I∆V = I 2 R. There are four resistors in the circuit, add the power up,
We have to be careful as we calculate the power provided by the batteries. If current flows from − to +, the
battery provides I∆V power. If current flows from + to −, the battery is charged by the circuit and removes
I∆V power from the circuit. For the current directions drawn on the figure, both batteries would be charged by
positive current, so
∆V1
R4 ∆V2
f e d
Solution
Definitions
Strategy: Use Kirchhoff’s Loop and Junction Equations to produce a sufficient set of independent equations to
find the currents.
Section 1 - Draw the Circuit Draw the circuit to be reduced as given (shown above).
Section 2 - Reduce any Simple Series/Parallel Combinations
Rs = R1 + R2 = 6.0Ω
a Rs b R3 c
∆V1
∆V2
R4
f e d
f e d
I1 + I3 = I2
I1 I2
∆V1
R4 ∆V2
Loop 1 Loop 2
I3
f e d
Section 6 - Write Loop Equations For each loop, write a loop equation. Make sure each circuit element appears
in at least one loopPequation. Furthermore, recall the sign convention for potential differences. Kirchhoff’s Rule
for circuit loops is loop ∆V = 0.
(a) Write Loop Equation 1: The sum of the potential differences for loop 1 is
From f to a, the loop goes through the battery from − to +, so the potential increases ∆Vf a = +∆V1 . From a to
b, the current goes through a resistor in the same direction as the loop causing a potential drop, ∆Vab = −I1 Rs .
From b to e, the loop goes through the resistor in the opposite direction as the loop, causing a potential rise,
∆Vbe = +I3 R4 . From f to e, there is no circuit element. The potential difference across a perfect wire is zero,
∆Vef = 0. The loop equation for Loop 1 is
∆V1 − I1 Rs + I3 R4 = 0
(b) Write Loop Equation 2: The sum of the potential differences for loop 2 is
From e to b, the loop goes through a resistor in the same direction as the current, so the potential decreases(drops)
∆Veb = −I3 R4 . From b to c, the current goes through a resistor in the same direction as the loop, causing a
potential drop, ∆Vab = −I2 R3 . From c to d, the loop goes through a battery from negative to positive causing a
potential rise, ∆Vbe = +∆V2 . From d to e, there is no circuit element. The potential difference across a perfect
wire is zero, ∆Vde = 0.
The loop equation for Loop 2 is
−I3 R4 − I2 R3 + ∆V2 = 0
Section 7 - Solve the Complete Set of Equations Use mathematical techniques to solve for the unknown
currents. We have three unknown currents and three equations, so we’re in business.
(a) System of Equations to be Solved: Solve the three independent equations:
I1 + I3 = I2 (1)
2
I3 = A
5
(f) Solve for I1 : Solve (2) for I1
I3 R4 + ∆V1
I1 =
Rs
( 25 A)(8Ω) + 6V 16V + 30V
= =
6Ω 30Ω
23
I1 = A
15
(g) Solve for I2 : Use (1) to compute I2 :
2 23
I2 = I1 + I3 = A+ A
5 15
29
I2 = A
15
Section 8 - Calculate Potential Differences Across Resistors
(a) Use Ohm’s Law, V = IR, to calculate the potential difference across each resistor.
Using Ohm’s Law:
29
∆VR3 = I2 R3 = V
5
16
∆VR4 = I3 R4 = V
5
46
∆VRs = I1 Rs = V
5
The potential difference across Rs is actually the potential difference across two resistors, R1 and R2 . Since they
are in series, we know that the current is the same through both, I1 . Use Ohm’s Law for these two resistors to
obtain
46
∆VR1 = I1 R1 = V
15
92
∆VR2 = I1 R2 = V
15
Section 9 - Check for Consistency
(a) Check that the junction equation and the loop equations work with the numeric answers. I often make
mistakes in Kirchhoff’s Law problems, checking the loops allows me to find the errors and fix them.
symbolic equation numerical equation checks?
23 2 29
I1 + I3 = I2 15 + 5 = 15 yes
23 2
−I1 Rs + I3 R4 + ∆V1 = 0 − 15 · 6.0 + 5 · 8.0 + 6.0 = 0 yes
−I3 R4 − I2 R3 + ∆V2 = 0 − 52 · 8.0 − 15
29
· 3.0 + 9.0 yes
Capacitive Circuits
Let’s consider two parallel plate capacitors as shown below. There are two ways to make a connection to the
battery; the series and the parallel circuit.
238
22.1. CAPACITOR NETWORKS CHAPTER 22. CAPACITIVE CIRCUITS
+ _
_ +
+ _
+ _
+ _ _ +
+ _
+ _ _ +
+ _
+ _
+ _
+ _
_ +
+ _
+ _
_ +
1 2 3 4
_ +
For two identical parallel capacitors, the two-capacitor system simply increases the plate area by a factor of two
(imagine sliding the capacitors together). For parallel plate capacitors, the capacitance is C = ε0 A/d, so doubling
the plate area doubles the capacitance. In general, for a parallel combination Cparallel = C1 +C2 ; the capacitance
adds.
In the series combination of two identical capacitors, we have halved the potential difference, since half the
potential difference of the battery is established across each capacitor. In general for a series combination,
1/Cseries = 1/C1 + 1/C2 . This expression is derived in the next section.
In the series combination, where does the charge on plates 2 and 3 come from? Charge cannot be moved
from plate 1 to plate 2 or from plate 3 to plate 4 because the airspace between the plates is an insulator. The
negative charge on plate 2 must be taken from plate 3 leaving a positive charge on plate 3.
Q1 + Q2 = QR .
CR = C1 + C2 ,
The potential drop across each capacitor can be found from the ∆V1 QR
definition of capacitance. CR
Q1 Q2 QR ∆V2
∆V1 = ∆V2 = ∆VR = ∆VR
C1 C2 CR C2
Substituting yields Q2
Q1 Q2 QR
+ =
C1 C2 CR
The charge on each plate is the same (in magnitude) due to
conservation of charge
Q1 = Q2 = QR .
Example 22.1 Find the Equivalent Capacitance and Circuit Properties of a Network of
Capacitors
Problem: Three capacitors comprise a network, as shown
at the right. Two capacitors, C1 = 12nF and C2 = 6.0nF, a
are connected in series. These are then connected in parallel
with another capacitor, C3 = 3.0nF, to complete the network
of capacitors. A potential difference of Va − Vb = 12V is
established from a to b. Find the equivalent capacitance, the
potential difference, energy, and charge of each capacitor and C1
sub-network.
C3
C2
Solution
Definitions
a
b
Strategy: Working from the smallest elements, use series and parallel equations to replace them with equivalents
and keep working outward. Then apply the definition of capacitance to each sub-network to compute the potential
difference and charge of each element.
(a) Working from the individual capacitors, reduce parallel and series
combinations. Redraw the circuit using equivalents for simple parallel a
and series combinations. Use the formula for capacitors in series on
C1 and C2 :
1 1 1
= +
Cs C1 C2
C1 + C2 Cs C3
=
C1 C2
C1 C2
Cs =
C1 + C2
12 × 10−9 F 6.0 × 10−9 F
Cs = = 4.0 × 10−9 F
12 × 10−9 F + 6.0 × 10−9 F
b
Cs = 4.0nF
Redraw the circuit.
(b) Examine the redrawn circuit for simple series and parallel combi-
nations and continue reduction. Use parallel capacitor formula on Cs a
and C3
Cp = Cs + C3
Cp = 4.0 × 10−9 F + 3.0 × 10−9 F
Cp = 7.0nF Cp
Don’t get carried away and reduce a part of the circuit which is not a
simple combination. Don’t do more than one step at a time, you will
make mistakes.
(c) Begin with the equivalent circuit, and apply the definition of ca-
pacitance. We know that the potential difference across the equivalent a
capacitor is ∆Vp = ∆V0 = 12V. The charge on the equivalent capac-
itor is
−Qp
Cs C3
(f) Once the voltage and charge on each capacitor in the network is known we can determine the energy stored
in each one. Use the energy stored in a capacitor, U = 12 Q∆V to compute the energy in each capacitor,
1 1
U1 = Q1 ∆V1 = (48 × 10−9 C)(4.0V) = 96nJ
2 2
1 1
U2 = Q2 ∆V2 = (48 × 10−9 C)(8.0V) = 192nJ
2 2
1 1
U3 = Q3 ∆V3 = (36 × 10−9 C)(12V) = 216nJ
2 2
Once the equivalent capacitance is found, it can be used to find the total energy stored in a circuit and the
total charge delivered by the battery. You can work backward to find the charge and potential difference of each
of the individual capacitors.
Solution
Capacitors add in parallel, so the equivalent capacitance of the parallel combination is Cparallel = 12pF + 12pF =
24pF. This parallel combination is in series with the 36pF capacitor. Using the series formula for capacitance,
the equivalent capacitance of the circuit is
1 1 1 5
= + =
Ceq 36pF 24pF 72pF
RC Circuits
245
23.1. REASONING ABOUT RC CIRCUITS CHAPTER 23. RC CIRCUITS
+ _
e− e−
+ _
−
e−
e + _
I I
∆V0
R2 C2
R1
Solution to Part(a)
Immediately after S1 closes, the capacitors offer no resistance to the flow of current and so the potential difference
across the capacitors is zero, therefore ∆VC1 = 0 and ∆VC2 = 0. Because it is in parallel with a capacitor,
∆VR2 = 0. Then by Kirchhoff’s Loop Equation, ∆V0 = ∆VR1 = 10V. The currents in the circuit are found by
Ohm’s Law, I1 = ∆V1 /R1 = 10mA. All this current flows through C2 because at the time switch S1 is closed,
it presents zero resistance while R2 has finite resistance, so IR2 = 0 and IC2 = I1 . This current must also flow
through C1 , so IC1 = I1 .
Solution to Part(b)
After a long time, C1 becomes fully charged and blocks all current; the current in all elements of the circuit
becomes zero. By Ohm’s Law, if zero current is flowing, ∆VR1 = ∆VR2 = 0. This implies ∆VC2 = 0. By
Kirchhoff’s Loop Equation, this means ∆VC1 = ∆V0 = 10V.
Solution
Definitions
R1
∆V = 12V ≡ Potential difference between battery terminals
R1 = 3.0MΩ ≡ Resistance of first resistor
I1
R2 = 6.0MΩ ≡ Resistance of second resistor
C = 3.0µF ≡ Capacitance of capacitor
C I2 R2 I1 ≡ Current through first resistor
∆V
I2 ≡ Current through second resistor
∆V1 ≡ Potential difference across first resistor
∆V2 ≡ Potential difference across second resistor
∆VC ≡ Potential difference across capacitor
Strategy: Use the fact that at long times, capacitors do not pass any current, analyze the resulting DC Circuit,
and use the voltage drops to get the capacitor voltage, and the definition of capacitance to get the charge.
(a) Redraw the circuit leaving connection points for the capaci-
tors, but exclude the capacitors. Since at long times the capacitors
draw no current, they do not contribute to the circuit properties.
The circuit given in the sample problem is simplified in the long- R1
time limit to be without the capacitor, leaving the two resistors
in series with the battery. I1
I2 R2
∆V
(d) Compute Voltage Drop Across Capacitor: Use the network properties (found in the previous section) to
calculate the voltage drop between the points where the capacitor was connected before you redrew the circuit.
This gives the potential difference across the capacitor. The capacitor is in parallel with R2 , so they have the
same potential drop
∆VC = ∆V2 = 8.0V
(e) Compute the Charge: Use the Definition of Capacitance to compute the charge on the capacitor. By
definition of capacitance,
QC = C∆VC = 3.0 × 10−6 F · (8.0V) = 24µC
Definitions
R1
∆V = 12V ≡ Potential difference between battery terminals
R1 = 3.0MΩ ≡ Resistance of first resistor
I1
R2 = 6.0MΩ ≡ Resistance of second resistor
C = 3.0µF ≡ Capacitance of capacitor
C I2 R2 I1 ≡ Current through first resistor
∆V
I2 ≡ Current through second resistor
∆V1 ≡ Potential difference across first resistor
∆V2 ≡ Potential difference across second resistor
∆VC ≡ Potential difference across capacitor
Strategy: Use the fact that at short times, the capacitors do not resist the current. Analyze the resulting DC
circuit.
(a) Redraw the circuit replacing the capacitors with wires. Since
the wire has less resistance than the resistor it is in parallel with,
R2 , the current flows through the wire instead of R2 .
R1
I2 ≈ 0A
I1
∆V2 ≈ 0V
Since the capacitor is in parallel with R2 , they have the same
I2 R2
potential difference across the terminals ∆V
∆VC = ∆V2 ≈ 0V
(b) Solve the DC Circuit: The resulting circuit is a simple DC circuit and the currents can be solved for using
techniques for DC circuits. We can use Ohm’s Law, ∆V = IR, to find the current through R1 . Since R1 is
the only circuit element across the battery, the potential drop across it is the same as that across the battery,
∆V = ∆V1 .
I1 = ∆V1 /R1 = 12V/ 3.0 × 106 Ω = 4.0µA
23.2 RC Circuits
23.2.1 General Behavior of RC Circuits
Both charging and discharging RC circuits have exponential time dependencies. The rate of change of any
t
factor in the circuit is controlled by an exponential factor e− τ where t is the time and τ is the time constant. All
circuit quantities, charge, current, and potential difference change with this time dependence. The number e is
not the electric charge, it is the base of the natural logarithm, ln(e) = 1. For quantities which decrease with time,
the quantity decays to 1/e ≈ 1/2.72 ≈ 1/3 of its initial value in one time constant. Quantities which increase
with time reach 1 − 1e ≈ 23 their final value in one time constant.
Time Constant, τ : The time constant of an RC circuit is τ = RC, where R is the total
resistance the capacitor charges or discharges through and C is the total capacitance.
Base of the Natural Logarithm: The number e is the base of the natural logarithm.
To a few decimal places e = 2.7182818. e is a constant that is programmed into some
TI calculators, or you can use the function e1 = e.
Solution
1 τ
e−1 ≈ = e− τ ,
3
so τ ≈ 10s.
exp(−t/τ )
which at time t = 0 is one and decreases to 0 at long times. The graph below might be
the voltage across a resistor in a discharging RC circuit with initial voltage, ∆V0 = 10V,
and time constant τ = 10s. τ is the time required to the initial voltage ∆V0 to decay
to ∆V0 /e = ∆V0 /2.72 = 0.37∆V0 .
∆V0 10
6
∆VR (t) (volts)
1
τ
0
0 5 10 15 20 25 30
t (sec)
Increasing Exponential Curve: Some features of RC circuits increase with time. They
have a general time dependence of
1 − exp(−t/τ ),
which at time t = 0 is zero and increases to 1 at long periods of time. The graph
below might be the voltage across a charging capacitor with final voltage, ∆Vf = 10V,
and time constant τ = 10s. The time τ is the time required for the voltage to reach
1 − 1e = 0.63 of the final voltage.
10 ∆Vf
7
∆Vc (t) (volts)
2
τ
1
0
0 5 10 15 20 25 30
t (sec)
A D
The potential differences ∆VBC = ∆VDA = 0 because the potential difference across a perfect wire is zero. The
potential difference ∆VAB = +IR because the potential drops as current crosses a resistor. The potential across
the capacitor is ∆VCD = −Q/C because potential decreases in the direction of the electric field. Substituting
into the loop equation gives,
Q
IR − =0
C
For the direction of I chosen, the current is related to the charge on the positive capacitor plate by
dQ
I=−
dt
because a decrease in the charge on the plate causes a positive current.
dQ Q
−R − =0
dt C
Rearranging gives,
dQ 1
=− dt
Q RC
Integrate from time 0 when the capacitor has charge Q(0) to time t,
Z Q(t) Z t
dQ 1
=− dt
Q(0) Q RC 0
−t
ln(Q(t)) − ln(Q(0)) =
RC
Use a property of logarithms, ln(B) − ln(A) = ln(B/A),
−t
ln(Q(t)/Q(0)) =
RC
Exponentiate both sides. We will write the exponential both as exp(x) and ex .
exp ln(Q(t)/Q(0)) = exp(−t/RC)
Use the property of exponentials that exp(ln(x)) = x, to yield the final result
t
Q(t) = Q(0)e− τ
where τ = RC. All other circuit properties are related to Q(t) as shown in the table below:
Solution
Definitions
R1 ≡ Resistance of Resistor 1
R
∆V0 ≡ Initial Potential Difference Across C
C ≡ Capacitance of the capacitor
I
I(t) ≡ Current in the circuit
C Q(t) ≡ Charge on the Capacitor
I0 ≡ Initial Current
Q0 ≡ Initial Charge on Capacitor
(c) Use Discharging Form for Charge: Since the charge on the capacitor starts at its highest value Q0 and
decays toward zero, the correct form of the time dependence is a decaying exponential,
Q(t) = Q0 exp(−t/τ )
(d) Compute Initial Current, I0 , Through Resistor: Initially, the voltage across the capacitor is VC , so the
initial current is by Ohm’s Law
∆VC 2
I0 = = 6.0V/9.0MΩ = × 10−6 A
R1 3
(e) Use Discharging Form of Current: Since the current in the circuit starts at its highest value I0 and decays
toward zero, the correct form of the time dependence is a decaying exponential,
I(t) = I0 exp(−t/τ )
2
× 10−6 A exp −t/ 27 × 10−3 s
I(t) =
3
100
90
80
70
60
V C (t) (Volts)
50
40
30
20
10
0 2 4 6 8 10 12 14 16 18 20
t (sec)
Solution to Part(a)
The potential difference across the capacitor decays as a function of time, so the time dependence is
t
∆VC (t) = ∆VC0 e− τ
where ∆VC0 = 100V which can be read directly from the graph. The time constant τ = RC is the time for the
voltage to decay to e−1 = 0.37 of its initial value, so it is the time when the voltage crosses (0.37)(100V) = 37V,
which can be read from the graph as τ = 3s.
Solution to Part(b)
Solution to Part(c)
1
The time for the voltage to fall to 4 of its initial value is found by solving the decay equation,
∆VC t
0.25 = = e− τ
∆VC0
t
ln(0.25) = −
τ
t = −τ ln(0.25) = −(3s)ln(0.25) = 4.2s
In the table, Q(∞) is the charge on the capacitor at long times and ∆VC (∞) is the potential difference across
the capacitor at long times.
Solution
Definitions
R1 ≡ Resistance of Resistor 1
R1
∆V1 ≡ Potential Difference Across Battery
I C ≡ Capacitance of the capacitor
I(t) ≡ Current in the circuit
∆V1 Q(t) ≡ Charge on the Capacitor
C
I0 ≡ Initial Current
Qf ≡ Final Charge on Capacitor = Q(∞)
(c) Use Charging Form for Charge on Capacitor: Since the charge on the capacitor starts at zero and
increases toward Qf , an increasing exponential time dependence is appropriate,
Q(t) = Qf (1 − exp(−t/τ ))
(d) Compute Initial Current, I0 , Through Resistor: Use Ohm’s Law to calculate initial current, the current
is I0 = ∆V1 /R1
2
I0 = 6.0V/9.0MΩ = × 10−6 A
3
(e) Use Charging Form of Current: The current has its maximum value at t = 0 and then decays toward
zero, so the proper form of the time dependence is
I(t) = I0 exp(−t/τ )
2
× 10 A exp −t/ 27 × 10−3 s
−6
I(t) =
3
Solution to Part(a)
Solution to Part(b)
For a charging capacitor, the potential difference starts at zero and charges to the applied voltage,
t
∆VC (t) = ∆Vf (1 − e− τ )
Solution to Part(c)
Determine Time Dependence from Graph: If the plot increases toward a maximum
value for a long period of time, then the time dependence is 1 − exp(−t/τ ). If the plot
begins at its highest value and decays toward zero, then the time dependence has the
functional form exp(−t/τ ).
Compute Time Constant from Point: The time constant can be approximated by
taking any point off the graph and using this point to solve for τ , as shown in Example
23.9 Use a Potential versus Time Plot to Determine RC Circuit Properties.
Example 23.9 Use a Potential versus Time Plot to Determine RC Circuit Properties
Problem: A capacitor is charged to some value, then connected in series with a 100Ω resistor and allowed to
discharge. A measurement of the potential difference across the capacitor is shown in the figure below. Compute
the time constant of the circuit, the capacitance of the capacitor, and the initial charge stored on the capacitor.
12
10
6
V
0
0 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007
t(sec)
Solution
Strategy: Determine the time constant from the plot, then use the definition of the time constant and the
definition of capacitance.
(a) Determine Form of Time Dependence: Examine the graph and determine the functional form of the
time dependence, whether it is exp(−t/τ ) or 1 − exp(−t/τ ). Read the overall constant off the curve. Since the
curve begins at its maximum value and decays toward zero, the time dependence has the form
t
∆VC = ∆V0 exp − .
τ
(d) Compute the Initial Charge Stored: Use the definition of capacitance, C = Q/∆V , to determine the
initial charge stored on the capacitor. The initial charge is given by Q0 = C∆V0
Solution to Part(a)
Solution to Part(b)
12V
5000sec time(sec)
V V
t t
The voltage step in the circuit with capacitance is smoothed out. If we return to the plumbing analogy
developed in lab, a capacitor acts like a pressure tank. When pressure rises in a system the capacitor takes up
261
24.2. REVISITING SHIELDING BY CONDUCTORS CHAPTER 24. FINAL TOPICS TEST 2
some of the excess. When pressure falls, the capacitor acts as a reservoir to maintain the pressure.
conductor
+
+
_
_
_ + _
+ + No Field +
_
_
We can reverse the situation for the conducting shell and place the charge inside as shown in the example
below.
Conductor
Solution to Part(a)
The drawing should have the following features: The field lines
should leave λ symmetrically and radially. The lines should be
evenly spaced about λ. The lines should curve and hit the inner
surface of the conductor perpendicular to the surface. An equal
number of lines should leave the outer surface of the conductor + _ +
radially. + charge should be drawn where lines begin and − charge
should be drawn where lines end. _
+Q
b a _
c d
+ _ +
Conductor
Solution to Part(b)
Finally, we can consider grounding the conductor. The surface charge densities not held in place by the charged
objects will escape to ground, completely isolating one region from the charge in the other region.
conductor No Field
conductor
_
_ No Field
_
+ _ +Q
_
_
Biot-Savart Law
Electric charges produce electric fields. Moving electric charges produce magnetic fields. This might come
as something of a surprise, since as a kid you probably discovered (without really saying it) that “magnets make
magnetic fields.” Why don’t we start from this point, using magnets as our basic building blocks? We might
hope for a nice analogy between electric charge as the source of electric field and something else as the source
of magnetic field, but it turns out that there is no such thing as magnetic “charge”. Magnets can be explained
in terms of the motion of the electric charges they contain. So we begin with the more basic concept of moving
charges. We will come back to magnets later.
Magnetic Flux Through any Closed Surface (Maxwell II): The mathematical ex-
pression of non-existence of magnetic charge is that the magnetic flux, φm , through
any closed surface is always zero.
Z
φm = (B ~ · n̂)dA = 0
S
where B ~ is the magnetic field on the surface. Note the similarity to Gauss’ Law, where
the flux from the electric field is equal to Qenclosed /ε0 , but for magnetic fields, the
magnetic charge enclosed is always zero. This is the second Maxwell equation, Gauss’
law is the first.
The fact that there is no magnetic charge is a restriction on the kind of magnetic field map we can draw.
Solution
264
25.2. BIOT-SAVART LAW CHAPTER 25. BIOT-SAVART LAW
This cannot be a valid field, since it produces a net flux out of a parallel-piped with sides ⊥ to the x-axis.
Solution
Since net flux exits the region, it must be an electric field map. The net flux exiting any region for a magnetic
field is zero.
Maxwell’s Equations
~ · n̂)dA = Qenclosed
R
Gauss’ Law: S
(E ǫ0
~ · n̂)dA = 0
R
No Magnetic Monopoles: (B S
Ampere’s Law:
Lorentz Force
F~ = q E
~ when there are no magnetic fields
One Maxwell’s equation is simply missing and is the integral expression of the Biot-Savart law of next section.
A number of the equations have restrictions that we will remove over the next few weeks.
We will rarely use this definition to compute the cross product. It is usually easier to compute the magnitude of
the cross product, then use the right hand rule to compute the direction.
The above means that the cross-product of two parallel vectors or a vector with itself is zero. The magnitude
~ = |A
of the cross-product of two perpendicular vectors is |C| ~ × B|
~ = |A||
~ B|.
~
Avoid these pitfalls: Make sure you are using your right hand in all these manipulations! (We laugh ourselves
silly when we see students’ left hands in motion during a test.) Also make sure you do things in the right order;
A~ ×B ~ is not the same as B~ × A–they
~ differ by a minus sign–so you must align your fingers with the first vector
in the cross product, then curl them into the second vector in the cross product, not the other way around.
Solution
~ =A
C ~×B ~ = (Ay Bz − Az By )x̂ + (Az Bx − Ax Bz )ŷ + (Ax By − Ay Bx )ẑ
~ = (2 − 6)x̂ + (15 − 1)ŷ + (2 − 10)ẑ = −4x̂ + 14ŷ − 8ẑ
C
Because of the form of the physical laws governing magnetic fields, we have to work in three dimensions.
When a three-dimensional coordinate system is drawn, you must make choices about the direction of the axes.
This choice must be made by some criteria or all the signs of the numbers we compute are arbitrary.
Right and Left Handed Coordinate Systems: When we draw a three-dimensional
coordinate system on a flat piece of paper, we can choose the positive directions for
two of the axes in any direction we want. The third axis must be chosen so x̂ × ŷ = ẑ.
Such a coordinate system is called a right-handed coordinate system and will give us
correct signs. If we choose the third axis so that x̂ × ŷ = −ẑ, we have drawn a left
handed coordinate system and all the quantities we compute will be incorrect.
z
+z out-of page
x +z into page x
z
When I do a cross-product, I draw a right-handed coordinate system and use it to work out the results of
crossing various combinations of unit vector.
Example 25.5 Dr. Stewart’s Cross-Product Method
Problem: The vector A ~ = (2, 0, 3) = 2x̂ + 3ẑ is crossed with the vector B
~ = (4, 3, 0) = 4x̂ + 3ŷ. Calculate the
~ ~ ~
cross-product C = A × B.
Solution
z out-of page
x
~ =A
C ~×B
~ = (2x̂ + 3ẑ) × (4x̂ + 3ŷ)
Multiply out the cross-product without changing the order of the terms.
~ = (8x̂ × x̂ + 9ẑ × ŷ + 12ẑ × x̂ + 6x̂ × ŷ)
C
Since the angle between a vector and itself is zero, a vector crossed with itself is zero, so x̂ × x̂ = 0. To do the
other cross-products, use the right hand rule on the coordinate system drawn above, x̂ × ŷ = ẑ, ẑ × ŷ = −x̂, and
ẑ × x̂ = ŷ.
C~ = −9x̂ + 12ŷ + 6ẑ
Units of the Magnetic Field: The units of the magnetic field are the Tesla(T).
Ns
1T = 1
Cm
The unit Gauss (G) is also used for magnetic fields. It must be converted to Tesla
before any calculations are performed.
1G = 1 × 10−4 T
Tm
µ0 = 4π × 10−7
A
The strength of the magnetic field is set by µ0 .
~ 0 = µ0 q~v × r̂10
B
4π r10 2
where ~r10 is the vector from the location of the charge to the location where the field is
calculated, and µ0 = 4π × 10−7 Tm A . This formula is an approximation valid for particle
speeds that are small compared to the speed of light.
Solution
Definitions
(b) Use the Magnetic Field of a Moving Particle: Let P be the point where the field is computed. The
Biot-Savart Law gives the magnetic field of a moving particle as
~ 0P = µ0 q~v × r̂0P
B
4π 2
r0P
~P | = µ0 |q||~v |
|B 2
4πr0P
(d) Compute Direction of Magnetic Field: First compute the direction of ~v × r̂0P using the Right Hand
Rule. Pointing the fingers of the right hand in the ~v direction and curling the fingers in the r̂0P direction gives
~ P is into the page in the −ẑ direction.
the direction of ~v × r̂0P out of the page. Since q < 0, B
(e) Substitute and Compute:
~P | = (4π × 10−7 Tm
A )(1.609 × 10
−19
C)(0.001)(3 × 108 ms )
|B 2
= 4.827 × 10−19 T
4π(0.1m)
Combine this with the direction of the field to write the magnetic field as a vector
~ P = −4.827 × 10−19 Tẑ
B
In physics, we routinely deal with charge moving through space, but in everyday life, most of the magnetic
fields we work with are generated by currents flowing in wires. The reformulated expression for the magnetic field
due to a current is called the Biot-Savart law. There is a subtlety though, a current is continuous (Kirchhoff’s
Junction Law), so it really does not make sense to talk of an isolated current at a point. We therefore express the
Biot-Savart law in terms of the infinitesimal magnetic field dB ~ generated by an infinitesimal element of current
Id~ℓ where the current, I, flows in the direction ~ℓ and the field is generated by a segment of length dℓ of the
current.
~ 0 produced by the current element Id~ℓ is
Biot-Savart Law: The magnetic field dB
~0 = µ0 Id~ℓ × r̂10
dB 2
4π r10
where ~r10 is the vector from the current element to the location where the field is com-
puted, and µ0 = 4π × 10−7 Tm A . This is the fundamental starting point for computing
the magnetic field of any current distribution. We imagine the distribution as a collec-
tion of current elements, use the Biot-Savart law to find the field contributed by each
current element, then sum all the contributions to get the total net field.
~ 0|
Biot-Savart Law - Magnitude Only: The magnitude of the magnetic field |dB
produced by the current element Id~ℓ is
where ~r10 is the vector from the current element to the location where the field is
computed and θ is the angle between d~ℓ and ~r10 and the angle is measured from d~ℓ to
~r10 .
Representing Magnitude of Field Into or Out of Paper: There are two ways to
represent the strength of the field into or out of the paper. The magnetic field is stronger
where arrows are closer, so a stronger field can be represented by arrows spaced closer
together. We will also sometimes ask for the magnetic field at isolated points, in this
case the spacing of the arrow heads/tails does not mean anything. Use larger or smaller
circles to represent relative magnitudes in this case.
We can use this notation to show the z axis direction in a right-handed coordinate
system. The figure to the right shows a right-handed coordinate system. y
z x
The figure below shows the magnetic field of an infinite straight wire viewed from the end and viewed from
the side.
I
I
Solution
(b) Compute the Magnitude: The magnitude of the magnetic field is given by the formula for the magnetic
field of an infinite wire,
µ0 I
|B| =
2πR
where R is the distance from the wire.
(4π × 10−7 Tm
A )(3A)
|B| = = 6 × 10−6 T
2π(0.1m)
Consider the segment of wire from A to B shown below. The vector ∆~ℓAB points in the same direction as
the wire and has length the distance between points A and B. As such the vector ∆~ℓAB = ~rAB the displacement
vector from A to B. So if we have a well-drawn figure, we can just read ∆~ℓ off the figure. For example, if the
large blocks on the figure below represent 1cm then ∆~ℓAB = (2cm, −2cm, 0).
∆ AB
B
y
~ 0 produced by
Magnetic Field of a Finite Current Element: The magnetic field B
~
the finite current element I∆l can be approximated by
~ ~
~ 0 = µ0 I∆l × r̂10 = µ0 I∆l × ~r10
B
4π 2
r10 4π 3
r10
where ~r10 is the vector from the center of the current element to the location where
the field is computed. The vector ∆~l points in the direction of the current and has
magnitude ℓ, the length of the current element. The approximation improves as the
distance to the field point becomes large compared length of the current element. The
second expression uses the definition of the unit vector, r̂10 = ~r10 /r10 .
The field of the finite current element is simply the Biot-Savart law before the limit ∆~ℓ ⇒ d~ℓ is taken.
Solution
(a) Compute the Magnitude: The displacement vector from the origin to the point (5cm, 0, 5cm) makes an
angle of 45◦ to the current. The displacement vector from the origin to the point P (see figure below) has length
p √
r0P = (5cm)2 + 0 + (5cm)2 = 5 2cm
~
~ P = µ0 I∆ℓ × r̂0P
B
4π 2
r0P
The current element given is 1mm in length and flows in the +ẑ direction.
Substitute
−4
~ P = µ0 (5 × 10 Amẑ) × r̂0P
B
4π 2
r0P
The magnitude of the cross product is
1
|ẑ × r̂0P | = |ẑ||r̂0P | sin(45◦ ) = sin(45◦ ) = √
2
where 45◦ is the angle between ẑ and r̂0P . Substitute and compute the magnitude
~ P | = 7.07 × 10−9 T
|B
y into page
Let’s rework the above example by letting the vectors do the work for us.
Solution
∆ x
y into page
(b) Compute the Current Element: The path element given is ∆ℓ = 1mm in length and flows in the +ẑ
direction, I∆~ℓ = I∆ℓẑ
(c) Compute the Field: The finite current element approximation to the Biot-Savart law is
~ ~
~ P = µ0 I∆ℓ × r̂0P = µ0 I∆ℓ × ~r0P
B
4π 2
r0P 4π 3
r0P
Substitute
~ P = µ0 (I∆ℓẑ) × ~r0P = µ0 ((0.5A)(0.001m)ẑ)
B √
× (0.05mx̂ + 0.05mẑ)
4π 3
r0P 4π (5 2 × 10−2 m)3
~ P = 7.07 × 10−9 T(ẑ × (x̂ + ẑ)) = 7.07 × 10−9 T(ẑ × x̂ + ẑ × ẑ)
B
Therefore,
~ P = 7.07 × 10−9 Tŷ
B
since ẑ × ẑ and ẑ × x̂ = ŷ which I found using the right-hand-rule on my diagram.
Field of an Infinite Solenoid: The magnetic field inside an infinitely long solenoid
wound with n of wire turns per unit length carrying current I is
~ = µ0 nI
|B|
The field is uniform and directed along the axis of the solenoid in the following right-
handed sense: if you curl the fingers of your right hand to follow the current in the
solenoid windings, your right thumb points in the direction of the field inside. Outside
of the infinite solenoid, the field is zero. Avoid a notorious trap— n is not the number
of turns of wire wrapped around the solenoid! It’s a number of turns divided by the
length over which they are wrapped along the solenoid cylinder.
N
n=
L
where N is the number of wraps and L is the length.
It is a pain drawing all those curves and it is also difficult to determine direction of the field. I will represent
a solenoid by the cut-away drawing shown below where the current direction is clearly shown. If you don’t want
to use the right-hand rule given above (and I don’t), the direction of the field can be found by using the right-
hand-rule for a wire on any of the loops. The field is uniform inside and zero outside. I have drawn the field of
one of the individual loops as a dashed line. The direction of the field of one of the loops, inside the solenoid, is
the same as the direction of the solenoid’s field.
I out-of page
I into page
Solution to Part(a)
The magnitude of the magnetic field of an infinite straight solenoid is
N
B = nµ0 I = µ0 I
L
where n is the turns per unit length, N is the total number of turns of wire, and L is the length of the solenoid.
If the solenoid carries 0.5A and has length L = 30cm and is wound with a total of N = 500turns of wire, then
500turns Tm
B= (4π × 10−7 )(0.5A) = 1.05 × 10−3 T
30cm A
Solution to Part(b)
Solve the expression for B for the current and substitute 1T for the magnetic field,
BL (1T)(0.30m)
I= = = 477A
N µ0 (500turns)(4π × 10−7 Tm
A )
Magnetic Field Lines Are Closed Curves: Since there is no magnetic charge, mag-
netic field lines are closed curves. Recall that field lines begin and end on charge and
there is no magnetic charge.
The best way to learn to draw a magnetic field map is from an example.
Example 25.11 Draw the Magnetic Field Lines for Two Parallel Currents
Problem: Two wires are parallel and carry currents of equal magnitude in the same direction. Draw the field
map for this situation.
Solution
I I
I I
I I
Example 25.12 Draw the Magnetic Field Lines for Two Anti-Parallel Currents.
Problem: Two wires are parallel and carry currents of equal magnitude in opposite directions. Draw the field
map for this situation.
Solution
I I
I I
Example 25.13 Draw the Magnetic Field Lines Inside, Outside, and Near the Ends of
a Solenoid.
Problem: A number of loops are tightly wound to create a long solenoid. Draw the magnetic field lines inside,
outside, and near the ends of the solenoid. Use one line per wire.
Solution
(b) Sum the Magnetic Fields Far Inside and Outside: Using our
technique for infinite wires, bend loops of opposite orientation apart
and combine loops of the same orientation.
(c) Reason About the Strength of the Field: The field above and
below the solenoid partially cancels. It is weaker than the field inside,
so push the lines outside the solenoid farther from the solenoid. We
would have seen the cancellation if we had included more lines.
I− I+ r+ x
B+
y into page
283
26.1. ADDING MAGNETIC FIELDS CHAPTER 26. COMPUTING MAGNETIC FIELDS
Strategy: Compute the magnetic field of each wire; then, add, using linear superposition.
(a) Select Right Handed Coordinate System: For x̂ × ŷ = ẑ, +ŷ must be into the page, therefore the
currents are as drawn.
(b) Use the Right Hand Rule for a Wire to Compute the Fields: The field lines of an infinite wire form
circles around the wire. Use the right hand rule for a wire for I+ which flows into the page. Imagine grabbing
the wire with your right hand, with your thumb pointing in the direction of the current. Your fingers curl in the
direction of the magnetic field, which for I+ gives a clockwise field. This field points down at the point P where
we wish to compute the magnetic field, so the magnetic field from I+ points in the −ẑ direction. Repeating the
reasoning for the I− current, gives the direction of the magnetic field of the wire carrying current out of the page
as +ẑ as drawn. B+ is slightly larger than B− , because the I− current is farther away.
(c) Compute B ~ + : The distance from I+ to P is 2cm − 0.25cm = 1.75cm. The magnetic field of an infinite
straight wire is
µ0 I (4π × 10−7 TmA )(20A)
B+ = = = 2.29 × 10−4 T
2πr+ 2π(0.0175m)
~ + = −2.29 × 10−4 Tẑ
B
~ − : The distance from I− to P is 2cm + 0.25cm = 2.25cm. The magnetic field of an infinite straight
Compute B
wire is
µ0 I (4π × 10−7 Tm
A )(20A)
B− = = = 1.78 × 10−4 T
2πr− 2π(0.0225m)
~ − = 1.78 × 10−4 Tẑ
B
Compute the Total Field: By linear superposition, the total field due to the two wires at point P is
~P = B
B ~+ + B
~ − = −2.29 × 10−4 Tẑ + 1.78 × 10−4 Tẑ = −0.51 × 10−4 Tẑ
(a)On the side view diagram, draw the magnetic field due to the solenoid only.
(b)At the point P in the top view diagram, draw the magnetic field vector from the solenoid, from the
straight wire, and the total magnetic field vector.
(c)Compute the magnetic field of the solenoid.
(d)Compute the magnetic field of the straight wire at point P .
(e)Compute the total magnetic field at point P .
I
P P
d
y
d x
Use the Right Hand Rule for a Wire on any of the solenoid wires to get the direction of the uniform field in the
solenoid.
Bwire BP
Bwire I
P P Bsole
d
y
d x
Use the Right Hand Rule for a Wire to get the field of the wire at point P. The field lines are circles about the
current.
Solution to Part (c)
= 4 × 10−5 Tŷ.
~P = B
B ~ wire + B
~ sole
~P = B
B ~ 1P + B
~ 2P + B
~ 3P + B
~ 4P + B
~ 5P + B
~ 6P .
What physical principle allows you to add the fields in this manner?
~ 1P , from the segment 1, between
(b)What is the contribution to the total magnetic field at point P , B
points A → B?
Because of the symmetry of the problem, the total field is
~ P = 2(B
B ~ 1P + B
~ 2P + B
~ 3P )
8cm
6cm
A 1 B P 6
y
F G
2
4cm
5
C 3 D 4 E
2cm
Solution to Part(a)
The magnetic fields of the individual current elements add because of the principle of linear superposition.
Solution to Part(b)
For segment A → B, the current element I∆~ℓ1 is parallel to r̂1P and therefore the field is zero.
Solution to Part(c)
The vectors ∆~ℓ2 , ∆~ℓ3 , ~r2P , and ~r3P are drawn on the figure below and can be read directly from the figure.
∆~ℓ2 = (2cm, −2cm, 0)
8cm
6cm
A 1 B P 6
y r2P F G
2
4cm
r3P 5
∆ 2 ∆ 3
C 3 D 4 E
2cm
Solution to Part(d)
∆~ℓ3 = (2cm, 0, 0)
Solution to Part(g)
The vector from the center of segment 3 to the point P is from the diagram ~r3P = 1cmx̂ + 2cmŷ.
Solution to Part(h)
The magnetic field from segment 3 is given by the finite element approximation to the Biot-Savart law,
~
~ 3P = µ0 I ∆ℓ3 × r̂3P
B
4π 2
r3P
p
therefore we need r̂3P and the length r3P . The length of the vector is given by r3P =
√ (1cm)2 + (2cm)2 =
5cm. The unit vector is by definition
~r3P 1 2
r̂3P = = √ , √ ,0
r3P 5 5
The expression for the magnetic field requires the calculation of the cross product,
1 2 4cm 4cm
∆~ℓ3 × r̂3P = (2cmx̂) × ( √ x̂ + √ ŷ) = √ (x̂ × ŷ) = √ ẑ
5 5 5 5
Now finish the calculation,
~ −7 Tm
~ 3P = µ0 I ∆ℓ3 × r̂3P = 4π × 10
B A
(2A) √
1 0.04m
√ ẑ
4π 2
r3P 4π ( 5 × 10−2 m)2 5
~ 2P = 7.15 × 10−6 Tẑ
B
Solution to Part(i)
So by linear superposition,
~ P = 2(0 + 5.06 × 10−6 Tẑ + 7.15 × 10−6 Tẑ) = 24.42 × 10−6 Tẑ
B
y
B
x
P
This means rather than working with the full vector expression for the Biot-Savart law, we can work with the
magnitude alone, since we already know the direction,
µ0 Idℓ sin θ
dB0 = 2
4π r10
where θ is measured from d~ℓ to ~r. We do have to be careful of the sign, but this is a lot of progress.
Solution
Strategy: Imagine dividing the wire into small segments (i.e., current elements), use the Biot-Savart Law on
each segment, and then use linear superposition to add up the field contributions of all the segments.
(a) Divide the Wire into Small Segments: Divide the wire
into small current elements I∆~ℓ, where d~ℓ points along the wire
y
in the direction of current flow.
I∆
ri0
B0 x
+z Out of Page
(b) Use Linear Superposition: The total magnetic field can be found by adding the fields of the current
~ i,
elements, B
~0 = ~ i.
X
B B
i
(c) Use Biot-Savart Law: The field produced by the individual segment i is in the +ẑ direction by the
right-hand-rule. Its magnitude is given by the Biot-Savart Law
µ0 I∆ℓ sin θ
Bi = 2
4π ri0
where ~ri0 is the vector from the current element to the origin and θ is the angle between ∆~ℓ and ~ri0 . The angle
between these vectors is always θ = 90◦ , so sin θ = 1. The distance from the element i to the origin is ri0 = R,
the radius of the circle. Since the field of each segment points in the +ẑ direction,
~ i = µ0 I∆ℓ ẑ
B
4π R2
The total field at the origin is X
~0 = µ0 I∆ℓ
B ẑ
i
4π R2
(d) Write Sum as Integral: Let ∆ℓ become infinitely short.
Z
µ0 Idℓ µ0 I
Z
B~0 = ẑ = dℓ ẑ
2 4π R2 C
C 4π R
Magnetic Field of Current Loop at the Origin: The magnitude of the magnetic field
at the origin of a flat circle of wire of radius R that carries current I is
µ0 I
B0 =
2 R
The magnetic field of a finite current element is an approximation to the exact expression for the field of a
finite wire. It is accurate in the limit where the distance from the wire is large compared to the length of the
segment, which will be the case any time the wire is broken in many segments.
Solution
riP
θi
i
∆x x
d
(b) Compute Needed Vectors: The displacement vector from the chunk i to the point P is
(c) Compute the Angle: The angle between the current and the displacement vector for the ith segment, θi ,
can be expressed as
R
sin θi = p
(d − xi )2 + R2
I found this by using sine = opposite/hypotenus on the dashed triangle.
(d) Compute the Field: The total field is the sum of the fields of each segment.
X
BP = BiP
i
The field of each segment is given by the Biot-Savart law restricted to the plane.
X µ0 I∆x sin θi
BP = 2
i
4π riP
X µ0 I∆x R X µ0 I∆xR
= 2 2
p = 3
i
4π (d − xi ) + R (d − xi )2 + R2
i
4π ((d − xi )2 + R2 ) 2
(e) Convert the Sum to an Integral and Hope for the Best:
Z L
µ0 IRdx µ0 IR L dx
Z
BP = 3 =
−L 4π ((d − x) + R ) 2
2 2 4π −L ((d − x)2 + R2 ) 23
Solution
Definitions
Strategy: Use the formula for the magnetic field of a short wire.
(a) Use Formula for the Magnetic Field of a Short Wire: The magnetic field of a short wire is given by
µ0 I
B= sin θ1 + sin θ2
4πR
where θ1 is the angle shown in the figure, θ2 = 0 because of the geometry of the problem, and R is the
perpendicular distance to the wire.
(b) Compute sin θ1 : The sin = ho is opposite over hypotenuse. The length of the hypotenuse is d =
√ p
R + L = (5cm) + (10cm)2 = 11.2cm. The length of the opposite side is L = 10cm, therefore
2 2 2
10cm
sin θ1 = = 0.893
11.2cm
(c) Substitute and Compute: Substitute the value for sin θ1 into the expression for the magnetic field,
µ0 I (4π × 10−7 Tm
A )(5A)
B= sin θ1 = (0.893)
4πR 4π(5cm)
B = 8.9µT
A (a, b) (0,0) x
B P
C
I
(c,d)
D
The displacement vector and current element are parallel for segment AB, so the magnetic field at point P is
zero.
The vector ∆ℓBC points from the point (a, b) to the point (c, d). This vector is
The displacement vector from the center of the segment to the point P is
A (a, b) (0,0) x
B P
rBCP
∆ BC
C
I
(c,d)
D
Divide the segment BC into N small pieces. Let the center of the ith piece be (xi , yi ). The line from B to C
has the function, yi = d − 2xi . The displacement vector from the segment i to the point P is
If we divide the segment BC into N pieces, then vector ∆~ℓ for the segment i is ∆~ℓ = ∆~ℓBC /N . The distance
along the x axis from B to C is −a, so N = −a/∆x, and the
∆~ℓBC ∆x∆~ℓBC
∆~ℓ = = = ∆x(1, −2)
N −a
The total field at P is the sum of the fields of the segments by linear superposition
~ BCP = ~ iP
X
B B
i
Convert to an integral
0
~ BCP = − dµ0 I ẑ dx
Z
B p
4π a ( x2 + (−d + 2x)2 )3
Integrals.com
dx −2d + 5x
Z
p = √
2
( x + (−d + 2x) ) 2 3 d d2 − 4dx + 5x2
2
0
~ dµ0 I ẑ −2d + 5x
BBCP = − √
4π d2 d2 − 4dx + 5x2 a
dµ0 I ẑ −2 −2d + 5a µ0 I ẑ −2d + 5a
=− − √ = 2 + √
4π d2 d2 d2 − 4da + 5a2 4πd d2 − 4da + 5a2
Work out the messy bit, a = −1cm and d = −2cm
~ BCP = µ0 I ẑ 2 + (−1) = µ0 I ẑ
B
4πd 4πd
−7 Tm
~ BCP = µ0 I ẑ = (4π × 10
B A )(0.3A)ẑ
= 1.5 × 10−6 Tẑ
4πd 4π(0.02m)
A substantial correction.
The distance from the segment CD to the point P is 2cm. The field points out of the page at point P by the
right-hand rule for a wire. The field of an infinite straight wire,
−7 Tm
~ CDP = µ0 I ẑ = (4π × 10
B A )(0.3A)
ẑ = 3 × 10−6 Tẑ
2πr 2π(0.02m)
Divide the wire into small segments. An element of the segment i is ∆~ℓ = ∆xx̂. The location of the ith segment
is (xi , d) and the displacement vector from the ith segment to the point P is
~ CDP = ~ iP
X
B B
i
Convert to an integral
∞
~ CDP = − µ0 Idẑ dx
Z
B p
4π 0 ( (x + d2 ))3
2
Integrals.com
dx x
Z
p = √
2
( (x + d ))2 3 d x2 + d2
2
∞
~ CDP µ0 Idẑ x µ0 Idẑ µ0 I ẑ
B =− √ =− =−
4π d2 x2 + d2 0 4πd2 4πd
Exactly half the infinite wire result, which was what we were expecting.
~ CDP = 1.5 × 10−6 Tẑ
B
~ = Bx x̂ = µ0 2πR2 I
B x̂
4π (x2 + R2 )3/2
R
B
P x
r
z
Solution
The magnetic field of a loop of current of radius R = 5cm, a distance z = 2cm along the axis of the loop is given
by
µ0 2πR2 I (4π × 10−7 Tm
A ) 2π(5cm)2 (10A)
B= =
4π (z 2 + R2 ) 23 4π 3
((2cm)2 + (5cm)2 ) 2
B = 1 × 10−4 T
By the right hand rule, the magnetic field at (0, 0, 2cm) is in the +ẑ direction, therefore
~ = 1 × 10−4 Tẑ
B
Field of a Finite Solenoid: The magnetic field at a point a distance a from one end
and b from the other end along the central axis of a finite solenoid with n turns per
unit length carrying current I and having radius R is
~ = 1 µ0 nI √ b
|B| +√
a
.
2 b2 + R 2 a2 + R2
Avoid the trap – n is not the number of turns of wire wrapped around the solenoid!
It’s the number of turns divided by the length that they are wrapped along the solenoid
cylinder.
R P
x
a b
Ampere’s Law
Ampere’s Law: The current, IC through the surface bounded by a closed curve, C, is
related to the line integral of the magnetic field B along the curve by
I
~ · d~l = µ0 IC ,
B
where IC is positive if the thumb of the right hand is pointed in the direction of the
current and the fingers curl in the direction of d~l.
299
27.1. AMPERE’S LAW CHAPTER 27. AMPERE’S LAW
Amperian paths
I is positive I is negative
I I
Direction of traversal
of Amperian path
Maxwell’s Equations Part II: Maxwell’s Equations and the Lorentz force as introduced
to this point are:
Maxwell’s Equations
~ · n̂)dA = Qenclosed
R
Gauss’ Law S
(E ǫ0
~ · n̂)dA = 0
R
No Magnetic Monopoles (B S
Lorentz Force
F~ = q E
~ No Magnetic Fields
Current
Magnetic Field
Amperian Path
Solution
There is current flowing through the gray surface which is bounded by the Amperian path. Since the field points
generally in the direction d~ℓ, the integral I
B~ · d~ℓ = µ0 I 6= 0
C
so a current flows through the surface bounded by the curve.
Solution
(a) Draw the Conductors: Draw the conductors, label the regions.
IV
III
II
I
(b) Draw Currents: Draw the currents using ⊗ for current into the
page and ⊙ for current out of the page. Select a number of ⊙ and
⊗ proportional to the current. Select the current in the inner cylinder IV
1.0mA to be 2 ⊙, giving 8 ⊗ flowing in the outer cylinder.
III
II
(c) Draw Fields: Use Right-Hand Rule for a wire on the total current
passing through the surface bounded by the Amperian path in each
region to determine the direction of the field. The field lines will be
closed circles. In region I and II the total current enclosed is out of
the paper, so the field is counterclockwise. Somewhere in region III,
the total current enclosed is zero and the field changes orientation. In
region IV , the total current enclosed is into the paper, so the direction
of the field is clockwise. We do not draw the field in region I and III
because we have not been asked to calculate the field in these regions.
IV
Bi ≡ Magnitude of Magnetic Field in Region i
III
d~l ≡ Element of length pointing along path
II
IC ≡ Current enclosed by path
I ~r ≡ Radius of Path
Ii = 1.0mA ≡ Current through the inner conductor
Io = 4.0mA ≡ Current through the outer conductor
III ≡ Total Current enclosed in Region II
IIV ≡ Total Current enclosed in Region IV
Strategy: Argue that the field is constant along a circular path. Use Ampere’s Law and divide by the length of
the path to get the field, then use Right-Hand Rule for a Wire to get direction of B.
(a) Draw Good Figure: The diagram was drawn in Example 27.2 Drawing Magnetic Fields for Cylindrical
Geometry. The problem only asks for the field in Regions II and IV . Draw a sample Amperian path on the
diagram.
(b) Apply Ampere’s Law: Apply Ampere’s Law to a circular path. The current, IC , through the surface
bounded by a closed curve, C, is related to the integral of the magnetic field B along the curve by
I
~ · d~l = µ0 IC .
B
~ is constant and parallel to every d~l in the path, we can write Ampere’s Law for this situation as
Since B
I
B dl = µ0 IC
The integral in this case is always the circumference of the Amperian loop, 2πr, so the magnitude is given by
µ0 IC
B=
2πr
and the direction is found by using the right-hand rule on the net current enclosed.
(c) Apply Ampere’s Law the Region II: Apply Ampere’s Law to a circular Amperian path in region II. Use
the right-hand rule for Ampere’s law to select out-of the page as the positive direction for current. The current
passing through the surface bounded by an Amperian path in region II is
IC = Ii = 1.0mA
where Ii is the current flowing in the central conductor. Using the form of Ampere’s Law specialized to the
symmetry in Region II,
B~ II = µ0 Ii counterclockwise
2πr
where the direction can be deduced from the field map or from the direction of the Amperian path and the positive
current direction.
(d) Apply Ampere’s Law in Region IV: Apply Ampere’s Law to a circular Amperian path in region IV . The
total current flowing through the surface bounded by the Amperian path is
IIV = Ii − Io = −3.0mA.
where out-of the page represents positive current. Outside of the cylinders, in Region IV , the negative sign of
the enclosed current indicates that the direction of the magnetic field is opposite to that in Region II. Applying
Ampere’s Law gives
~ IV = µ0 (Ii − Io )
B counterclockwise
2πr
or since Io > Ii
~ IV = µ0 (Io − Ii )
B clockwise
2πr
Solution
Definitions
III
Magnetic Field
Amperian Path
Strategy: Argue field constant along circular path. Use Ampere’s Law and divide by the length of the path to
get field, then use Right Hand Rule for a Wire to get direction of B. Compute the current enclosed in the tube
as the ratio of the enclosed area to the total area.
(a) Specialize Ampere’s Law for Cylindrical Symmetry: A sample Amperian path is drawn on the figure, by
symmetry the field must point along the path. The current, IC , through a surface bounded by the closed curve,
C, is related to the integral of the magnetic field B along the curve by
I
~ · d~l = µ0 IC .
B
~ is constant and the field is parallel to ~l, we can write Ampere’s Law for this situation
Since the magnitude of B
as I
B dl = µ0 IC
The integral in this case is always the circumference of the Amperian loop, 2πr, so the magnitude is given by
µ0 IC
B=
2πr
and the direction by the right-hand rule using the net current enclosed.
(b) Apply Ampere’s Law to Region I: In region I, the total current encircled by any Amperian path is zero,
so the magnetic field is zero.
~I = 0
B
(c) Apply Ampere’s Law the Region II: The total current encircled by an Amperian path of radius r through
region II changes as the radius of the path changes. The total current flowing through the conductor is I. The
fraction of this current encircled by the Amperian path is
Aenclosed
IC = I
Atube
where Aenclosed is the cross-sectional area enclosed by the Amperian path and Atube is total cross-sectional area
2 2
of the conducting tube. The two areas are Atube = πrouter − πrinner and Aenclosed = πr2 − πrinner
2
, therefore
πr2 − πrinner
2
IC = I 2 2
πrouter − πrinner
Substituting this value into the expression for Ampere’s Law appropriate for this symmetry allows the calculation
of the magnetic field:
2 2
B~ II = µ0 I r − rinner counterclockwise
2
2πr router 2
− rinner
The direction of the field is found by pointing the thumb of your right hand in the direction of the current, then
your fingers curl in the direction of the field, that is by applying the Right Hand Rule for a wire to the tube.
(d) Apply General Ampere’s Law the Region III: The total current encircled by an Amperian path in Region
III is IC = I = 5A. Substituting this value into Ampere’s Law yields
~ III = µ0 I
B counterclockwise
2πr
The direction of the field the same as BII because the direction of the total enclosed current in the same.
(a)Using figure (b) and the fact that symmetry implies the field is uniform and parallel to the current
sheet, draw the magnetic field of the sheet. I have modelled the sheet as a collection of parallel wires
to help you figure the direction of the field.
(b)The appropriate Amperian path is drawn in figure (a). Write the general form of Ampere’s law and
break up the integral along the four parts of the path. The integral along the path from b to c and
from d to a can be discarded. Why?
(c)Let the distance from a to b (and from c to d) be ℓ. By symmetry the field must have an equal
magnitude on the two parts. Do the integral and report the formula for the magnetic field of a sheet
of current.
(d)Use your formula. If 20A flows down a sheet that is 2m wide, calculate the magnetic field near the
sheet.
Region I
d c
a b
Region II
Using the right hand rule for a wire and the information given in
the problem the field is as drawn.
Z Z b Z d
~ · d~ℓ =
B ~ · d~ℓ +
B ~ · d~ℓ = µ0 Ienc
B
C a c
The magnetic field is constant along both paths and parallel to the path
Z b Z b Z b
~ · d~ℓ =
B Bdℓ = B dℓ = Bℓ
a a a
So
2Bℓ = µ0 Ienc
The current passing through the path is λℓ
2Bℓ = µ0 Ienc = µ0 λℓ
µ0 λ
B=
2
µ0 λ (4π × 10−7 Tm
A )(10A/m)
B= = = 2π × 10−6 T
2 2
27.4 Solenoids
Ampere’s law can also be used to calculate the field of an infinite solenoid. The reasoning is very different
from the cylindrical case. Unless one counts the toriodal solenoid, which is a solenoid wrapped in a circle, there
are not any other systems where the reasoning applies. You are required to understand the argument, but I will
not ask you to reproduce it.
Solution
Strategy: Choose Amperian Path along axis of solenoid, argue that only the inside leg contributes, then use
Ampere’s Law.
(a) Draw the System and Select the Path: The figure with
the directions of the currents labelled is shown to the right. The
Amperian path is drawn.
d c
a b
Amperian Path
(b) Use Symmetry: The solenoid is cylindrically symmetric about its axis, so the field must have that symmetry.
There are a few possibilities: (1) Radially outward like the electric field of an infinite line charge, (2) Circles around
the axis of the solenoid like the magnetic field of a long wire running down the axis of the solenoid, or (3) Straight
lines parallel to the axis with field strength that depends only on the distance from the axis, B ~ = B(r)x̂ where x̂
is the direction of the axis. The first possibility implies a line of magnetic charge down the axis and therefore is
physically impossible. The second possibility violates Ampere’s law. If an Amperian path were to follow one of
the field circles there would be a net field around the path, but no current flowing through the path. Therefore
the field lines must be straight lines parallel to the axis of the solenoid of the form B ~ = B(r)x̂ .
(c) Reason Field Along Path: The field is perpendicular to the Amperian Path from b to c and from d to a, so
B~ · d~l = 0. What about the leg c to d? Choose an Amperian path where the leg c to d is at infinity. Infinitely far
from the solenoid all the current in the solenoid appears to be flowing at the axis. Since each element of current
flowing in one direction has an equal and opposite current flowing in the opposite direction, their magnetic fields
cancel if we are far enough from the solenoid and the field at infinity is zero. Therefore the contribution of leg c
to d is zero.
(d) Compute Current Encircled: If the length of a to b is L, then the current enclosed by the path is IC = nLI.
Since n is the number of turns per unit length, nL is the number of times the wire passes through the surface.
If this is multiplied by the current, the total current through the Amperian path is found IC = nLI.
(e) Use Ampere’s Law: The current, IC , through a closed curve, C, is related to the integral of the magnetic
field B along the curve by I
~ · d~l = µ0 IC
B
therefore
Bab L = µ0 IC = µ0 nLI
Bab = nµ0 I
Exactly what we expected.
(f) Apply Right-Hand Rule for Wire: Use the Right-Hand Rule for a Wire to get the direction of the field.
Grab one of the wires with your right hand so your thumb points in the direction of current flow. The tips of your
fingers inside the solenoid point in the direction of the field inside. This is because the field inside must add up
from the individual elements of current.
We actually found this result directly from the Biot-Savart law last chapter by first finding the field of a ring,
then integrating to get the field of the finite solenoid, then letting the solenoid length go to infinity. This was
profoundly easier.
We know the number of wires intersecting the surface bounded by the Amperian path, so we simply multiply this
by I to get the encircled current:
Ienc = N I
By the symmetry of the situation, the field inside the solenoid will always be perpendicular to the radial vector,
which makes it parallel to our Amperian path, and the field will have constant magnitude along the path. The
direction of the field will of course depend on the direction of the current, but if we assume that the line integral
~ · d~ℓ = Bdℓ, by
is evaluated in the same direction as the field. Since the field and the path are always parallel, B
the properties of the dot product of parallel vectors. Since the field is constant on the path, we can re-write the
integral part of Ampere’s law as I I I
B~ · d~ℓ = Bdℓ = B dℓ
The remaining integral is just the length of the path, the circumference of the circle, 2πr
I
~ · d~ℓ = B(2πr)
B
Ampere’s Law is I
~ · d~ℓ = µ0 Ienc
B
and from this relationship we can now solve for the field.
B(2πr) = µ0 N I
µ0 N I
B=
2πr
This equation tells us that, unlike an infinite straight solenoid, the field inside a toroidal solenoid is not constant,
but is stronger closer to the inner surface. The field outside the solenoid is zero.
Magnetic Force
F~ = q~v × B.
~
Remember × represents a vector cross product—this makes the magnetic force on the
particle perpendicular to both the particle’s velocity and to the magnetic field it is
moving through. Magnetic forces act “sideways,” to put it crudely.
Force on a Small Segment of Current-Carrying Wire: A current-carrying wire feels
a magnetic force if it is in a magnetic field. The force, dF on a small segment depends
on the current I through the wire, the segment’s length and orientation d~ℓ, and the
~
magnetic field B
dF = Id~ℓ × B.~
If the magnetic field is constant over the length of a straight wire, the total force on
the wire is
F = I ~ℓ × B,
~
where |~ℓ| is the length of the wire. Note: The vector ~ℓ points in the direction of the
current flow.
To apply the Lorentz force, we have to work out the same kind of cross products we have been using throughout
magnetic fields. Both + and − charges feel the magnetic force, so we have to be careful of the sign of q when
evaluating the direction of the Lorentz force.
311
28.1. MAGNETIC FORCE CHAPTER 28. MAGNETIC FORCE
Solution to Part(a)
z out of page
Solution to Part(b)
Solution
Definitions
y B
~ = 4.0 × 10−7 Tẑ ≡ Constant, uniform magnetic field
B
I = 3.0A ≡ Current
I wire ℓ = 12cm ≡ Length of current-carrying wire
F~m ≡ Magnetic force on the wire
z x
(a) Use Lorentz Force: The force on the wire is given by the Lorentz Force
F~m = I~l × B.
~
(b) Use Right-Hand Rule: Use RHR to get the direction of the force, point the fingers in ~l direction and curl
~ direction, giving the direction of F~m as the −ŷ direction.
them in the B
(c) Use Magnitude Form of Cross Product:
Fm = IℓB sin θ
The angle between the current and the magnetic field is 90◦ , and sin 90◦ = 1, so
Fm = IℓB
Example 28.3 Figuring Out the Direction of the Field You Need
Problem: A 10cm wire carrying a current of 6A in the +ŷ direction lies parallel to the y-axis.
(a)What direction must a uniform magnetic field be in to generate a force on the wire in the +z
direction?
(b)What must the magnitude of this field be to exert a force of 0.10N on the wire segment?
Solution to Part(a)
so for a force in the +ẑ direction if ~ℓ = |ℓ|ŷ, we need ẑ = ŷ × B̂. Executing the right hand rule gives B̂ = −x̂.
Solution to Part(b)
The magnitude of the force for the geometry given above is F = ILB, so
F 0.1N
B= = = 0.166T
IL (6A)(10cm)
With the Lorentz force, we complete the set of equations that govern the motion of charged particles in static
electromagnetic fields.
Maxwell’s Equations: Maxwell’s Equations and the Lorentz force as introduced to this
point are:
Maxwell’s Equations
~ · n̂)dA = Qenclosed
R
Gauss’ Law S
(E ǫ0
~ · n̂)dA = 0
R
No Magnetic Monopoles (B S
Lorentz Force
F~ = q E
~ + q~v × B
~
Magnetic Field Does No Work: The magnetic force is always perpendicular to the
velocity, F~ = q~v × B,
~ which implies that the magnetic force is always perpendicular
to ∆~x, the particle’s displacement during a short time interval ∆t . Therefore, the
magnetic field does no work, since the dot product of perpendicular vectors is always
zero—
W = F~ · ∆~x = 0.
Speed Does Not Change when Travelling in a Magnetic Field: If the magnetic
field does no work, then it does not change the energy of the particle. If the energy
stays the same, the magnitude of the velocity cannot change since the kinetic energy is
1 2
2 mv . Therefore, a particle moving in a magnetic field changes direction, but moves
at a constant speed.
We will use these properties later in the section to describe the motion of a charged particle in a uniform
magnetic field. First, though let’s consider the motion of charged particle where the magnetic force is balanced
by some other force.
Solution
Definitions
Earth
Strategy: Use Newton’s Second Law to balance the gravitational force against the Lorentz Force.
(a) Write Force Balance: The problem states the magnetic force balances the gravitational force, so
F~g + F~m = 0,
and therefore
|F~g | = |F~m |.
The direction of the gravitational force is downward and therefore the direction of the magnetic force is upward.
(b) Compute Gravitational Force: The magnitude of the force of gravity is Fg = mg.
(c) Compute Magnetic Force: The magnetic force on the proton is F~ = q~v × B. ~ We will select a magnetic
field perpendicular to the velocity, so the magnitude of the magnetic field is Fm = qvB.
(d) Compute the Magnetic Field: Balance the magnetic and gravitational force,
Fm = qvB = mg = Fg .
The direction of the fields can be found by using ~v × r̂ from the Biot-Savart Law and the Right Hand Rule. The
proper directions are shown in the figure. Since the displacement vector from Q2 to Q1 in figure (b) is parallel to
the velocity of Q2 , the field at Q1 due to Q2 is zero.
(a) (b)
B12 v
Q2 v Q2 F12
F12 B12
F21
B21 = 0
Q1 Q1
v v
B21
so the direction is found with the right hand rule. In figure (b), Q2 produces zero magnetic field at Q1 , so at the
instant in time represented by the figure, the motion of Q1 does not interact with the field of Q2 , so there is no
resulting Lorentz force. See the figure above.
Newton’s Third Law does not apply to magnetic forces. It seems to apply in the figure (a), but this is a coincidence.
It clearly does not apply in figure (b).
Strategy: Use Newton’s Second Law to relate the magnetic force to the centripetal acceleration and solve for
the radius.
(a) Use Right Hand Rule to Get Direction of Circle: Using the RHR on the q~v0 × B ~ factor in the Lorentz
force gives an initial force to the right and a trajectory as drawn in the diagram.
(b) Use Formula For Centripetal Acceleration: The acceleration of a particle moving in a circle at a constant
speed is
v2
ac =
r
and is directed toward the center of the circle.
(c) Compute Magnetic Force: The force on a particle moving in a field is given by the Lorentz force,
F~m = q~v × B.
~
~ and we’ve already picked the correct orientation of the circle, Fm = qvB will be
Since ~v is perpendicular to B
directed inward, toward the center of the trajectory circle.
(d) Apply Newton’s Second Law: The force on the proton equals the mass times the acceleration, by Newton’s
Second Law. Since both the force and acceleration are directed toward the center of the trajectory circle, we can
write
v2
Fm = mac = m
r
mv 2
qvB =
r
mv
r=
qB
(e) Substitute and Solve:
1.67 × 10−27 kg · 0.50 × 103 ms
r= = 0.13m
1.602 × 10−19 C · 4.0 × 10−5 T
Compute the Period: Since magnetic fields do no work, the velocity is constant (Conservation of Energy).
The period is the time it takes the proton to go around the circle, so it’s the circumference of the circle divided
by the magnitude of the proton’s velocity,
2πr 2π(0.13m)
T = = = 1.6 × 10−3 s.
v0 0.5 × 103 m/s
For calculating the period, it doesn’t matter which way the particle is going around the circle. Periods aren’t
negative, so we always just use the magnitude of the velocity.
Problem: For the wires to the left, which carry the same
magnitude of current:
P
B Fm
From the drawing, the magnetic field points to the left at P . The Lorentz force is F~m = q~v × B.
~ Using the right
~ points upward at point P , so F~m points downward since q < 0 for an electron.
hand rule, ~v × B
Solution to Part(a)
FB
Solution to Part(b)
Use F~ = q~v × B
~ and the right hand rule to get the force drawn in the figure above.
Magnetic Dipoles
Shape of Magnetic Dipole Field and Dipole Moment Vector: The shape of a
magnetic dipole field resulting from an object with magnetic dipole moment vector m
~
is shown below. The vector m ~ will also be called the magnetic moment.
321
29.1. DRAWING AND READING MAGNETIC FIELD MAPS CHAPTER 29. MAGNETIC DIPOLES
An Isolated Loop of Current Produces a Dipole Field: The field of a ring of wire
which carries current I in the direction drawn is shown below. To find the direction of
the magnetic moment use the right hand rule for the wire to figure out the direction of
the field interior to the loop, the magnetic dipole moment points in the same direction
as the field in the center of the loop.
m
~ = N IAn̂ m
Solution
Strategy: Draw field stubs for North and South poles, then connect the field lines both outside and inside the
magnet.
(a) Draw the Field in the Magnet: Draw field line stubs
leaving the North pole end of the magnet and entering South
pole end of the magnet. The lines should extend through the
magnet.
N S
S N
(b) Draw Field Map: Using the same technique used to draw
electric field lines for fixed charge, draw a field map outside of
the magnets by connecting and smoothly bending the line stubs.
Connect the field lines between the magnets since the field lines
point in the same general direction where they would (but cannot)
intersect. N S
S N
Magnetic Dipole Field: The full mathematical form of the field of a magnetic dipole,
m
~ is:
~ r) = µ0 3r̂(m
B(~
~ · r̂) − m
~
.
4π r3
This is the field of a point dipole or the field far from a system with a magnetic dipole
moment.
If we chose the +ŷ direction as the direction of the magnetic dipole, we can calculate the magnetic field along
the x and y axis.
Simplified Magnetic Dipole Fields: The expression above for the magnetic dipole
field can be simplified if a direction for the dipole moment is chosen and only the
strength of the field along the axes is computed. If m
~ = mŷ, then along the +ŷ axis,
~ y, 0) = µ0 2mŷ .
B(0,
4π |y|3
The magnetic moment of a permanent magnet depends on the shape and material of the magnet and must
be measured in most cases.
S
(b)Draw the orientation about its fixed center so
that the magnet would come to rest, as if there
was some force which slowed its motion.
Solution to Part(a)
Solution to Part(b)
The magnet will rotate so the north pole points along the field, that is so the magnetic moment aligns with the
field.
S
m
S
N
(a) (b)
Current Into Page Current Out of Page Current Into Page Current Out of Page
A coil carrying a current also has a magnetic moment. You can use the right-hand rule for magnetic moments
to determine the direction of the moment OR use the right-hand rule for a wire to figure out what the field looks
like, then figure out what direction the magnetic moment must point to produce the field.
Wire
Use the right-hand rule to figure out the direction of the magnetic field. Point the thumb along the direction of
the current and the fingers will curl in the direction of the field.
Solution to Part(b)
F1
F2
m
B1
Wire
~ri × F~i
X
~τ =
where ~ri is a vector pointing from the center of mass of an object (or the point where it is pivoted) to the point
where the force F~i is applied. If an object has net torque applied to it, its speed of rotation will increase. If an
object is not rotating, then the net torque on the object is zero.
~τ = m ~
~ × B.
Direction of Rotation Due to Torque : If you grab the torque vector with your right
hand and point your thumb along the vector, your fingers will curl in the direction that
the object is angularly accelerated, or in the direction that it would rotate if it started
from rest.
Since an external agent has to exert a torque to rotate a magnet in a magnetic field away from its equilibrium
position, the magnet’s potential energy changes as it is rotated. The form of the magnetic potential energy is
very similar to the electric potential energy of a dipole in an electric field.
U = −m ~
~ ·B
First, let’s do something that does not require us to understand torque mechanically just to get our feet wet.
x
Side view of
loop.
Solution to Part(a)
The magnetic moment of a current loop is defined as m ~ = N IAn̂ where N is the number of turns, n̂ is the
normal, I is the current, and A = πr2 is the area of the loop with radius r. Substituting gives,
~ = (1)(0.1A)(π)(0.15m)2 = 0.0071Am2
|m|
Solution to Part(b)
The direction of the magnetic moment is normal to the loop, n̂, as shown below. The direction is found by curling
your fingers in the direction of the current, using your right hand. Your thumb will point in the direction of the
normal. From the diagram below, you can see the normal makes an angle of θn = 120◦ + 90◦ = 210◦ with the
field. The torque on a current loop is given by ~τ = m
~ × B.~ Using the right hand rule, the direction of the torque
is into the page, which is the −ŷ direction. The magnitude of the torque is
Solution to Part(c)
x
Side view of
loop.
Now, let’s use the magnetic torque to do something, like lift a weight.
Solution
Definitions
B
m
~ ≡ Magnetic Moment of Loop
I ≡ Current in loop
Current r M ≡ Mass of Object hung from Loop
loop
pivot m F~g ≡ Force of Gravity
~τg ≡ Torque Due to Gravity
~τm ≡ Torque Due to Magnetic Field
N = 50 ≡ Number of Turns of Wire
z mass M
y into page Fg
x
Strategy: Balance the torque due to gravity against the torque applied by the hanging weight.
(a) Compute the Magnetic Moment of the Loop: The magnitude of the magnetic moment of a current loop
is the current multiplied by the area. The direction of the magnetic moment is found by curling your fingers in
the direction of current and your thumb points in the direction of the moment. Therefore, the magnetic moment
of our current loop is
~ = N Id2 (−ẑ).
m
~ = B x̂ is
(b) Compute the Torque Due to the Magnetic Field: The torque due to a magnetic field, B
using the right-hand rule. The torque points out of the page, so the loop would tend to rotate in the direction
that your fingers curl, if your thumb points in the direction of the torque—counterclockwise. (Notice that this
will always give you the same direction as the shortest distance from m ~ This rotation is balanced by a
~ to B.)
mass hanging from the loop.
(c) Compute the Torque Due to Gravity: The vector from the center of the loop to the edge is ~r = d2 x̂ and
the force of gravity is F~g = −M gẑ, so the torque due to gravity is
d dM g
~τg = ~r × F~g = x̂ × (−M gẑ) = ŷ
2 2
using the right-hand rule.
(d) Balance the Torques: If the loop does not rotate, then the total torque is zero so
dM g
~τg + ~τm = 0 = ŷ − N IBd2 ŷ
2
Solve for M ,
2N dIB 2(50)(0.1m)(2A)(0.1T)
M= = = 0.2kg
g 9.81 sm2
which is pretty good.
m m
You also saw that one magnet attracted another magnet, if their poles were aligned. The magnets repelled if the
poles were anti-aligned. Torque was discussed first, because it is actually the primary (highest order) effect of a
magnetic field on a dipole. Unless the field is non-uniform, there is no magnetic force on a dipole.
Uniform Magnetic Field Exerts Zero Net Force on a Magnetic Dipole: If the
magnetic field is uniform and the same at all points in space, the net magnetic force on
a magnetic moment is zero.
Therefore, if we are discussing the repulsion or attraction of magnetic moments, we are discussing non-uniform
fields. Our experience with the attraction or repulsion of permanent magnets can be used to predict the behavior
of magnetic moments in non-uniform fields. Recall:
m m
m m
With these rules in mind, all we have to do to understand the motion of a dipole in a magnetic field is to replace
one of the magnets with the field it produces.
F F
S N S N S N
m m
Anti-aligned Dipole Feels Force Toward a Weaker Field: If the magnetic moment
of a permanent magnet or coil points in the opposite direction to the magnetic field, it
will feel a force toward decreasing magnetic field strength, as shown in figure (a) below.
This can be deduced by imagining a permanent magnet that could produce the field,
figure (b), and then using the properties of magnets.
F F
N S S N N S
m m
Solution
S
Fg
Solution
(a) Use either the right hand rule for a wire or the magnetic
moment of the current loop to figure out the direction of the
field. The magnetic moment of the permanent magnet points
in the opposite direction of the magnetic field (at the magnet); F
therefore, the magnet is repelled from the region of strongest field N
and is pushed toward weaker fields. The force, therefore, points m
upward. S
S
N
The magnitude of the force is given by an expression similar to the one used to find the force an electric field
exerts on an electric dipole.
Force of Magnetic Field on a Magnetic Dipole: The net force on a magnetic dipole
with dipole moment m ~ = B(y)ŷ,
~ in a magnetic field that points in the ŷ direction, B
is
dB
Fnet = m (cos θ)
dy
where θ is the angle between the dipole moment vector and the y axis.
Magnetic Materials
It seems reasonable that the total magnetic moment of either a permanent magnet or the temporary magnetic
moment of the refrigerator depends on the size of the magnet or the amount of the refrigerator that is magnetized.
Since the total magnetic moment of the material, m,~ depends on the amount of material, it makes sense to define
a magnetic moment density.
~ , of a
Magnetization: The magnetization density or simply the magnetization, M
material is the magnetic dipole moment per unit volume,
~ =m
M
~
V
where m
~ is the magnetic moment of the material and V is its volume.
The above is the case for a permanent magnet; it has a magnetization without applying an external magnetic
field. If we place any material in an external field, there will be a magnetic response producing a magnetization
M~ . The field due to this magnetization must be added to the external field to yield the total field in the material.
334
30.1. INTRODUCTION TO MAGNETIC MATERIALS CHAPTER 30. MAGNETIC MATERIALS
~ in a material
Magnetic Field in Material in an External Field: The magnetic field, B,
~ in an external applied field B
with magnetization M ~ 0 is
~ =B
B ~ 0 + µ0 M
~
We will find that for some materials the magnetization is in the same direction as the applied field and therefore
the total field is greater than the applied field. For some materials, the magnetization is in the opposite direction
as the applied field and the total field in the magnet is lower than the applied field. For materials with a weak
magnetic response (what we will call paramagnetic and diamagnetic materials) there is an approximately linear
relation between the magnetization and the applied field at low applied fields.
~
~ = χm B0
M
µ0
Relation Between Applied Field and Total Field for Linear Magnetic Materials: If
a magnetic material has a linear relation between the applied field and the magnetization,
that is if χm is constant, then the total field is related to the applied field by
~ =B
B ~ 0 + µM
~ =B
~ 0 + χm B
~ 0 = (1 + χm )B
~ 0 = Km B
~0
Permeability and Permittivity of a Material: We have been working with the per-
mittivity of free space, ε0 , and the permeability of free space µ0 . It is common practice
to use the permeability of a material as µ = Km µ0 and the permittivity of the material
as ǫ = κε0 .
For all normal paramagnetic and diamagnetic materials the effect is so small that it is difficult to observe.
The value of χm is so small that when we tested copper and aluminum in lab we saw no magnetic response. A
permanent magnet will slightly attract a paramagnet and slightly repel a diamagnet. A superconductor is a perfect
diamagnet, Km = 1 + χm = 1 + (−1) = 0 and reduces the magnetic field in its interior to zero. Therefore,
a superconductor is strongly repelled from a permanent magnet which is why a superconductor floats over a
permanent magnet.
30.2 Ferromagnetism
30.2.1 Ferromagnetism
Both diamagnetism and paramagnetism are small effects and can be understood in terms of the behavior of
individual atoms. Ferromagnetism is an effect involving many atoms and in certain materials is a gigantic effect.
Ferromagnetism arises when the atomic magnetic moments on different atoms tend to align with one another.
This produces large regions involving billions of aligned moments called domains. When a field is applied, the
domains align producing a large magnetization. This magnetization remains after the field is removed, producing
a permanent magnet. Ferromagnetism occurs in only a few materials; iron, nickel, cobalt, and rare earth elements
like gadolinium and dysprosium.
We can represent the process of magnetizing and then de-
magnetizing a ferromagnet by the curve to the right. The curve B
plots the applied field Bapp against the resulting field in the ma- Bs b
terial B. The material starts out at point (a) where there is zero saturation
applied field. As Bapp increases the magnetic field in the material Br
c
rapidly increases. The slope of the curve at point a is the ini-
tial relative permeability, Km0 = B/Bapp . In ferromagnets, Km0
ranges from 200 to 10, 000, so we get a lot of field in the ferromag-
d
net for a small applied field. As the applied field is increased, the a
Br Bapp
field in the magnet increases with a maximum slope of the curve
between point (a) and (b) that ranges up to Km = 1, 000, 000.
The magnetic field in the material continues to increase until all
the domains are aligned. When this happens the magnet is said
to be saturated, and the magnetic field can no longer increased
by increasing magnetization. Saturation happens at point (b).
If we then start to turn off the applied field at the point (b), the magnetization decreases until the applied
field is zero at point (c). There is still a magnet field at the point (c), called the remnant magnetic field Br .
This is the magnetic field of the ferromagnet acting like a permanent magnet. The ferromagnet does not return
to its state of zero field when the applied field is removed. It remembers a field has been applied, it remembers
its history. The curve captures this memory of history and is called a hystereses curve. To erase this remnant
magnetization, apply a field in the opposite direction until at point (d), the magnetic field drops to zero. The
field we have to apply to remove the remnant magnetic field is called the coercive force, Bc . The coercive force
is naturally not a force but a magnetic field.
The table below reports the magnetic properties of some important ferromagnetic materials. Note how small
the coercive force, Bc , is compared to the remnant magnetization and how large the relative permeabilities are.
This gives you an idea of how large a magnetic field you can produce for a small applied field.
Material Km0 Km maximum Br Bs Bc
Iron Commercial 99.9%Fe 250 6000 1.3T 2.16T 8.8 × 10−5 T
Iron Pure 99.9%Fe 10,000 100,00 1.3T 2.16T 1.0 × 10−6 T
Permalloy 4,000 100,000 0.7T 1.05T 5 × 10−5 T
Superpermalloy 100,000 1,000,000 0.7T 0.79T 1.88 × 10−7 T
Cobalt 99% pure 70 250 0.5T
Nickel 99% pure 110 600 0.4T
Mu-metal 50,000 200,000
Mu-metal is a material used for magnetic shielding. As you can see in the table for iron the ferromagnetic
properties of a material are very sensitive to the way the material is processed, which accounts for the difference
between pure iron and commercial iron.
Solution
The field is much less than Bc so we can use the initial Km = Km0 = 250, therefore the field is B = Km B0 =
(250)(1 × 10−5 T) = 250 × 10−5 T
Solution
(a) Compute the Magnetic Moment: The magnetic moment is the magnetization density multiplied by the
volume
m = M V = πr2 hM = π(0.005m)2 (0.001m)(1.02 × 106 A/m) = 0.08Am2
(b) Compute the Magnetic Field: Far from the magnet, we can use the field of a point dipole. If the y axis
is the axis of the dipole, along the axis of a point dipole,
µ0 2m 4π × 10−7 Tm 2
A 2(0.08Am )
B(0, y, 0) = = = 1.28 × 10−4 T
4π |y|3 4π (0.05m)3
Solution
= 0.0283Am2
m
(b) Compute the Magnetic Field: The magnetic field of a dipole is strongest in the direction of the dipole
moment. The magnitude of the magnetic field of a dipole at a distance y along its axis is
µ0 2m (4π × 10−7 Tm
A ) 2(0.0283Am )
2
B= = = 4.52 × 10−5 T
4π |y|3 4π |0.05m|3
(c) Compute Maximum Torque: The torque a magnetic field exerts on a dipole is given by τ = m ~ × B.~ The
maximum torque occurs when the magnetic moment is at right angles to the field. This occurs for the orientation
drawn above. At maximum torque, the magnitude of the torque is
Faraday’s Law
So far we know how to work with systems where the electric or magnetic fields are constant. It would be
natural to assume this would let us work with fields which are changing, by simply recalculating with the new
currents or charges. Unfortunately, the universe is not that simple—changing fields cause additional effects.
where d~ℓ points along the path C and ~v is the velocity of the path. The first term, C E ~ · d~ℓ, is the emf that
R
results from a net electric field along the path and is the negative of the potential difference if the field is generated
by static charges. For static charges, if the path is closed this term is always zero because of R independence of
path. Changing magnetic fields can make this term non-zero for a closed path. The term, C ~v × B ~ · d~ℓ, is a
result of the path’s motion through the magnetic field and is called the motional emf .
where the integral is taken around the path, E ~ is the electric field, and ~ℓ points along
the path. For electric fields due to static electric charges, this integral was always zero
(for example, Kirchhoff’s loop theorem). For paths within changing magnetic fields,
this is no longer the case.
339
31.1. ELECTROMOTIVE FORCE CHAPTER 31. FARADAY’S LAW
Solution
The electric field has a component in the opposite direction to the path all around the path, so the electric field
adds along the path, and there is a negative emf around the path.
B weaker B stronger
B - Out of Page
Solution
Consider what would happen to a positive charge at a few points in the circuit. The charge can’t get out, so it’s
dragged along with the velocity of the circuit, ~v . The charge, therefore, feels a force F~ = q~v × B.
~ In figure (a),
these forces cancel if you take the integral around the loop, so there is no induced EMF—that is no work is done
by the magnetic force as the charge is moved around the loop, so there will be no current. In figure (b), the sum
of the forces around the loop is not zero and there will be a flow of current.
v v
B weaker B stronger
B - Out of Page current
Magnetic Force
where in this case N = 1 and L is the length of the wire in the field.
++
v v
Fm Fm
__
v
y
z x
The forces on the top and bottom are perpendicular to the wire and can do no work along the direction of the
wire. The forces on the left and right add to produce a clockwise EMF. A + charge has positive work done to it
if it travels around the loop. Therefore a current will flow in the clockwise direction around the loop.
The electromotive force on the left side of the loop will drive the current in the clockwise direction and has
magnitude
~ · d~ℓ = N vBL = (100)(10 m )(0.2T)(10cm) = 20V
Z
emflef t = N ~v × B
s
The electromotive force on the right side also tends to drive the current in the clockwise direction and has the
same magnitude as the emf on the left side. The total emf is the sum of the two
Solution
~ · d~ℓ. Since the velocity and the
R
The motional emf for a moving conductor in a magnetic field is emf = C ~v × B
~ = vB. If you are driving to the west, the direction of ~v × B
magnetic field are perpendicular, |~v × B| ~ is downward.
~
Therefore, ~v × B is perpendicular to a path between the bumpers and to a path between the car doors, so the
~ and the emf along the path
emf along these paths is zero. A path from the roof to the floor is parallel to ~v × B
is
~ · d~ℓ = vBℓ = (36 m )(2.2 × 10−5 T)(1.5m) = 0.0012V = 1.2mV
I
emf = ~v × B
s
an observable, but not particularly useful voltage. Do not worry about the sign. You would get the opposite sign
if you assumed you were driving toward the east.
Units of Magnetic Flux: Flux is measured in Webers (Wb) where 1Wb = 1Tm2 .
You can visualize the magnetic flux through a surface as the number of magnetic field lines passing through
the surface.
Example 31.5 Field through Flat Loop
Problem: Our best UPII lab magnets produce a field of about 0.25T. Compute the magnetic flux due to this
field if the direction of the field is parallel to the normal of a circular loop of wire 1cm in radius.
Solution
The magnetic flux through a loop of wire is φm = N BA, if the field is parallel to the normal of the loop. Since
the number of turns is not given, the number of turns is N = 1. The magnetic field is B = 0.25T, and A = πr2
is the area of the loop. Substituting gives
φm = N BA = N Bπr2 = (1)(0.25T)(π)(1cm)2 = 7.9 × 10−5 Wb
Example 31.6 Magnetic Flux through a Coil Tipped with Respect to the Field
Problem: A circular coil of wire with 20 turns of wire and radius 20cm has a normal which makes an angle of
20◦ with a uniform magnetic field of magnitude 0.25T. Compute the magnetic flux through the coil.
Solution
~ · n̂, where A = πr2 is the area, N = 20
If the magnetic field is uniform, the magnetic flux is given by φm = N AB
is the number of turns, and B = 0.25T is the magnetic field. We are given the angle the normal makes with the
~ · n̂ = B cos 20◦ . Substituting gives
field as 20◦ , so B
φm = N πr2 B cos 20◦ = (20)(π)(20cm)2 (0.25T) cos 20◦ = 0.6Tm2
t ≡ Time in Seconds
N = 1 ≡ Number of Turns
A(t) = πR(t)2 ≡ Area of Loop at Time t
dR/dt
R(0) = 0.12m ≡ Initial Radius
R(t) ≡ Radius at Time t
φm (t) ≡ Magnetic Flux at Time t
dR mm
= −3.0 ≡ Rate of Change of Radius
dt s
B = 4.0µT ≡ Magnitude of Magnetic Field
Strategy: Compute the flux as a function of time by writing the area as a function of time.
(a) Apply the Definition to Magnetic Flux: Since the magnetic field is constant, the flux is φm = N B ~ · n̂A.
~ · n̂ = |B| and φm = N BA.
(b) Compute the Dot Product: The normal to the loop is parallel to the field, so B
(c) Compute Time Dependent Area: Since the radius is decreasing at 3mm/s,
dR
R(t) = R(0) − t.
dt
2
dR
A(t) = πR2 (t) = π R(0) − t
dt
(d) Put it All Together:
φm (t) = N BA(t)
2
dR
φm (t) = πN B R(0) − t
dt
2
mm
φm (t) = π(4.0µT) 0.12m − 3 t
s
Qualitative Statement of Faraday’s Law: If the magnetic flux through a closed path
is changing, there is an emf induced around the path. The faster the flux changes, the
larger the emf . The direction of the induced current is given by Lenz’ law.
Lenz’s Law: The induced emf and induced current are in such a direction as to oppose
change in the magnetic flux.
Changing magnetic flux can be caused by a moving or distorting loop or a changing magnetic field. If the
magnetic field is changing, the emf is caused by an electric field, so a changing magnetic field generates an electric
field.
Solution
By Faraday’s Law, an emf can be induced in the loop by creating a changing magnetic flux, φm , through the
loop. A changing magnetic flux can be created by the variation of the area of the loop, the orientation of the
~ · ndA. Since
R
loop, the magnitude of the magnetic field, or the direction of the magnetic field; since φm = N s B
the loop must remain fixed, only the magnitude or orientation of the magnetic field can be changed. This can be
done, for example, by moving a permanent magnet in a circular path around the plane of the loop.
We can also reason about the size and sign of the voltage we would measure across a loop experiencing a
change in flux. Faraday’s law states the faster the flux changes, the greater the voltage. Lenz’ law gives the
direction of the current that sets up a magnetic field, which resists the change in flux. If the direction of induced
current is known, a potential difference can be selected to establish that current. How do we use Lenz’ Law?
Given a situation with changing flux through a loop, use the Right Hand Rule for a wire, grabbing the circuit and
making sure your fingers curl so that the field IN the loop opposes the change in field. Your thumb points in the
direction of the induced current. Select a sign for the voltage, remembering the current flows from high potential
to low potential, so as to cause the current to flow in the correct direction.
Solution
Solution
emf
31.3.2 Drawing the Electric Fields Resulting From Changing Magnetic Fields
If a changing magnetic field is generating the emf , then the emf is caused by an electric field generated by
the changing magnetic field. For simple systems, we can draw this electric field.
Solution
Eddy Currents : Eddy currents are currents induced in a solid piece of conducting
material by a changing magnetic flux.
Consider the following case: a steel disk is thrown into a magnetic field. As it enters and leaves the field, there
is a changing magnetic flux through the disk, inducing an emf , which causes a current to flow.
Solution
As the disk enters the magnetic field, the flux through the disk
changes, causing an emf around paths though the disk, by Fara-
day’s law. Since the disk is conducting, this emf will cause a
current to flow. The current direction will act to oppose the
change in flux. The flux is increasing, so the current will circulate
in a direction to produce a magnetic field opposite to the uniform
field. I used the right hand rule for a wire to figure out the di-
rection of the current flow to produce this opposing field. Once
motion
inside the uniform field, the flux is no longer changing, so eddy
currents stop.
Faraday’s Law: For any closed path, C, the emf around the path is related to the
magnetic flux through the surface, S, bounded by the path by
dφm d
Z
emf = − =− B~ · n̂dA,
dt dt S
that is, EMF equals the time rate of change of magnetic flux.
Faraday’s Law for a Fixed Path: A changing magnetic flux through the surface
enclosed by a stationary path causes a net electric field around the path,
~ · d~ℓ = − dφm = − d
I Z
emf = E ~ · n̂)dA,
(B
C dt dt S
where E ~ is the electric field, emf is the electromotive force, C is the path bounding
the surface S, ~ℓ points along the path, φm is the magnetic flux through the surface, n̂
is the normal to the surface, and t is the time.
Faraday’s law allows us to extend the Maxwell equation relating to independence of path to systems with a
changing magnetic field. All that is left to do to complete Maxwell’s equations is to complete Ampere’s law.
Maxwell’s Equations Part III: Maxwell’s Equations and the Lorentz force as intro-
duced to this point are:
Maxwell’s Equations
~ · n̂)dA = Qenclosed
R
Gauss’ Law S
(E ǫ0
~ · n̂)dA = 0
R
No Magnetic Monopoles (B S
~ · d~ℓ = − dφm = − d ~
H R
Faraday’s Law (Independence of Path) C E dt dt S B · n̂dA
Lorentz Force
F~ = q E
~ + q~v × B
~
Solution
Definitions
y
B
y
z x
Strategy: Compute the flux through the loop, apply Faraday’s Law to get the voltage, and Lenz’ Law to get
the sign.
(a) Compute the Flux: Since the magnetic field is uniform over the area of the loop and the loop is perpendicular
to the field, the flux is φm = N BA = N Bℓ2 .
d
emf = − φm
dt
d d
= − (N BA) = − (N Bℓ2 )
dt dt
d d Tm2
emf = −N ℓ2 B = −N ℓ2 (Ct2 ) = −N ℓ2 [2Ct] = −(0.2T/s2 )(100)(0.1m)2 t = −0.2 2 t
dt dt s
The sign of the EMF is related to the choice of direction the path traverses the circle, either clockwise or
counterclockwise; the sign of the flux depends on whether the loop normal points into or out of the page.
Needless to say, this involves another Right Hand Rule. Therefore, use Lens’ law to figure out the current
direction.
(c) Use Lenz’ Law: Since the magnetic field is increasing in the +ẑ direction, the induced EMF will produce a
current that produces a field, that points in the −ẑ direction. By the right-hand rule, the current is clockwise as
viewed from the +ẑ direction.
Solution
Definitions
y
B
y
z x
Strategy: Compute the flux as a function of time, apply Faraday’s Law, then apply Lenz’ Law to get the sign.
(a) Compute the Flux: Since the magnetic field is uniform over the surface bounded by the loop, the flux
through the loop is φm = N (B ~ · n̂)A. The number of turns is not given in the problem so N = 1. Since the
~ · n̂ = B. The flux is then φm = BA(t) where the
normal to the loop is in the same direction as the field, B
area of the loop has been written as a function of time. The loop is circular, therefore the area of the loop is
A(t) = πR(t)2 . The radius R(t) is changing with time and can be written
dR
R(t) = R0 + t = R0 + vt
dt
where R0 is the radius of the loop at time zero and v is the rate the radius is changing. Substituting everything
back together yields
φm = Bπ(R0 + vt)2
(b) Use Faraday’s Law: The flux linking the circuit changes with time and therefore there is an induced EMF
given by Faraday’s Law.
d
emf (t) = − φm
dt
d
= − Bπ(R0 + vt)2
dt
emf (t) = −2Bπ(R0 + vt)v
Evaluate this at t = 0
0.01m
emf (0) = −2BπR0 v = −2(−0.5T)π(0.1m)( ) = 0.052mV = 5.2 × 10−5 V
60s
(c) Use Lenz’ Law: Since the loop is linking more of the magnetic field in the −ẑ direction, the induced field
will be one that points in the +ẑ direction. By the right-hand rule, the current is counter-clockwise as viewed
from the +ẑ direction.
I
~ = 0.1 T t2 ẑ ≡ Magnetic Field
B
s2
φm (t) ≡ Magnetic Flux
A ≡ Area of Loop
t ≡ Time
N = 50 ≡ Number of Turns
Strategy: Compute the flux through the loop as a function of time, then apply Faraday’s Law to compute the
emf. Use Lenz’ Law to compute the direction of the induced current.
Compute Magnetic Flux: Since the field is normal to the loop, the magnetic flux is defined as
2
T 2 Tm
φm (t) = N AB(t) = (50)(5cm)2 (0.1t2 ) = 0.0125t
s2 s2
Apply Lenz’ Law: The magnetic field is increasing in strength out of the page, therefore the magnetic flux out
of the page is increasing. By Lenz’ Law a current will be induced that creates a magnetic field to counteract the
change in flux. The flux is increasing out of the page, so the induced field must point into the page. Using the
right hand rule for a wire gives the induced current direction drawn.
31.4 Generators
A loop rotating in a magnetic field will produce an emf that is a sine wave. To evaluate the motion of the
loop, we need to recall certain definitions about rotational motion and calculate the flux through a rotating loop.
1Hz = 1s−1
Angular Frequency: The angular frequency, ω, is the number of radians an object
rotates per second,
ω = 2πf
Solution
Definitions
B y
rotation axis
y
t=0
z x
Strategy: Compute maximum flux, then use a cosine time dependence and fix the phase angle.
(a) Use Definition of Magnetic Flux: The magnetic flux for a coil in a uniform field is by definition
~ · n̂)A
φm = N (B
This can be re-written using a property of the dot product as
φm = N BA cos θ
For a turning loop in a constant field the only thing changing with time in this expression is the angle the loop
makes with the field, therefore θ is a function of time θ(t). The quantity N BA is the maximum flux, φmax ,
through the loop which happens when the normal of the loop points in the same direction as the magnetic field.
φmax = N BA = N Bπr2 = (10)(0.25T)π(10cm)2 = 0.0785Wb
(b) Write θ as a Function of Time: The function θ(t) gives the angle the normal makes with the field at all
times. This can be written in terms of the angular velocity, ω, and the phase angle, δ.
θ(t) = δ + ωt
Since the normal of the loop is perpendicular to the field at t = 0, θ(0) = 0 = δ + ω(0) therefore δ = 0, and
θ(t) = ωt
(c) Compute Angular Frequency: Convert the period to seconds
T = 1min = 60s
The angular frequency is then
2π 2π rad
ω= = = 0.105
T 60s s
(d) Put it All Back Together:
φm (t) = φmax cos(ωt) = N Bπr2 cos(ωt)
rad
φ(t) = 0.0785Wb cos[0.105 t]
s
Definitions
n̂ ≡ Normal to Loop
A ≡ Area of Loop
B ≡ Magnetic Field
θ
Loop ω ≡ Angular Frequency
n t ≡ Time
f ≡ Frequency
emfmax ≡ Maximum Induced EMF
Direction
of Rotation
Magnetic Field
Strategy: Write magnetic flux as a function of time, then apply Faraday’s Law to get the EMF.
Solution to Part(a)
(a) Compute Time Dependent Flux: The magnetic flux is φm = N AB ~ · n̂, if the field is uniform. The angle,
θ, between the magnetic field and the loop is changing at a constant rate; this can be written as θ = ωt, where
ω = 2πf is the angular frequency and f is the frequency. We can then write
~ · n̂ = N AB cos θ = N AB cos(ωt)
φm (t) = N AB
This choice for the time dependence of θ assumes the angle between the loop normal and the magnetic field is
zero at time t = 0.
(b) Apply Faraday’s Law: The emf produced by the coil is given by applying Faraday’s Law to the time
dependent flux,
dφm (t) d
emf = − = −N AB cos(ωt) = N ABω sin(ωt).
dt dt
The maximum emf occurs when sin(ωt) = 1, so
(c) Solve for Frequency: We wish to compute the frequency at which the generator must turn to produce
emfmax = 10V. Solving for the frequency gives
emfmax 10V
f= = = 113Hz
2πN AB 2π(200)(π)(1.5cm)2 (0.1T)
Substituting into the expression derived in part (a), the maximum emf at f = 60Hz is
emf (t)
emfmax
0V
t
− emfmax
Almost all electrical power is produced by generators, spinning coils in magnetic fields.
Inductance
32.1 Inductance
In electrostatics, a complex system of charge and field is reduced to a single parameter—the capacitance—
when analyzing electric circuits. The magnetic properties of a system of conductors are reduced to a parameter
called the inductance, when analyzing electric circuits. This inductance is the ratio of the flux through the circuit
to the current flowing in the circuit. We can compute the inductance of a single circuit called the self-inductance
or the inductance between two circuits, called the mutual inductance.
φm = LI.
Wb Tm2
1H = 1 =1 .
A A
The only case where we will actually compute the inductance from scratch is the case of an infinite solenoid. The
strategy is very similar to that of computing a capacitance. We take a system of conductors (usually wires) that
have no current running through them and apply an arbitrary current I. We compute the flux resulting from I,
then apply the definition of inductance, and cancel the arbitrary current.
Solution
356
32.1. INDUCTANCE CHAPTER 32. INDUCTANCE
Definitions
Strategy: Introduce a current and compute the magnetic field. Compute the flux and apply definition of
inductance.
(a) Introduce a Current I: Allow a current I to flow in the coil.
(b) Compute the Magnetic Field: The magnetic field in the solenoid in the infinite solenoid approximation is,
B = nµ0 I,
φm
L= = µ0 n2 Aℓ.
I
(e) Substitute and Compute: The number of turns per length, n, is n = N/ℓ. The cross-sectional area of the
solenoid A = πR2 = πD2 /4.
N 2 πD2
L = µ0 ℓ
ℓ2 4
2 2
µ0 N πD
L=
4ℓ
(4π × 10−7 AN2 )(10000)2 (π)(1 × 10−3 m)2
L=
4(15 × 10−2 m)
L = 0.66mH
32.2 Inductors
Inductors are circuit elements that resist a change in current. If we substitute the definition of inductance
into Faraday’s Law we get the following expression for the potential difference across an inductor.
Faraday’s Law for Inductor : The emf across an inductor with inductance, L,
carrying current, I, is
dI
emf = −L ,
dt
where t is the time. An inductor in a circuit carrying a constant current has zero
potential difference. This emf will behave just like a potential difference in an
electric circuit.
Symbol For Inductor: An inductor is represented in an
electric circuit by the symbol to the right.
Solution
d 0.1As2 (0.2s3 )
∆V = −L = V
dt t2 t3
Definitions
L ≡ Inductance
VA t ≡ time
I(t) = γt3 ≡ Current As Drawn
Um ≡ Energy Stored in Inductor
L I(t)
VB
dI(t) dγt3
∆V = −L = −L = −3γLt2 .
dt dt
0 t
Solution
0 t
V(t)
0 t
Solution to Part(a)
We compute the flux in the solenoid in part (c), use that and the definition of inductance as
Solution to Part(b)
The magnetic field of an infinite solenoid is B = nµ0 I, where n = 100/20cm = 500m−1 is the turns per unit
length and I = 0.1A. Substituting gives
Tm
B = nµ0 I = (500m−1 )(4π × 10−7 )(0.1A) = 6.3 × 10−5 T
A
Solution to Part(c)
The magnetic flux linking the solenoid is φm = N AB where A is the area and N = 100 is the number of
turns. The area of the solenoid is A = πr2 . Substituting gives
φm = N AB = (100)π(2cm)2 (6.3 × 10−5 T) = 7.9 × 10−6 Tm2
Solution to Part(d)
Solution to Part(e)
Transformer Output Voltage: If a transformer is formed from two coils where the
first coil is wound with Na turns and the second coil is wound with Nb turns then
the ratio of the input voltage ∆Va to the output voltage is ∆Vb is
∆Vb Nb
=
∆Va Na
Solution to Part(a)
φm 1 × 10−4 Tm2
L= = = 2 × 10−5 H
I 5A
Solution to Part(b)
Solution to Part(c)
Since the loop and the flux are the same M12 = φm,2 /I1 = 2 × 10−5 H
Solution
Strategy: Apply formula for magnetic field of solenoid, then definition of self and mutual inductance.
Solution to Part(a)
The magnetic field of a solenoid is B = nµ0 I where n is the turns per unit length and I is the current. The
turns per unit length is the number of turns, N , divided by the length, ℓ, n = N/ℓ = 400/0.25m = 1600m−1 .
So
B = nµ0 I = (1600m−1 )µ0 I
Solution to Part(b)
To compute the inductance, we need to compute the magnetic flux in terms of an arbitrary current. Using
the field computed above, φm = N AB where A = πr2 is the area of the solenoid, with r the radius of the
solenoid. Substituting everything gives
Solution to Part(c)
Solution to Part(d)
The mutual inductance, Msolenoid,loop , from the flux through the single loop generated by the field of the
solenoid is defined as
Msolenoid,loop = φloop /Isolenoid
where φloop is the magnetic flux through the loop and Isolenoid is current through the solenoid. The magnetic
2
flux through the loop is φloop = Asolenoid B = πrsolenoid B since there is only one turn of wire and the magnetic
field only exists inside the solenoid. Substituting the expression for the magnetic field of the solenoid gives
2
φloop = πrsolenoid nµ0 Isolenoid
Magnetic Energy
Energy Density of Magnetic Field : The energy density of the magnetic field, ηm ,
is
B2
ηm = ,
2µ0
where B is the magnitude of the magnetic field.
365
33.1. MAGNETIC ENERGY CHAPTER 33. MAGNETIC ENERGY
The total energy, Um , in a 1cm cubic volume, V , is density multiplied by the volume
J
Um = ηm V = (6.37 )(0.01m)3 = 6.37 × 10−6 J
m3
RL Circuits
367
34.1. LIMITING BEHAVIOR OF RL CIRCUITS CHAPTER 34. RL CIRCUITS
After an inductor has been connected to a battery for a long time, the current reaches a constant final
value. If the current through an inductor is unchanging, then the potential difference across the inductor is
zero.
Long-Time Charging Behavior: At long times the current in the circuit has reached
a steady value (no change), which means that there is no electric potential difference
across the inductor. This allows the inductor to be replaced by a straight wire.
We can use these observations to analyze the long and short time behavior of RL circuits.
R2
Strategy: Use the fact that at short periods of time, the resistors can be replaced by straight wires. Analyze
the resulting DC circuit.
∆V1 ≈ 0V ∆V2 ≈ 0V
∆V L
(b) Solve the DC Circuit: The resulting circuit is a simple DC circuit and the potential differences can be
solved for using techniques for DC circuits. Since the terminals of the battery are connected directly to the
terminals of the inductor, the potential difference across each is the same,
∆V L = ∆V = 12V
Solution
Definitions
R2
Strategy: Use the fact that at long periods of time, the inductor passes the current unimpeded. Analyze
the resulting DC Circuit.
∆V L
R2
∆V1 = 4.0V
∆V
4
× 10−6 A · 6.0 × 106 Ω
∆V2 = IR2 =
3
∆V2
∆V2 = 8.0V
R2
(d) Compute Voltage Drop Across Inductor: Since the current does not change in the long-time limit,
the potential difference across the inductor is zero.
∆VL = 0
Since the inductor is in series with the resistors, the current through it is the same as through them.
4
I= × 10−3 A
3
34.2 RL Circuits
This section covers computing the functional form of the current and voltage in an RL circuit. The time
dependence of current through the circuit and the potential differences across the inductor and resistor are
the same increasing or decreasing exponential curves we encountered in RC circuits. The time dependence
is characterized by a time constant τ . For an RL circuit,
Time Constant, τ : The time constant of an RL circuit is τ = L/R, where R is the
total resistance and L is the inductance.
Since all the time dependence curves are exponentials involving the time constant, the key step in an RL
circuit problem is to determine whether the time dependence of the quantity of interest is an increasing or
decreasing exponential.
1 − exp(−t/τ ),
which at time t = 0 is zero and increases to 1 at long periods of time. The graph
to the right might be the potential difference across a resistor in a “charging” RL
circuit with a battery with voltage ∆Vb = 10V, and time constant τ = 10s.
10 Vf
7
∆VR (t) (volts)
0
0 5 10 15 20 25 30
t (sec)
exp(−t/τ ),
which at time t = 0 is one and decreases to 0 at long periods of time. The graph
to the right might be the potential difference across a “discharging” inductor with
initial potential difference ∆V0 = 10V, and time constant τ = 10s.
V0 10
6
∆VL (t) (volts)
0 5 10 15 20 25 30
t (sec)
Solution
Definitions
I
R1 = 9.0Ω ≡ Resistance of Resistor 1
∆V1 = 6.0V ≡ Potential Difference Across Battery
L = 3.0mH ≡ Inductance of the inductor
R1
I(t) ≡ Current in the circuit
If ≡ Final current through the resistor
∆V1
L
Strategy: Use formula for charging inductor, and fix the constant from the long-time behavior.
(a) Compute Final Current, If , Through the Circuit: Use Example 34.2 Analyze Long Time Behavior of
an RL Circuit, to get If , which for the circuit in the figure is simply If = ∆V1 /R1 .
I(t) = If [1 − exp(−t/τ )]
The current has its minimum value at t = 0, and then increases to its final value, so the proper form of the
time dependence is
I(t) = If [1 − exp(−t/τ )]
I(t) = 2/3A 1 − exp −t/0.33 × 10−3 s
When two objects are alone in series, their potential drops must be equal and
opposite and the current flowing through them must be the same. The potential
difference across the resistor is always IR. The potential difference across the
inductor is also IL (t)R or
where ∆V0 = I0 R. You also get this result from directly applying ∆VL (t) = −L dI(t)
dt .
Solution
Definitions
I
R1 = 9.0Ω ≡ Resistance of Resistor 1
L = 3.0mH ≡ Inductance of the inductor
R1 I(t) ≡ Current in the circuit
I0 = 4.0A ≡ Initial Current
τ = L/R.
I(t) = I0 exp(−t/τ ).
Since the current in the circuit starts at its highest value I0 and decays toward zero, the correct form of the
time dependence is a decaying exponential.
I(t) = I0 exp(−t/τ )
music source
375
35.1. DEVICES UTILIZING THE LORENTZ FORCE CHAPTER 35. FINAL TOPICS TEST 3
F Fnet F
2 Draw the Force Vectors
35.1.3 Motors
We also built a motor in lab. The motor was composed of a
coil of wire with one end completely sanded and one end half
sanded suspended by a paper clip. Paper Clips
Coil
Magnet
Clay
• (2) A magnetic field exerts torque on the loop that causes it to rotate toward equilibrium. This force
is called the Lorentz Force.
• (3) The current is turned on and off so torque only acts in one direction, causing rotation.
B
I
B
B
F
Paper Clip
Draw the field so it extends to the coil, draw the direction of the field at both ends of the coil, then the
direction of the force using the Lorentz force. Common mistakes made in the description of speakers and
motors.
• (1) Magnetic fields DO NOT interact to produce force. Force is produced by the action of the field on
current, or sometimes atomic current in permanent magnets.
35.2.2 Generators
A generator converts mechanical energy into electrical energy.
Displacement Current
+ _ A
_
I C
I + I
ath
np
a
+ _
p eri
E am
Surface 1 Surface between plates
C
+ _
+ _
_ I
I + I
+ _
Ampere’s Law applies to any surface bounding the path C. Two possible surfaces bounded by C are
drawn above. Surface 1 passes between the plates and surface 2 passes outside of the plates. Since both
surfaces are bounded by C, Ampere’s law should give the same result for both surfaces.
379
36.1. DISPLACEMENT CURRENT CHAPTER 36. DISPLACEMENT CURRENT
Apply Ampere’s law to surface 1. Since no current passes between capacitor plates the current passing
through the surface is Iencircled = 0 and applying Ampere’s law
Z
~ · d~ℓ = 0
B
C
For surface 2, the current flowing through the surface is Iencircled = I and applying Ampere’s law gives
Z
B~ · d~ℓ = µ0 I Whoops!
C
Since both surfaces are bounded by C, both surfaces can be used in Ampere’s law and the results for both
surfaces must be the same.
We are going to fix Ampere’s Law by adding another term that makes the contribution from surface 1
the same as that from surface 2. This term will be called the displacement current, Id .
I
B~ · d~ℓ = µ0 (I + Id )
C
Comparing Ampere’s law with Faraday’s law, we would expect this additional term to depend on a changing
electric flux. The electric flux, φe , through surface 1 is,
σ
φe = EA = A
ǫ0
where I have used the electric field of equal and opposite planes as E = σ/ε0 . The charge density on the
positive plate is σ = Q/A so
A Q Q
φe = =
ǫ0 A ǫ0
Solving for Q gives
Q = ǫ0 φe
The current is the derivative of the charge
dQ dφe
≡ I = ǫ0 = Id
dt dt
The displacement current, Id , is a quantity with the dimensions of current needed to complete Maxwell’s
equations.
dφe d
Z
Id = ǫ0 = ǫ0 ~ · n̂)dt
(E
dt dt S
A changing electric flux creates a magnetic field.
Ampere’s Law for Electrodynamics: For situations where there is a changing elec-
tric field, Ampere’s Law must be modified as follows,
I
B~ · d~ℓ = µ0 (I + Id ),
C
where B~ is the magnetic field, C is a closed curve, ~ℓ points along the curve, I is
the real current (the moving charged particles) through the surface, and Id is the
displacement current for the surface enclosed by the curve.
A changing electric flux through a surface S produces a net magnetic field around the curve bounding
the surface.
The Displacement Current is Not a Real Current: The displacement current has
the dimensions of current but does not involve the transfer of charged particles
through the surface.
as a function of time.
Solution
The displacement current, Id , is defined as
dφE d N m2 2 N m2
Id = ǫ0 = ǫ0 (100 )t = 2ǫ0 (100 )t
dt dt Cs2 Cs2
A
Id = (1.77 × 10−9 )t
s
+ _
Strategy: Use the parallel plate capacitor formula for the electric field and identify the current as the time
rate of change of the charge on the capacitor plates.
(a) Compute Electric Field in Capacitor: The electric field in a parallel plate capacitor is given by
σ
E=
ǫ0
where the surface charge density σ can be related to the total charge Q by Q = σA where A is the area.
This gives an electric field of
Q
E= .
Aǫ0
In the above expression only Q changes with time, so we can write the derivative of E as
dE 1 dQ
= .
dt Aǫ0 dt
(b) Equate the Current as Time Rate of Charge: The current is charge per unit time so at any instant
the current is
dQ
I=
dt
(c) Substitute and Compute: Substitute the expression for current into the derivative of the electric field,
dE 1 dQ I
= = .
dt Aǫ0 dt Aǫ0
Substitute and solve using the area of the plates as A = πr2 = π(9cm)2 = 0.0254m2
dE I 5A V
= = C2
= 2.22 × 1013
dt Aǫ0 2
(0.0254m )(8.85 × 10−12
Nm2 )
ms
Solution
Strategy: Qualitatively decide on the direction of the displacement current, then use the right hand rule
for a wire to get the direction of the magnetic field.
(a) Sketch Circular Paths: Sketch a number of circular paths
in the region containing the field, these are the boundaries of
the surfaces we will compute the flux through and the Ampe-
rian paths along which we will compute the magnetic fields.
(b) Compute Displacement Current through Surface: In the figure, the field points out of the page. If we
compute the electric flux, φe , with the normal in the direction of the field, then the flux is positive with this
normal. Since the field is increasing, dφe /dt > 0, the displacement current Id is directed out of the page.
(c) Reason About Shape of Magnetic Field: Since magnetic field lines must close, and the region is
cylindrical, the magnetic field lines must be circles.
(d) Use Right Hand Rule (RHR) for Wire: The displacement current acts like a real current in Ampere’s
Law (even though no charge is flowing in the region), so we can use the RHR to compute the direction. Since
the displacement current points out of the page, the B field is oriented as drawn, by the RHR for a wire.
Faraday’s Law :
~ · d~l = − d
I Z
E ~ · n̂)dA
(B
C dt S
Ampere’s Law:
~ · d~l = µ0 I + µ0 ǫ0 d
I Z
B ~ · n̂)dA
(E
C dt S
~ · n̂)dA = Qenclosed
I
(E
S ǫ0
Solution to Part(a)
Gauss’ Law.
Solution to Part(b)
The electric flux through any closed surface is proportional to the net charge enclosed by the surface.
Electromagnetic Waves
385
37.1. A SOLUTION TO MAXWELL’S EQUATION CHAPTER 37. ELECTROMAGNETIC WAVES
~ · d~ℓ = − d
Z Z
E ~ · n̂)dA
(B
C dt S
Now compute the flux through the surface bounded by the path. If the path is thin (∆x is small),
then the magnetic field does not change much over the slice and the magnetic flux can be approximated by
φm = (B ~ · n̂)A. The positive normal for the curve selected is n̂ = ẑ. The area of the surface bounded by the
curve is A = L∆x. The magnetic flux is then
~ · n̂)A = B0 ẑ cos(kx0 − ωt) · ẑ(L∆x) = B0 cos(kx0 − ωt)L∆x
φm = (B
The negative time derivative of the flux is
dφm d
− = − B0 cos(kx0 − ωt)L∆x = −LB0 ∆xω sin(kx0 − ωt)
dt dt
Substitute everything back into Faraday’s law,
~ · d~ℓ = − d
Z Z
E ~ · n̂)dA
(B
C dt S
E0 cos(k(x0 + ∆x) − ωt)L − E0 cos(kx0 − ωt)L = −LB0 ∆xω sin(kx0 − ωt)
Cancel L and divide by ∆x,
cos(k(x0 + ∆x) − ωt) − cos(kx0 − ωt)
E0 = −B0 ω sin(kx0 − ωt)
∆x
Now let ∆x ⇒ 0,
cos(k(x0 + ∆x) − ωt) − cos(kx0 − ωt) d
E0 = E0 cos(kx − ωt) = −E0 k sin(kx − ωt)
∆x dx
−E0 k sin(kx − ωt) = −B0 ω sin(kx0 − ωt)
So the wave equations satisfy Faraday’s law if
E0 k = B0 ω
Maxwell IV - Ampere’s Law Now try the wave equations in
Ampere’s law, z Magnetic Component of Wave
∆x c
d
~ · d~ℓ = µ0 I + µ0 ε0 d
Z Z
B ~ · n̂)dA
(E
C dt S
Let the left side of the path be at x0 from the origin. For b → c, d~ℓ = ẑdz and B
~ · d~ℓ = B0 cos(k(x0 + ∆x) +
ωt)dz. For d → a, d~ℓ = −ẑdz and B ~ · d~ℓ = −B0 cos(kx0 + ωt)dz. Substituting back into the integrals gives,
Z Z Z
~ · d~ℓ =
B B~ · d~ℓ + ~ · d~ℓ
B
C b→c d→a
Z Z Z
~ · d~ℓ =
B B0 cos(k(x0 + ∆x) − ωt)dz + −B0 cos(kx0 − ωt)dz
C b→c d→a
The cosine may be brought out of the integral because it depends on only x and t.
Z Z Z
~ · d~ℓ = B0 cos(k(x0 + ∆x) − ωt)
E dz − B0 cos(kx0 − ωt) dz
C b→c d→a
Now compute the electric flux through the surface bounded by the path. If the path is thin (∆x is
small), then the electric field does not change much over the slice and the electric flux can be approximated
by φe = (E~ · n̂)A. For a positive normal for the curve selected n̂ = −ŷ. The area of the surface bounded by
the curve is A = L∆x. The electric flux is then
~ · n̂)A = E0 ŷ cos(kx0 − ωt) · (−ŷ)(L∆x) = −E0 cos(kx0 − ωt)L∆x
φe = (E
B0 k = µ0 ε0 E0 ω
E0 k = B0 ω
Rearrange the second equation to yield,
ω
E0 = B0 .
k
Substitute back into the first equation,
ω
B0 k = µ0 ε0 B0 ω
k
Cancel B0 and rearrange,
2
1 ω
=
µ0 ε0 k
r
1 ω
=
µ0 ε0 k
ω
The quantity k = λf is the speed of the wave
ω 1
= λf = v = √
k µ0 ε0
where λ is the wavelength, f is the frequency, and v is the wave velocity. Therefore, to be an electromagnetic
wave, the wave must travel at a velocity v = √µ10 ε0
A second result, not as important, but somewhat unusual, is found by substituting back into the second
relation derived from Maxwell equation
ω 1
E0 = B0 = B0 √ = vB0
k µ0 ε0
Therefore, the magnitude of the electric component of an electromagnetic wave is the velocity of the wave
multiplied by the magnitude of the magnetic component.
~ = 0,
or if we use ∇ · B
∂2B ~
~ = ε0 µ0
∇2 B
∂t 2
1 1 m
v=√ =q = 3 × 108
µ0 ε0 (4π × 10−7 Tm
2
−12 C ) s
A )(8.85 × 10 Nm2
Speed of Light in Vacuum: The speed of light in vacuum is denoted by the symbol
c and is always c = 3 × 108 ms .
The velocity of an electromagnetic wave, computed from Maxwell’s equations and the universal constants
ε0 and µ0 is the speed of light, c. Therefore, light is an electromagnetic wave!!!! In this class, you have
measured both ε0 and µ0 and have mastered the techniques to carry out the above calculation, so you can
show this crucial insight about the universe.
~ is defined as
The Poynting Vector : The Poynting vector, S,
~ ~
~ ≡ E×B
S
µ0
Solution
Strategy: Represent the wave as a series of arrows, use the Poynting vector for the direction of propagation.
(a) Think About Real Wave: A plane wave fills space producing an oscillating electric and magnetic field
everywhere. We can’t draw this. What we can draw is the magnitude of the electric and magnetic field
along the coordinate axis in the direction of propagation.
(b) Compute Direction of Propagation: Use the Poynting vector to compute the direction of the other
field component or the direction of propagation if given both E~ and B.
~
(c) Draw the Wave: Draw oscillating electric and magnetic fields as in the figure, making sure that E ~ ×B~
points in the direction of propagation and that the spacing of the crests is the wavelength.
y(E) Direction
z(M) x
(a)What is the magnitude of the magnetic field at t = 0 at the point where the electric field was
measured?
(b)What is the Poynting vector?
(c)What is the direction of propagation?
Solution to Part(a)
The magnitude of the magnetic component of an electromagnetic wave is related to the magnitude of the
electric component by
E 100 N
B= = C
= 3.33 × 10−7 T.
c 3 × 108 ms
Solution to Part(b)
~ ~ N −7
~ = E × B = (100 C x̂) × (3.33 × 10 Tẑ) = 26.5 W (x̂ × ẑ).
S
µ0 4π × 10−7 TmA
m2
The cross product x̂ × ẑ can be evaluated using the Right Hand Rule, x̂ × ẑ = −ŷ.
~ = −26.5 W ŷ
S
m2
Solution to Part(c)
The wave propagates in the direction of the Poynting vector. Therefore, the wave propagates in the −ŷ
direction.
Equation of a Moving Wave: Waves that move in a single direction vary in both
space, x, and time, t. The equation for a wave which varies sinusoidally is
where Amax is the maximum amplitude, k is the wave number, ω is the angular
frequency, and δ is the phase difference. The amplitude of the wave at any specified
point in space and time is h(x, t). Amax is the maximum amplitude of the wave as
measured from its average amplitude (this is always positive).
f = 1/T
t
−Amax
1Hz = 1s−1
Frequency and Angular Frequency: The frequency, f , of a wave has a circular
analog called the angular frequency, ω, defined as
ω = 2πf
The angular frequency is the time rate of change of the argument of the sine function
representing the wave.
k = 2π/λ x
Speed of a Moving Wave: The speed, vwave , of any moving wave is related to its
wavelength, λ, and frequency, f , by
ω
vwave = λf =
k
Solution to Part(a)
Solution to Part(b)
1
h(meters)
-0
-1
-2
-3
-4
-5
0 1 2 3 4 5 6 7 8 9 10
x(meters)
Solution
The period of the wave is given as T = 2s and therefore the frequency is f = 1/T = 0.5Hz and the angular
frequency ω = 2πf = πs−1 . The wavelength, which can be read off the graph, is λ = 4m, which gives a
wavenumber k = 2π/λ = 0.5πm−1 and the wave has a maximum at x = 0, so it is a cosine wave, a sine wave
with phase difference δ = π2 . The amplitude of the wave can be read from the graph.
π 2π
h(x, t) = A sin(kx − ωt + ) = A cos( x − 2πf t) = (5m) cos((0.5πm−1 x) − (πs−1 t))
2 λ
Note how the units cancel in the argument of the cosine function.
U (t) P (t)
I(t) = =
At A
1
~ 2
I(t) = ~ 2 + |B| c
ε0 |E|
2 µ0
√
Using E = cB and c = 1/ ε0 µ0 this can be rewritten,
~ 2
|B| B2
I(t) = c = c 0 sin2 (kx − ωt)
µ0 µ0
Light waves oscillate vary rapidly, so rapidly that we don’t notice the instantaneous changes in intensity, so
it is more useful to know the average intensity, Iave . If we average the intensity over one period, T , we find
T
B02
Z
Iave = I(t)dt = c
0 2µ0
Average Intensity: In general, when we say intensity, we are referring to the average
intensity of an electromagnetic wave. The average intensity is the intensity averaged
over one period of the wave. If we ever mean the instantaneous intensity, we will
explicitly say so.
Intensity of an EM Wave : The average intensity, I, of an EM wave can be related
to the average magnitude of the Poynting vector
~ave |
I = |S
~2
E ~2
cB ~2
cε0 E
0 0 0
I= = =
2µ0 c 2µ0 2
Light waves also carry momentum. For light the total energy in the wave ET OT is related to the total
momentum, p, by p = ET OT /c. The intensity divided by the speed of light is the momentum per unit
area per unit time transferred by the wave. If the light falls on a surface, I/c momentum per unit time is
transferred to the surface per unit surface area. But momentum transfer per time is force, so I/c force per
unit area is exerted on the surface by the light wave. We call force per unit area, pressure.
(a)What is the intensity of the waves at 1km = 1000m from the station?
(b)What is the maximum amplitude of the electric field at 1km from the station?
(c)What is the maximum amplitude of the magnetic field at 1km?
Solution to Part(a)
Intensity, I, is power, P , per unit area; so divide the total power by the surface area of a sphere with radius
r = 1000m.
P 50000W W
I= = = 3.98 × 10−3 2
4πr2 4π(1000m)2 m
Solution to Part(b)
The the intensity in related the magnitude of the electric component of the field by I = 12 cǫ0 E02 , where E0
is the amplitude of the electric field. So for the field,
s
W
r
2I 2(3.98 × 10−3 m 2) N
E0 = = m C 2 ) = 1.73 .
cǫ0 8
(3 × 10 s )(8.85 × 10 −12
Nm2
C
Solution to Part(c)
Align Dipole Antenna for Maximum Reception : The antenna will generate the
maximum signal when the greatest potential difference exists between the two pieces
of the antenna. The maximum potential difference occurs when the axis of the
antenna (as shown above) aligns with the electric field of the wave.
Align Loop Antenna for Maximum Reception : For maximum signal, we need
maximum induced EMF in the loop, thus maximum change in magnetic flux, so
the magnetic field of the electromagnetic wave must align with the normal of the
surface enclosed by the loop.
(a)What kind of antenna would you use to detect the electric field of this wave?
(b)How would you place the antenna for maximum reception of the electric component of the wave?
(c)What kind of antenna would you use to detect the magnetic field of this wave?
(d)How would you place the antenna for maximum reception of the magnetic component of the
wave?
Solution to Part(a)
Since the electric field causes a force (and therefore a potential difference) in the direction of the field, a
dipole antennae is best.
Solution to Part(b)
The antenna’s axis should be parallel to the electric field, in the ŷ direction.
Solution to Part(c)
Since the electric potential difference here is caused by a changing flux through an area, a loop works best.
Solution to Part(d)
To maximize the changing flux, you want the normal of the loop along the magnetic field in the ẑ direction.
Therefore the loop should lie in the x − y plane.
Solution
Strategy: Compute the flux as a function of time, then apply Faraday’s Law to get the EMF.
(b) Compute Magnetic Flux through Loop: Compute the flux, φm (t), as a function of time assuming the
~
wave changes little over the area of the loop, so φm = A|B(t)|, ~
where A is the area of the loop and |B(t)| is
2 2
the magnitude of the field. The area of the loop is A = πr = 0.13m . The flux is then
~ = 2.6 × 10−10 Wb sin((12.6s−1 )t)
φm = A|B(t)|
37.5 Polarization
Light is an electromagnetic wave and is described by the propagation of oscillating electric and magnetic
fields. The polarization of a wave is usually expressed by the direction of the oscillating electric field vector; if
the electric field oscillates in a plane the wave is said to be linearly polarized. Light can easily be a collection
of electromagnetic waves with random polarizations. Such light is unpolarized.
37.5.1 Polarization
For most light, the direction of the electric vectors is random, changing with time and from place to
place in the wave. By passing the light wave through a polarizer, we can arrange for the electric vector to
oscillate in a single direction like the x̂ direction. The simplest polarizer is regular array of conducting wires,
as shown below. If a light wave with random electric vector is incident on the polarizer as in figure (a), the
component of the electric field parallel to the wire, E ~ k causes charge to flow and loses energy. Therefore,
after passing through the polarizer as in figure (a), only the component of the electric field perpendicular to
the wire survives, E ~ ⊥.
Figure(a) Field Before Encountering Polarizer Figure(b) Field After Encountering Polarizer
Eperp
E|| E0
θ Eperp
transmission axis
wires wires
We can use the above picture to determine how the intensity of the light wave is affected as it passes
through the polarizer. The magnitude of the electric field changes from E0 to E0 cos θ, where θ is the
angle between the incident field and the transmission axis. Since the magnetic field is proportional to the
magnetic field and both the magnetic and electric energy density depends on the square of the field, the
intensity changes from I0 to I0 cos2 θ. If unpolarized light is shined on a polarizer, polarized light of half the
initial intensity is produced.
Malus’ Law: Consider polarized
light of a certain intensity, Ii , in-
incident polarization axis
cident at a certain angle, β, to
the transmission axis of a polar-
β
izer. The intensity of light trans- Ii
mitted through the polarizer It is
given by Malus’ Law
It = Ii cos2 β
transmission
It
Unpolarized Light Halved by Polarizer: With a bit of calculus, we find that the
transmitted intensity for initially unpolarized light is half that of the incoming light.
Solution to Part(a)
If the middle sheet makes an angle θ1 = 15◦ with the first sheet, it makes an angle θ2 = 90◦ − θ1 = 75◦ with
the second sheet since the two outside polarizers must be at an angle of 90◦ with each other. A polarizer
reduces the intensity of unpolarized light by a factor of 2, so if I0 is the intensity of the light shining on
the first, then I1 = I0 /2 reaches the middle polarizer. The middle polarizer transmits light of an intensity
I2 = I1 cos2 (θ1 ). The third polarizer transmits I3 = I2 cos2 (θ2 ). So in terms of the incident intensity
Solution to Part(b)
The answer is the same as above since 75◦ + 15◦ = 90◦ . This just shines the light through the polarizer state
in (a) from the other side.
Polarization by reflection Any reflected light is partially polarized to an axis parallel to the reflection
interface. However, at a particular angle of incidence, called the polarization angle θp , the reflected
light is totally polarized. We will discuss this in Course Guide 38.
Polarization by scattering Light can be absorbed by a material, then re-radiated. This is called
scattering. Some of the scattered light travels in a direction perpendicular to the incident direction.
If the incident light is polarized, the reradiated light is polarized in the same direction. If the incident
light is unpolarized, the reradiated light which is perpendicular to the incident light has two directions
along which it is perfectly polarized. Air molecules absorb and re-radiate blue light better than red
light, which has interesting effects as we will see.
Polarization by birefringence Some materials will split a single beam of incoming light into two beams
travelling at different speeds with mutually perpendicular polarizations.
Light
Speed of Light in Material: The speed of light in a material, cmatter with dielectric
constant κ and relative magnetic permeability Km is
1
cmatter = √
κε0 Km µ0
The magnetic permeability can be very slightly less than 1 but is usually bigger than 1. The dielectric
constant is always greater than 1. The product κKm is always greater than one, so the speed of light in a
material is less than the speed of light in vacuum.
Light Slows Down in a Material: The speed of light in a material is lower than
the speed of light in vacuum.
The amount light slows down in the material is the only property of the material we need to do optics.
It is convenient to report the amount light slows as the ratio of the speed of light in a vacuum to the speed
of light in a material. This ratio is given the name the index of refraction of the material and is given the
symbol n.
Speed of Light in a Material: When light travels in a material, its speed (cn ) is
less than the speed of light in a vacuum. The ratio of these speeds is called the
index of refraction (n)
c p
n≡ = κKm
cn
Sizes of the Index of Refraction: The index of refraction has no units since it is a
ratio of velocities. The index of refraction in a vacuum is defined to be 1. The index
of refraction of air is very close to 1. Water has an index of refraction of 4/3 = 1.33.
The table below lists the index of refraction for some common materials at a few wavelengths.
402
38.1. LIGHT AND MATTER CHAPTER 38. LIGHT
Solution to Part(a)
The speed of light in air is approximately the speed of light in a vacuum, cair ≈ c = 3 × 108 ms .
Solution to Part(b)
The wavelength is given in the problem, you just had to be able to correctly interpret the phrase “a 500nm
laser”. λair = 500nm.
Solution to Part(c)
cair 3 × 108 ms
fair = = = 6 × 1014 s−1
λair 5 × 10−7 m
Solution to Part(d)
c 3 × 108 ms m
cLucite = = = 2 × 108
nLucite 1.5 s
Solution to Part(e)
The frequency of the light does not change when it enters a material, fLucite = fair = 6 × 1014 s−1 .
Solution to Part(f)
The wavelength of the laser light in the Lucite(TM) can be found using c = λf ,
cLucite 2 × 108 ms
λLucite = = = 333nm
fLucite 6 × 1014 s−1
or
λ 5 × 10−7 m
λLucite = = = 333nm
nLucite 1.5
that we will consider over the next few chapters, the wave character of light is hidden and light can be taken
to be a geometric ray travelling in a straight line.
When light strikes a surface, the electric and magnetic components are altered in the material as described
earlier. All of Maxwell’s equations must be satisfied at the surface, so Gauss’ law must be satisfied for a
pillbox enclosing the surface and Ampere’s law must apply to an Amperian path encircling the surface. To
satisfy all of Maxwell’s equation, a reflected wave that bounces off the surface and a refracted wave that
enters the material, but travels at a different angle are required. All three, the incident, the reflected, and
the refracted ray have the same frequency.
θi = θr
That is to say, light reflects from a point on the interface at the same angle it
arrived.
Angles and the Normal to the Interface: An interface is the boundary between
two materials. At any point on the interface, a line can be imagined which is
perpendicular (normal) to the interface. All angles in geometric optics are measured
with respect to the normal.
The incident and reflected rays are drawn to the right. When
you look into a mirror, the light appears to come from the line
traced back from the reflected ray. This direction is drawn as
a dashed line. All angles are measured from the normal.
incident
θi
θr
reflected
mirror
Law of Refraction (Snell’s Law): If the light travels from a material with index of
refraction, ni and is transmitted across an interface into another material with index
of refraction, nt , and the angle of incidence at the interface between the material
is, θi , to the normal then the angle of transmission θt is related by Snell’s Law
ni sin θi = nt sin θt
If you play with Snell’s law a bit, you find that as light travels from a lower index of refraction to a higher
index of refraction, the light is bent closer to the surface normal. As light travels from a higher to lower
index of refraction, the light bends farther away from the surface normal. Both cases are drawn below.
Light Travelling from Low Index to High Light Travelling from High Index to Low
index, bends toward the normal. index, bends away from the normal.
ni nt ni < nt ni nt ni > nt
θi θi
θt
θt
incident transmitted
incident transmitted
Total Internal Reflection : When light goes from a high index to a lower index
material, it is possible that the transmitted light will not leave the interface bound-
ary. The minimum incident angle for which this occurs is called the critical angle.
The critical angle is reached when θt = 90◦ . This means Snell’s Law becomes
ni sin θc = nt and solving for the angle gives
nt
θc = arcsin
ni
The light will be totally internally reflected for θi ≥ θc .
ni nt ni > nt
θc
θt = 90o
(a) 3 × 108 ms (b) 2.1 × 108 ms . (c-Answer) 1.6 × 108 ms . (d) Cannot be determined from the
information given. (e) Toby
Solution
Let the index of refraction of the material we’re seeking be n1 and the water be n2 . The critical angle is the
angle of incidence required to produce an angle of refraction of 90◦ .
4/3
n1 = = 1.82
sin 47◦
The speed of light in the material, c1 , is then
c 3 × 108 ms m
c1 = = = 1.6 × 108 .
n1 1.82 s
Let the first material in the problem be material 1 with light speed c1 = 1.5 × 108 ms and the second material
be material 2 with light speed c2 = 2.5 × 108 ms . The index of refraction of material 1 is by definition,
c 3 × 108 ms
n1 = = =2
c1 1.5 × 108 ms
Solution to Part(b)
The index of refraction of material 2 is by definition
c 3 × 108 ms
n2 = = = 1.2
c2 2.5 × 108 ms
Solution to Part(c)
The wavelength in a material is reduced by a factor of the index of refraction from the wavelength in a
vacuum. If λ is the wavelength in a vacuum, then λ1 = λ/n1 . Solving for
λ = λ1 n1 = (320nm)(2) = 640nm = 640 × 10−9 m
Solution to Part(d)
The wavelength in a material is reduced by a factor of the index of refraction of the material, λ2 = λ/n2 =
533.3 × 10−9 m.
Solution to Part(e)
Yes, since the first index of refraction is higher than the second, light travelling from a higher index of
refraction to a lower index can undergo total internal reflection.
Solution to Part(f)
Applying Snell’s Law with a transmitted angle of 90◦ gives
n1 sin θc = n2 sin 90◦
Solve for θc ,
n2
θc = sin−1 = 36.9◦
n1
Since 55◦ > 36.9◦ the light will be totally internally reflected.
Solution to Part(g)
Definition of Intensity: The intensity, I(t), of a wave is the energy, U (t), crossing
a unit area, A, per unit time. This is the same as power, P (t), per unit area,
U (t) P (t)
I(t) = =
At A
The intensity of the reflected and transmitted light rays can be calculated from Maxwell’s equations. The
reflected intensity changes in a complicated manner with the angle of incidence. The reflected intensity also
depends on the direction of the electric component of the wave. The amount of light reflected is different if
the electric component lies in the plane formed by the direction of propagation and the surface normal or is
perpendicular to the plane.
Electric Field Perpendicular to Plane of Incidence Electric Field Parallel to Plane of Incidence
T
1.0 T 1.0
0.5 0.5
R
R
θp
0.0 0.0
30 60 90 30 60 90
Angle of Incidence (degrees) Angle of Incidence (degrees)
The are two special cases where the same result is obtained for any direction of the electric component
of the field, normal incidence where the angle of incidence is zero and grazing incidence where the angle
of incidence is 90◦ . If the light ray strikes a surface at normal incidence, perpendicular to the surface, the
reflected intensity Ir has a simple relation to the indices of refraction of the two materials. Since energy
is conserved, the transmitted intensity, It , is the difference between the incident intensity and the reflected
intensity. We will only deal with normal incidence, because the expression becomes much more complicated
away from normal incidence.
Intensity of Reflected Light : In the special case that the incident light strikes the
interface at right angles (normal incidence), the intensity of the reflected light, Ir ,
is a fraction of the intensity of the incident light, Ii .
2
ni − nt
Ir = Ii
ni + nt
Energy is Conserved: If there is no absorption, then the reflected intensity and the
transmitted intensity must add up to the total intensity.
Ii = Ir + It
Grazing Incidence: If the angle of incidence is 90◦ , that is if the light just grazes
the surface, then all the light is reflected and no light is transmitted.
is stacked on top of a sheet of plain old glass with nglass = 1.4. To the first order, which means we ignore
the light which undergoes multiple reflections, compute the intensity of the transmitted light.
Solution
Definitions
Strategy: Apply the formula for the reflected intensity of normally incident light at each interface.
(a) Compute Transmitted Intensity at Air-Lucite Interface: The reflected intensity at the first interface
is 2
nair − nLucite
Iref lected = I0
nair + nLucite
2
1 − 1.5
Iref lected = I0 = 0.04I0
1 + 1.5
Since energy is conserved, the transmitted intensity at the Air-Lucite interface, ILucite , is
(b) Compute the Transmitted Intensity at the Lucite-Glass Interface: The reflected intensity at the
second interface is 2
nLucite − nglass
Iref lected,glass = ILucite
nLucite + nglass
2
1.5 − 1.4
Iref lected,glass = ILucite = 0.001ILucite
1.4 + 1.5
So the transmitted intensity through the Lucite-Glass interface,
(c) Compute the Transmitted Intensity at the Glass-Air Interface: The reflected intensity at the third
interface is 2
nglass − nair
Iref lected,air = Iglass
nglass + nair
2
1.4 − 1
Iref lected,air = Iglass = 0.028Iglass
1.4 + 1
Therefore, the transmitted intensity is
(d) Multiply It All Out: Combine the results of the three previous sections
W W
It = (0.96)(0.999)(0.972)I0 = (0.932)(1000 2
) = 932 2
m m
Solution
The light goes from ni = 1 to nt = 1.33, so the angle of polarization, θp , is given by Brewster’s law,
nt 1.33
θp = arctan = arctan = 53◦
ni 1
Definitions
n mostly
polarized m
vn = 1.5 × 108 ≡ Velocity of Light in Material
s
n=1 θi ≡ Incident Angle
θi θr θt ≡ Transmitted Angle
1 θr ≡ Reflected Angle
θp ≡ Brewster’s Angle
Material n ≡ Index of Refraction of Material
θt
Strategy: Use Snell’s Law to compute the angle of refraction, and Brewster’s Law the angle of polarization.
Compute the Index of Refraction: The Index of Refraction is the ratio of the velocity of light in the
material to the velocity of light in vacuum
c 3 × 108 ms
n= = =2
vn 1.5 × 108 ms
Compute Angle of Incidence: Use Snell’s Law to relate the angle of incidence to the angle of refrac-
tion(transmission).
nair sin θi = n sin θt
Using nair = 1 and n = 2, gives
sin θi = 2 sin 26◦
Solving for θi and
θi = arcsin(2 sin 26◦ ) = 61◦
θp = arctan(n) = 63◦ .
Compute Angle of Reflection: The Angle of Reflection equals the Angle of Incidence
θr = θi = 61◦
The reflected beam is close to Brewster’s Angle, so the reflected light is almost fully polarized. The direction
of polarization is perpendicular to the plane of incidence.
Geometric Optics
Light is an electromagnetic wave, but also a flux of photons. To learn more about the wave nature of light
take UPIII. To learn more about photons, take Modern Physics. We can understand the behavior of systems
composed of mirrors and lenses by modelling light as a ray. The study of light in situations where it behaves
as a ray is called geometric optics. Geometric optics is a good approximation as long as the wavelength of
light is small compared to the size of optical system. Visible light has a wavelength on the order of 500nm,
therefore systems composed of lenses and mirrors are well approximated by geometric optics. Chromatic
abberations, polarization, and interference effects cannot be explained by the geometric optics model. For
these, we need a model of light that recognizes its wave nature, that is to say, wave (physical) optics. For
some effects in nature, even the wave model is not adequate and we must take into account the corpuscular
nature of light and then we have a model called quantum optics.
diffuse
reflection
object object
Specular Reflection from a smooth surface
After bouncing off the object, the light will travel in straight lines called light rays. So if we place a screen
414
39.1. IMAGE FORMATION CHAPTER 39. GEOMETRIC OPTICS
somewhere in the path of these light rays without using some optical element to sort out the various rays,
each point on the screen will receive light rays from all points on the object and all that we’ll see is a general
impression of the color of the object.
screen object
If we place an optical system between the screen and the object—we can form an image on the screen, where
all light rays from the same point on the object strike the same point on the screen. When the screen is
placed where the light rays from a single point converge (where a sharp image is formed), we say that the
image is focused. The dashed lines drawn in the optical system do not represent the light rays. They allow
you to associate incoming and outgoing rays. The path through the optical system may be very complex.
image formed
optical system light source
screen object
The image will be fuzzy if the screen is moved closer to or farther away from the optical system, since the
light rays do not meet at the same point on the screen. In this case, the image is said to be out of focus.
screen object
Definition of Image: An image is a surface in space where all light rays coming
from the same point on an object intersect.
The light rays continue in a straight line after passing through the surface where the image forms. They
may eventually be detected by an optical detector like the human eye. The detector sees the rays originate
at the last point they all crossed, at the image. The detector only measures the direction the light rays are
travelling when they encounter the detector, their apparent direction. This direction may be very different
than the direction they were travelling when they left the object.
Apparent Direction: No
matter what (possibly apparent source
convoluted) path that
light may take from a
source to a detector, the
apparent source direc-
tion is the direction from light path
which the light enters the
detector. The line of
sight is along the appar-
ent source direction (the
dotted line in the figure
to the right).
detector
source
Object
Image
The image formed by an optical system is described by three features: whether it is bigger or smaller
than the object, whether it is right side up or upside down compared to the object, and whether it is formed
of physical light rays or the apparent direction of light rays. The first two of these features are captured in
a single number, the magnification, m. The magnification is positive if the image has the same orientation
as the object and negative if the image is upside down. The magnitude of the magnification is the ratio of
the image height to the object height. Therefore, if the image is twice as large as the object and right side
up, then the magnification is 2. If the image is one third the height of the object and upside down, then the
magnification is − 13 .
Magnification and Inversion : If the image is a different size than the object, then
it is magnified (reduced or enlarged, depending on the relative size). If the image
is oriented in the same direction as the object, then the image is upright. If the
image is oriented in the opposite direction, it is inverted. This classification can be
represented by a number, m, called the magnification of the image. The table below
consolidates the meanings of various values of Magnification for the black arrow as
the original object (images are in outline)
Real Image Formation : Consider light that is reflected from a single point on an
object. The reflected light can take different paths to a detector. If the light paths
themselves intersect, then a real image is formed at that intersection point. An
example of real image formation is shown below for a concave mirror.
object
image
Virtual Image Formation : Consider light that is reflected from a single point on
an object. The light can take different paths to a detector. If the light paths do not
intersect, but the apparent source directions do, then a virtual image is formed. An
example of virtual image formation is shown below for a diverging lens. In this case
the light travels through the optical element and so the detector is to the right.
object image
tools are the definition of image, the law of reflection, and Snell’s law. Using the definition of an image we
can find the location of the image by:
Locating the Image: To locate the image, follow two light rays emerging from the
same point on the object. The rays will intersect or appear to intersect at the image.
So we need to select two light rays. The ray that leaves the base of the object and strikes the mirror with
normal incidence is the easiest to trace since it is reflected directly back on itself. Select a second ray that
leaves the base of the object making an angle θ to the first ray. This ray will strike the mirror with angle
of incidence θ and be reflected at an angle θ as drawn below. The rays leaving the mirror never intersect so
a real image is not formed. The apparent direction of the outgoing rays cross behind the mirror. A virtual
image is located at the point they cross. By similar triangles, the image is the same distance behind the
mirror that the object is in front of the mirror.
image
θ θ
θ
normal
θ
mirror mirror
Apparent Direction
This analysis tells us nothing about the size and orientation of the image. To find this information, we
have to trace rays from two separate points on the object.
Describing the Image: To find the size and orientation of the image after the
location is found, trace one ray from the top of the object and one ray at the
bottom. These rays will pass through the top and the bottom of the image at the
image location.
We have already traced two rays from the base of the object
through the system, so we know where the base of the image
is. Now, trace a ray from the top of the object. The easiest ray
is the ray that strikes the mirror at normal incidence. This ray
is drawn to the right. The apparent direction of this ray must
pass through the top of the image. I have drawn the image in. object image
The image is the upright and the same height as the object, so
the magnification is m = 1.
θ
normal
θ
mirror
Image Formed by Flat Mirror: The image formed by a flat mirror is virtual, behind
the mirror, upright and the same size as the object (m = 1).
object image
image
θi θi
θi θi
normal normal
θt θt
surface surface
To actually solve for the image location quantitatively, we would have the solve Snell’s law for θt , then
express θi in terms of the object location. This is bad enough from a flat interface, but becomes intensely
terrible for a curved interface. To simplify matters, we will make an assumption that is approximately true
for almost any optical system of interest.
Paraxial Rays: Paraxial rays, nearly parallel rays, are rays the meet the interfaces
of an optical system at sufficiently small angles that the small angle approximation
is valid: sin θ ≈ θ, tan θ ≈ θ, and cos θ ≈ 1. We will use this approximation from
now on when analyzing optical systems.
Since most of our optical systems are not drawn to scale, most of our drawings will not look like this is
a good hypothesis, but I have to get the drawing on a sheet of paper.
Assuming the rays passing through the system are paraxial,
Snell’s law simplifies to ni θi = nt θt . I have redrawn the sys- Calculating the Image Location
tem to the right with s as the distance the object is from the
surface, s′ the distance the image is from the surface, and h surface
water air
the height the ray intersects the interface. Since the angles are nwater
small, we can write tan θi = h/s ≈ θi and tan θt = h/s′ ≈ θt . nair
Substituting into Snell’s law gives, object
h h
ni θi = ni = nt θt = nt ′
s s
where ni = nwater and nt = nair . Solving for the image dis- θi θt t
tance gives, s = (nair /nwater )s = 0.75s. So for water, the h
θi normal
fish is appears to be only 34 of the distance from the aquarium
θt
surface as it actually is. This is what makes grabbing things
underwater dicey. s
s’
object
axis
C
normal
mirror
To locate the image, we take two rays from the same point on the object and see where they go. We’ll use
two rays from the base of the object. The ray that strikes the mirror at normal incidence bounces directly
back on itself. A ray that leaves the object at an angle θo strikes the mirror at an angle of incidence of α.
It reflects from the mirror with an angle of reflection α. This reflected ray intersects the ray with normal
incidence at the image and makes an angle θi to the axis. The angle the normal makes with the axis, θc ,
will also be important.
object
image C axis
θc θi θo
normal
α h
α
A
s
∆s
s’
Using a little geometry and a lot of small angle approximation, we can find the mathematical location of
the image. First, another approximation that will save us a lot of work. The distance ∆s, the distance the
curvature of the mirror moves the point the second ray intersects the surface from the vertex, will cause us
no end of annoyance. Assume this distance is small compared to s and s′ .
Thin Lens/Mirror Approximation - Part I: The distance the lens’ or mirror’s surface
curves away from the vertex is small and can be ignored in all calculations.
Symbols for Optical Systems: We will use the following symbol convention. The
distance the object is from the mirror or lens will be called the object distance and
denoted by the symbol s. The distance the image is from the mirror or lens will be
called the image distance and denoted by the symbol s′ . The distance the center
of curvature is from the lens or mirror will be called the radius of curvature and
denoted by r.
Let point A be the point where the second ray strikes the mirror and h be the distance point A is from
the axis as drawn above. Using the small angle approximation we can write
h h h
tan θo = ≈ θo tan θi = ′ ≈ θi tan θc = ≈ θc
s s r
◦
There must be 180 in a triangle. For the triangle (image-C-A), this means
θi + (180◦ − θc ) + α = 180◦ → θi + α − θc = 0
and for the triangle (C-object-A)
θc + (180◦ − θo ) + α = 180◦ → θc + α − θo = 0
Use these two equations to eliminate α,
h h h 1 1 2
θo + θi = 2θc ⇒ + ′ =2 ⇒ + =
s s r s s′ r
Thin Mirror Equation: The location of the image of a slightly curved mirror is
found using the equation
1 1 2
+ ′ =
s s r
Since we have already traced the rays from the base of the object to the image, we can find the height and
orientation of the image by tracing a ray from its point. The ray that travels from the point the vertex has
incident angle β is reflected at equal angle β. The image is upside down, so the magnification is negative. The
ratio of the height of the image to the height of the object is given by similar triangles m = −hi /h0 = −s′ /s.
object
ho
image β axis
β
hi
s
s’
ni surface
water air nt
α A
β
object
h
image θi θo θc axis
apparent
direction C
h
s
∆s
s’
r
Locate the Image
This time we will use the fact that the sum on the angles in a straight line is 180◦ . Using the line formed
by the normal, we can write
and using the line formed by the outgoing ray and the apparent direction
Substituting the angles expressed in terms of the image and object distances gives
h h h
ni − nt ′ = (nt − ni )
s s r
and eliminating h,
ni nt nt − ni
− ′ =
s s r
We will adopt a sign convention next section and in that convention s′ will be negative for this situation, so
the formula for a single interface will be
ni nt nt − ni
+ ′ =
s s r
We can find the magnification by tracing the ray from the point of the object through the vertex. If the
incident angle is α and the transmitted angle is β, then the two angles are related by Snell’s law, ni α = nt β.
The object height is ho = αs and the image height hi = βs′ . The image is upright so the magnification is
positive. The magnification is
hi βs′ ni s′
m= = =
ho αs nt s
where I used β = (ni /nt )α from Snell’s law. Once again if we use a sign convention where s′ is negative for
′
this situation this will become m = − nnitss .
apparent
direction
surface nt
air
A
hi
object
ho
α axis
image β
water ni
s’
The image is virtual because it is formed of the apparent direction of the outgoing rays. The image is
upright and enlarged.
Sign Convention for Optics: The positive direction of the optical axis is in the
direction that light rays leave the optical element. The object distance to the first
interface is always positive.
Single Interface Equation (Thick Lens Equation): Consider now a spherical re-
fractive interface with a certain radius of curvature, r. If an object is in a material
of a certain index of refraction, ni , the light will impinge upon the interface to
another material, nt , and an image will be formed. The distance of the object from
the vertex of the interface, s, is related to the distance of the image, s′ , from the
same via the interface equation
ni nt nt − ni
+ =
s s r
ni s′
m=−
nt s
We worked out the location of the image for a flat surface earlier. Our new universal equation should
give the same answer back.
(a)How high is the insect above the surface from the frog’s viewpoint?
(b)Does the frog see the insect magnified, reduced, or the same size as if both were in air?
Solution to Part(a)
The frog is looking at the insect, so the light is going from the insect to the frog. Thus, ni = 1 and nt = 4/3.
The radius of curvature of the surface is infinity, since you have to have an infinitely big sphere for part of
its surface to appear flat. The single-interface equation is
ni nt nt − ni
+ ′ =
s s r
and reduces to
1 4/3
+ ′ = 0,
3cm s
since 1/∞ = 0. Solving the equation yields the image distance as s′ = −4cm, so the insect appears to be
4cm above the surface of the water (incident side).
Solution to Part(b)
ni s′
m=− = +1,
nt s
which will always be the case for a flat surface.
Solution to Part(a)
Light travels from the fish to the observer’s eye; therefore, the optical axis is as drawn. The index of refraction
of the incident material, water, is ni = 1.33 = 4/3. The index of refraction of the transmitted material, air,
is nt = 1. The mask bulges outward as drawn giving a positive radius of curvature once the sign convention
is applied r = +0.5m. The object, the fish, is located at s = 1.25m. The image distance is found using the
Single Interface Equation:
ni nt nt − ni
+ ′ = .
s s r
Substitute.
4/3 1 1 − 4/3
+ ′ =
1.25m s 0.5m
Solving yields an image distance of
s′ = −0.577m.
Solution to Part(b)
Solution to Part(c)
image fish
r 0 s' s
If we have multiple curved surfaces, the first surface light reaches forms an image, called an intermediate
image. This image becomes the object for the second surface. The total magnification of a multiple interface
system is just the product of the magnifications of each surface.
Magnification of Multiple Elements: The magnification, m, of an optical system
containing multiple elements is found by multiplying the magnification of each in-
dividual element.
m = m1 m2 ...mN
water
glass
nw ng
final image
a b
object
intermediate image
(a) Draw the Optical System: Drawn the real optical axis and place the optical elements on the axis.
Draw an optical coordinate axis for each interface.
(b) Compute the Image Distance as Formed by Interface a: Use Interface equation for the first interface
ni nt nt − ni nw ng ng − nw
+ ′ = −→ + ′ =
s s r sa sa ra
and solve for the intermediate image distance s′a . From the diagram the radius of the first interface (a) is
positive. Solve the interface equation for s′a
−1 −1
ng − nw nw 3 3/2 − 4/3 4/3
s′a = ng − ′ = − = −72cm
ra sa 2 24cm 48cm
The intermediate image forms farther from the first interface than the object, as drawn.
(c) Calculate the Magnification of Interface (a): The magnification for a single interface is
ni s′
m=−
nt s
where ni is the index of refraction where the light originates and nt is the index of refraction where light is
transmitted.
4
nw s′a (−72cm)
ma = − = − 33 = 1.33
ng sa 2 (48cm)
(d) Calculate the Object Distance for Interface (b): The object for the second interface is the image
formed by the first interface. The distance between the optical interfaces must be accounted for, which
means
sb = d − s′a
The image from the first interface is a virtual image.
(e) Calculate the Image Distance as Formed by Interface (b): Use Interface equation for the second
interface
ni nt nt − ni ng nw nw − ng
+ ′ = −→ + ′ =
s s r sb sb rb
and solve for the intermediate image distance s′b . Note from its placement on the optical coordinate axis,
rb < 0.
Special Note: Since the light comes through ng first, that is the index that is associated with the object
distance for the second interface. Solve the interface equation for s′b
−1 −1
′ nw − ng ng 4 4/3 − 3/2 3/2
s = nw − = − = 159cm = 1.6m
rb sb 3 −12cm 272cm
(f) Calculate the Magnification of Interface (b): The magnification for a single interface is
ni s′
m=−
nt s
where ni is the index of refraction where the light originates, this time the glass, and nt is the index of
refraction where light is transmitted, this time the water.
3
ng s′b (160cm)
ma = − = − 24 = −0.66
nw sb 3 (272cm)
(g) Calculate the Total Magnification: The total magnification is the product the magnifications of the
individual elements,
mT = m1 m2 = (1.33)(−0.66) = −0.88
The final image is inverted (mT < 0), reduced (|mT | < 1), and real s′b > 0.
n0
surface 1 surface 2 n0
object
image surface 1 n
C2 C1
surface 1 axis
s s1’ 0 r1
r2 s2 0 s’
surface 2 axis
If we add the two equations, the terms containing the intermediate image distance cancel out and we get
n0 n0 n − n0 n0 − n 1 1
+ ′ = + = (n − n0 ) −
s s r1 r2 r1 r2
The total magnification of the lens is the product of the magnifications of the individual surfaces,
n0 s′1 ns′ s′
m = m 1 m2 = − − ′ =−
ns n0 (d − s1 ) s
if d is small.
Image Location of a Thin Lens: The image formed by a thin lens is located by
1 1 (n − n0 ) 1 1
+ ′ = −
s s n0 r1 r2
where n is the index of refraction of the lens and n0 is the index of the material
outside of the lens. The magnification of the thin lens is m = −s′ /s.
Thin Lens Approximation Part II: A thin lens is a lens where the distance between
the curved surfaces is small compared to the intermediate image distance of the
first surface.
Types of Mirrors:
432
40.1. IMAGE FORMATION USING MIRRORS CHAPTER 40. THIN LENSES AND MIRRORS
Therefore, light rays from the sun are very, very nearly parallel. We can also make parallel light rays
using an appropriately constructed optical system. Parallel light rays are very useful. If we shine parallel
light rays on a curved mirror, the light rays, or the direction the light rays appear to come from, will converge
at a point. This point is called the focal point.
Definition of Focal Point: The point where incoming parallel light rays cross or
the apparent direction where incoming parallel light rays appear to cross. Note this
is the secondary focal point for a lens, the primary focal point is the point where a
point light source could be placed to generate parallel rays.
For the concave mirror system below, the parallel light rays all reflect to a point. For the convex mirror
system, the reflected light rays appear to originate from a point behind the mirror. The direction the rays
appear to come from is the apparent direction. The apparent direction is shown as dashed lines on the
diagram. The point where the light or the light’s apparent direction converges is called the focal point of the
optical system, and is labelled F .
apparent
direction
F F
where r is the radius of curvature. To find the focal length, we need to find where parallel rays are focused.
To produce parallel rays, use an object that is infinity far from the mirror. The image of this object will be
at the focal point, so s = ∞ and s′ = f .
1 1 2
+ =
∞ f r
so the focal length of a mirror is f = r/2.
Mirror Maker’s Equation: The focal length, f , of a spherical mirror is half the
radius of curvature, r, of the mirror,
r
f= .
2
Sign Convention for Optics: The positive direction of the optical axis is in the
direction that light rays leave the optical element. The object distance to the first
interface is always positive.
For mirrors, if the incident light comes in from the left, it reflects off the mirror and leaves to the left. For
lenses the light passes through the element, so if incident light is from the left, the light leaves the system
to the right.
The Optical Axis: The optical axis of an optical system is a line that passes through
each optical element perpendicular to the surface of the element. We will also use
an optical coordinate axis drawn below the optical system to show the locations of
the variables in our optical calculations.
Convex Mirror
incident light
reflected light
Thin Lenses and Mirrors: The mirrors and lenses we analyze are called thin lenses
and mirrors. This means the thickness or the lens or radius of curvature of the
mirror is such that the light rays appear to refract or reflect off a plane through
the center of the lens or mirror. An exaggerated cartoon of the lens or mirror will
usually be drawn, but along with it will be a plane that defines the mathematical
location of the thin lens or mirror.
The optical coordinate axis clearly shows the direction, which the sign convention gives as positive, and
the location of important distances in the optical system: radii of curvature, focal lengths, the object, and
image distances. The origin of the optical axis is the location where the optical axis intersects the mirror.
Both the real optical axis and the optical coordinate axis will be referred to as the optical axis.
The behavior of a spherical mirror as an optical element is determined by two points that lie on the
real optical axis, the center of curvature, C, and the focal point, F . With these points are associated two
distances along the optical coordinate axis, the radius of curvature, r, and the focal length f . Both these
distances have signs given by our sign convention. They are positive if they are on the positive side of the
optical coordinate axis (from zero) and negative if they are on the negative side.
A couple of examples to illustrate these definitions: first a convex mirror, then a concave mirror. In both
cases, we have to illustrate the direction of the incident light. We will save plane mirrors until we can
mathematically locate images, because both the focal point and the center of curvature of a plane mirror
are at infinity.
(a)Draw the mirror, the optical axis, the location of the focal point, and the location of the center
of curvature.
(b)Draw the optical coordinate axis, clearly labelling positive and negative on the axis. Then mark
the radius of curvature, r, and the focal point f on the optical axis.
(c)Compute the signed value of the focal length and the radius of curvature.
Draw the optical axis through the center of the mirror and a representation of the lens that bends in the
correct direction. The curvature of a convex mirror is as shown below. The center of curvature, C, for this
mirror is as drawn. For all mirrors, the focal point, F , is halfway between the center of curvature and the
mirror as drawn.
Convex Mirror
F - Focal Point
C - Center of Curvature
incident light
optical axis
F C
Draw the Optical Coordinate Axis and Locate Landmarks: The optical coordinate axis is drawn below the
mirror in the figure above. The sign convention tells us that positive on the optical axis is on the reflected
side of the mirror. Since light bounces off of a mirror, the reflected side is the same as the incident side, so
positive and negative on the optical axis are as drawn. The points f and r, the focal length and radius of
curvature, are located directly below the focal point and center of curvature.
Compute Focal Length and Radius of Curvature: Since the focal point and center of curvature are on the
negative part of the optical axis, the signed radius of curvature is r = −20cm and the signed focal length is
f = r/2 = −10cm.
(a)Draw the mirror, the optical axis, the location of the focal point, and location of the center of
curvature.
(b)Draw the optical coordinate axis, clearly labelling positive and negative on the axis, then mark
the radius of curvature, r, and the focal point f on the optical axis.
(c)Compute the signed value of the focal length and the radius of curvature.
Locate the Center of Curvature and the Focal Point: Draw the real optical axis through the center of
the mirror. Draw a cartoon (numerical value of radius not maintained) of the mirror that curves in the right
direction. The curvature of a concave mirror is as shown below. The center of curvature, C, for this mirror
is as drawn, and for all mirrors, the focal point, F , is halfway between the center of curvature and the mirror
as drawn.
Concave Mirror
incident light
C F
Draw the Coordinate Optical Axis and Locate Landmarks: The optical coordinate axis is drawn below
the mirror. The sign convention tells us that positive on the optical axis is on the reflected side of the mirror
as drawn. The points f and r, the focal length and radius of curvature, are located directly below the focal
point and center of curvature.
Compute the Focal Length: Since they are on the positive part of the optical axis, both are positive for this
mirror. Therefore, the signed radius of curvature is r = 60cm and the signed focal length is f = r/2 = 30cm
• Draw and label the focal point (F) and the center of curvature (C).
s′
m=−
s
The − sign comes from the fact that a real image (s′ > 0) is inverted (m < 0).
The following two examples illustrate using the mirror equation and the mathematical expression for
magnification of a convex or concave mirror.
Solution
s r s’ f 0
Definitions
s = 1m ≡ Object Distance
r = 45cm ≡ Radius of Curvature
s′ ≡ Image Distance
m ≡ Magnification
(a) Compute the Focal Length: The radius of curvature is given as r = 45cm, therefore the focal length is
f = r/2 = 22.5cm. For a concave mirror, the focal length is on the transmitted side and is therefore positive.
(b) Compute the Image Distance: Use the mirror equation to find the image distance. The information
given is object distance,
−1
1 1 1 1 1
+ = −→ s′ = −
s s′ f f s
−1
1 1
= −
0.225m 1.0m
s’ = 0.29m = 29cm
(c) Compute the Magnification: Use m = −s′ /s
0.29m
m=− = −0.29
1.0m
(d) Mathematically Interpret the Image: This is a real, s > 0, reduced |m < 1|, inverted m < 0 image.
(e) Draw Ray Diagram: Draw the mirror with the approximate radius of curvature, and the real optical
axis through the vertex. At the intersection of the vertex and the mirror, draw the plane from which the
rays will be reflected. Draw all principle rays.
• Parallel Ray (P) - Drawn parallel to the axis and bounces back through the focal point.
• Focal Ray (FF) - Drawn through focal point, bounces back parallel to the axis.
• Central Ray (CC) - Drawn to the vertex, bounces back with equal angle of reflection.
• Radial Ray (R) - Drawn through center of curvature, bounces directly back.
Where they intersect is the image, and should approximately agree with the calculated image distance
and magnification. If we do not draw the diagram to scale, tracing rays will not reproduce the answer. The
more accurate the scale of the drawing, the better the agreement should be. Draw the optical coordinate
axis below the figure.
F C
Apparent Direction
s 0 s' f r
Definitions
Strategy: Compute the image location, then draw the ray diagram. Use image location to compute the
magnification.
(a) Compute the Focal Length: The focal length is half the radius of curvature |f | = r/2 = 4cm. For a
convex mirror, the focal point is behind the mirror, so f = −4cm < 0.
(b) Compute the Image Location: The image location is given by the mirror equation
1 1 1
+ =
s s′ f
or solving for s′
1 1
s′ = 1 1 = 1 1
f − s −4cm − 6cm
12
s′ = − cm = −2.4cm
5
Solution to Part (b)
Describe the Image: The image is reduced, since |m| < 1, upright m > 0, and virtual s′ < 0. All of
this can be read from the ray diagram, the drawn image is smaller, the same orientation as the object, and
formed by the apparent direction of the rays.
Draw the Ray Diagram: Draw the mirror with correct direction of curvature and the plane through
the vertex from which we will draw the reflected rays. Draw the focal ray, central ray, parallel ray, and
radial ray for the mirror. This should converge to an image at approximately the image distance. This will
approximately happen if the object, focal point, and center of curvature are drawn to scale. The radial ray
is drawn from the object through the center of curvature C. The central ray is drawn, reflected at the point
the mirror meets the axis. The parallel ray is drawn parallel to the axis on the object side of the mirror,
then appears to bend toward the focal point. The reflection of this ray is in the same direction as the line
through the focal point. The focal ray is drawn from the object towards the focal point. The reflection of
this ray is parallel to the axis.
To locate the image, s′ , of a thin lens given the location of the object s, we found the equation
1 1 n − n0 1 1
+ = −
s s′ n0 r1 r2
If the index of refraction of the lens is n and the lens is placed in air, this becomes
1 1 1 1
+ ′ = (n − 1) −
s s r1 r2
To find the focal length, find the location where parallel light rays are focused, this will be the place an
object at infinity is focused, so if s = ∞ then s′ = f . In air,
1 1 1 1 1
+ = (n − 1) − =
∞ f r1 r2 f
Lens-maker’s Equation : For a converging (convex) lens, all light which comes in
parallel to the axis will be focused through the secondary focal point, F ′ . For a
diverging (concave) lens, the apparent direction of all light that comes in parallel
to the axis will pass through the secondary focal point. This leads to a relationship
that will describe the focal length of a lens—the lens-maker’s equation:
1 1 1
= (n − 1) −
f r1 r2
This yields a focal length with appropriate sign to locate the secondary focal point.
Focal Length Same for Both Sides: The focal length is the same for a thin lens
for either orientation of the lens with reference to the direction of incident light.
Converging and Diverging Lenses: A converging lens causes light rays from infinity
to be focused at the secondary focal point on the transmitted side of the lens by
causing the rays to bend closer together (to converge). A diverging lens causes the
apparent direction of light rays from infinity to focus at the secondary focal point
on the incident side of the lens. The focal length of a diverging lens is negative
and the focal length of a converging lens is positive.
Secondary Focal Point (F’): The focal point where the incident parallel light is
focused (converging lens) or the apparent direction is focused (diverging lens), is
the secondary focal point and is designated as F ′ on the ray diagram. The other
focal point is the primary focal point and is designated F on the ray diagram.
incident incident
light light
F F’ F’ F
0 f f 0
Primary Focal Point (F): The primary focal point is the point where a point light
source could be placed to produce outgoing parallel rays (converging lens) or where
light rays should be sent toward to produce outgoing parallel rays (diverging lens).
incident
Converging Lens light Diverging Lens
F F
incident
light
-f 0 0 -f
Types of Lenses: The figure below shows the types of lenses which can be formed
of sections of spherical surfaces. The radius of curvature and focal length of a plane
is infinity, therefore the contribution to the focal length in the lens maker’s equation
is 1/r = 1/∞ = 0. If one plays with the lensmaker’s equation, one will find for
lenses in air that a convex lens always has a positive focal length and is therefore a
converging lens and a concave lens always has negative focal length and is therefore
a diverging lens.
The radii of curvature for the lens (with appropriate signs) are r1 = 20cm and r2 = ∞. Apply the lens
maker’s equation,
1 1 1 1 1
= (n − 1) − = (1.66 − 1) − .
f r1 r2 20cm ∞
1 0.66
=
f 20cm
f = 30cm
Since the focal length is positive, the secondary focal point, F ′ , is on the transmitted side of the lens and
the lens is a converging lens.
F F’ C1
optical axis
0 f r1
The radii of curvature for the lens (with appropriate signs) are r1 = 20cm and r2 = 10cm. Apply the lens
maker’s equation,
1 1 1 1 1
= (n − 1) − = (1.66 − 1) − .
f r1 r2 20cm 10cm
1 0.66
=−
f 20cm
f = −30cm
Since the focal length is negative, this is a diverging lens. The secondary focal point, F ′ , is placed a focal
length from the lens on the incident side of the lens, with the primary focal point, F , symmetrically placed
on the other side of the lens.
F’ F C2 C1
optical axis
f 0 r2 r1
F' F'
Optical systems work backwards. If we place a point light source at the primary focal point of a converging
lens system, parallel rays will exit the system. This is how the ray boxes in lab work. If light is shone toward
the primary focal point of a diverging lens, parallel light rays exits the system.
converging lens
diverging lens
point
light
source
F F
• Draw and label the secondary focal point (F ′ ) . The secondary focal point is the location where
incoming parallel light is focused to a point.
• Draw and label the primary focal point (F ) . The primary focal point is on the opposite side of the
lens from the secondary focal point and the same distance from the lens.
• Draw the location of the points s and f on the optical axis. Indicate using > 0 and < 0 whether these
quantities are positive or negative.
• Draw the Parallel Ray and label it (P ). The parallel ray leaves the object parallel to the axis and is
refracted through the secondary focal point.
• Draw the Focal Ray and label it (F F ). The focal ray lies along the line passing through the primary
focal point and the object. At the lens, the ray is refracted parallel to the optical axis.
• Draw the Central Ray and label it (CC). The central ray travels straight through the center of the
lens.
• Draw the Apparent Directions. For each ray, draw the apparent direction—the direction the outgoing
ray appears to come from. Draw the apparent direction as a dashed line.
This will be illustrated by the examples which follow. If you need more detail on drawing the ray diagram
of either a lens or mirror consult the next chapter.
s′
m=−
s
(a)Draw the ray diagram and locate the image. Label the location of the primary and secondary
focal points.
(b)Indicate the positive direction on the optical axis. Mark important locations on the optical axis.
(c)From your ray diagram, what is the magnification and image location? Make sure to report
both with the correct sign.
(d)Describe the image based on your ray diagram.
(e)Compute the image location using the thin lens equation.
(f)Compute the magnification.
(g)Describe the image based on your computation. Tell why your calculation supports your de-
scription of the image.
Solution to Part(a)
The central ray is drawn from the object through the center of the lens. The parallel ray is drawn from the
object parallel to the axis and is refracted so that is passes through the secondary focal point. The focal ray
is drawn as if it comes from the primary focal point and is refracted parallel to the axis at the lens.
FF(Focal)
F F'
P(Parallel)
Solution to Part(b)
The optical axis is positive to the right, because that is the transmitted direction. Since the lens is convex,
the focal length is positive. Draw f at the appropriate point on the axis. Above f , draw the secondary focal
point, F ′ . Draw the primary focal point F at an equal distance from the lens on the opposite side of the
lens. Mark the location of the object s, the image s′ , and the lens 0 on the optical axis.
Solution to Part(c)
The image is 2.6cm tall and the object is 1.5cm tall, giving a magnification of m=2.6cm/1.5cm = 1.7. The
magnification is positive, because the image is upright. The image location is −4.3cm.
Solution to Part(d)
The image is upright (not flipped over), enlarged (it is taller), and virtual (it is formed of the apparent
directions of the light).
Solution to Part(e)
Solution to Part(f)
Solution to Part(g)
The image is virtual (s′ < 0), upright (m > 0), and enlarged (|m| > 1).
Example 40.7 Locate and Interpret the Image Formed by a Thin Concave Lens
Problem: An object (upright arrow) is 50.0cm from a diverging lens with a focal length with magnitude of
20.0cm. Using graphical and mathematical techniques, locate and interpret the image.
Solution
FF
CC
F’ F
s f s’ 0 −f
Definitions
(a) Draw all Possible Principal Rays: Central, Focal, and Parallel rays. Draw the situation to scale as in
the figure above.
(b) Interpret the Image: From the ray diagram, extrapolate the image and compare it to the object for
magnification and determine whether it is upright or inverted, reduced or enlarged, and a real or virtual
image. For each characteristic, give the corresponding mathematical (qualitative) interpretation. This is a
virtual s′ < 0, reduced |m| < 1, upright m > 0 image.
(c) Compute the Image Distance: Use the thin-lens equation,
1 1 1
+ = ,
s s′ f
to solve for the image distance. The image distance is
−1 −1
′ 1 1 1 1
s = − = − = −14.3cm
f s −20.0cm 50.0cm
The image distance is negative, which means a virtual image. This corresponds to the results obtained by
the graphical technique.
(d) Compute the Magnification of the Image: The magnification is found by using m = −s′ /s. The
magnification is
−s′ −14.3cm
m= =− = +0.286
s 50.0cm
Since this number is positive, the image is upright. Since it is less than one, the image is reduced. These
correspond to the results obtained by the graphical technique.
This chapter is a very detailed reference to the tracing of rays through systems of thin lenses and mirrors.
It should be used as a reference and consulted when the material of the previous chapter was not enough.
An additional example is provided for each type of optical element.
Concave Mirror
C F
s r f 0
451
41.1. RAY TRACING FOR CONCAVE MIRRORS CHAPTER 41. RAY DIAGRAM REFERENCE
Draw the Parallel Ray for a Concave Mirror : Incoming light rays that are parallel
to the axis are reflected through the focal point.
• Draw the incoming ray Choose a point on the object (object point) to analyze.
Draw a solid line (with arrow) that is parallel to the axis from this object point
to the mirror (mirror point).
• Draw the line of sight Draw a faint line through the mirror point and the
focal point. The focal point is in front of the mirror.
• Draw the outgoing (reflected) ray Draw a solid line (with arrow) along the
line of sight back from the mirror point.
• Draw the apparent source direction Draw a dashed line along the line of
sight, from the mirror point away from the reflected ray.
C F
s r f 0
Draw the Central Ray for a Concave Mirror : Incoming light rays that strike the
mirror at its vertex are reflected away at the same angle according to the law of
reflection, since the angle from the axis is the angle from the normal at this point.
• Draw the incoming ray Choose a point on the object (object point) to analyze.
Draw a solid line (with arrow) from the object point to the vertex of the mirror
(mirror point).
• Draw the outgoing (reflected) ray Draw a solid line (with arrow) along the
line of sight back from the mirror point.
• Draw the apparent source direction Draw a dashed line behind the mirror
point in the same direction as the reflected ray.
CC
C F
s r f 0
Draw the Focal Ray for a Concave Mirror : Incoming light rays that go through
the focal point are reflected parallel to the axis.
• Draw the incoming ray Using the guide line, draw a solid line (with arrow)
from the object point to the mirror point, passing through the focal point.
• Draw the line of sight Draw a faint line parallel to the axis, which extends
through the mirror point.
• Draw the outgoing (reflected) ray Draw a solid line (with arrow) along the
line of sight back from the mirror point.
• Draw the apparent source direction Draw a dashed line along the line of
sight, from the mirror point away from the reflected ray.
FF
C F
s r f 0
Draw the Radial Ray for a Concave Mirror : Incoming light rays that go through
the center of curvature are reflected directly back along the same line, since θi = 0
for this case.
• Draw the incoming ray Using the guide line, draw a solid line (with arrow)
from the object point to the mirror point, passing through the center of cur-
vature (C).
• Draw the line of sight For the radial ray, which reflects back on itself, the
ray is the line of sight.
• Draw the outgoing (reflected) ray Draw an arrow along the line of sight back
from the mirror point.
• Draw the apparent source direction Draw a dashed line along the line of
sight, from the mirror point away from the reflected ray.
C F
s r f 0
Put it All Together to Form an Image: There will be a point where either the
reflected rays or the apparent directions of the reflected rays intersect at a point.
This is the image. We draw another arrow to represent the image.
The location of the image is indicated with an s′ on the optical axis. The example which follows shows how
to use ray tracing to locate and describe the image.
s r 0
Solution
First, we location the focal point which is half the distance to the lens from the center of curvature. Draw
all possible principal rays for an object point: central, radial, focal, and parallel rays. The rays intersect at a
point where the image is formed. Draw in the image. Measure the image distance and adjust for the scale. I
get sscale = 12cm and s′scale ≈ 3.5cm by measuring the figure with a ruler. Since we are given s = 1m, then
1m
s′ ≈ (3.5cm) 12cm = 30cm. Compute the magnification as image height, 0.5cm, divided by object height,
1.7cm, as measured from the image below. m ≈ −0.5cm/1.7cm = −0.3. The − enters because the image is
inverted.
P
CC
R FF
C F
s r s’ f 0
Locate Landmarks for Convex Mirrors and Use Sign Convention : The center of
curvature C for the mirror is located a distance r from the mirror. The focal point
F is located a distance r/2 ≡ F from the mirror. The sign convention is as applied
to the figure at the right. Since both r and f are not on the transmitted side of the
mirror, r < 0 and f < 0.
Convex Mirror
F - Focal Point
C - Center of Curvature
F C
s 0 f r
Draw the Parallel Ray for a Convex Mirror : Incoming light rays that are parallel
to the axis are reflected “through” the focal point.
• Draw the incoming ray Choose a point on the object (object point) to analyze.
Draw a solid line (with arrow) that is parallel to the axis from this object point
to the mirror (mirror point).
• Draw the line of sight Draw a faint line through the mirror point and the
focal point. The focal point will be behind the mirror.
• Draw the outgoing (reflected) ray Draw a solid line (with arrow) along the
line of sight back from the mirror point.
• Draw the apparent source direction Draw a dashed line along the line of
sight, from the mirror point away from the reflected ray.
F C
Apparent Direction
s 0 f r
Draw the Central Ray for a Convex Mirror : Incoming light rays that strike the
mirror at its vertex are reflected away at the same angle according to the law of
reflection.
• Draw the incoming ray Choose a point on the object (object point) to analyze.
Draw a solid line (with arrow) from the object point to the vertex of the mirror
(mirror point).
• Draw the outgoing (reflected) ray Draw a solid line (with arrow) along the
line of sight back from the mirror point.
• Draw the apparent source direction Draw a dashed line behind the mirror
point in the same direction as the reflected ray.
F C
Apparent Direction
s 0 f r
Draw the Focal Ray for a Convex Mirror : Incoming light rays that go “through”
the focal point are reflected parallel to the axis.
• Draw a guide line Choose a point on the object (object point) to analyze.
Draw a faint line through the object point and the focal point, F .
• Draw the incoming ray Using the guide line, draw a solid line (with arrow)
from the object point to the mirror point.
• Draw the line of sight Draw a faint line parallel to the axis, which extends
through the mirror point.
• Draw the outgoing (reflected) ray Draw a solid line (with arrow) along the
line of sight back from the mirror point.
• Draw the apparent source direction Draw a dashed line along the line of
sight, from the mirror point away from the reflected ray.
F C
Apparent Direction
s 0 f r
Draw the Radial Ray for a Convex Mirror : Incoming light rays that go “through”
the center of curvature are reflected directly back along the same line, since θi = 0
for this case.
• Draw a guide line Choose a point on the object (object point) to analyze.
Draw a faint line through the object point and the center of curvature, C.
• Draw the incoming ray Using the guide line, draw a solid line (with arrow)
from the object point to the mirror point.
• Draw the line of sight The faint line from the mirror point to the center of
curvature, C, is the light of sight.
• Draw the outgoing (reflected) ray Draw a solid line (with arrow) along the
line of sight back from the mirror point.
• Draw the apparent source direction Draw a dashed line along the line of
sight, from the mirror point away from the reflected ray.
F C
Apparent Direction
s 0 f r
Put it All Together to Form an Image: There will be a point where either the
reflected rays or the apparent directions of the reflected rays intersect at a point.
This is the image.
The following example illustrates how to use ray tracing to locate and describe an image for a convex mirror.
Solution
Draw your diagram, carefully getting the focal length, radius or curvature, and the object distance to scale.
Draw your optical coordinate axis and label all these quantities. Draw an object on the diagram of a height
which makes the ray drawing convenient. Draw the four principle rays and their apparent directions. The
place where either the reflected rays or their apparent directions cross is the image. Draw an image at the
place the rays cross. Now, describe the image. The image below has the same orientation as the object and
is therefore upright. The image is smaller than the object and is therefore reduced. The image is formed by
the apparent direction of the reflected rays, so it is virtual. By measuring the image and the object height,
we can report a magnification of the image as
Image Height 1.33cm
m= = = 0.67
Object Height 2cm
where I have measured the image and object height with a rule. Depending on your printer, your diagram
may be re-scaled but you should get the same ratio.
Convex Mirror
P F - Focal Point
C - Center of Curvature
R
FF
F C
FF - Focal Ray
P - Parallel Ray
CC - Central Ray CC
R - Radial Ray
Apparent Direction
Guide Line
s 0 s' f r
converging
ni nt
C2 F F’ C1
s r2 −f 0 f r1
Draw the Parallel Ray for a Converging Lens : Incoming rays that are parallel to
the axis are refracted “through” the secondary focal point, F ′ .
• Draw the incoming ray Choose a point on the object (object point) to analyze.
Draw a solid line (with arrow) that is parallel to the axis from this object point
to the centerline of the lens (lens point).
• Draw the line of sight Draw a faint line through the lens point and the
secondary focal point.
• Draw the outgoing (refracted) ray Draw a solid line (with arrow) along the
line of sight that starts at the lens point and extends away from the object.
• Draw the apparent source direction Draw a dashed line along the line of
sight, which extends the outgoing ray in the opposite direction.
F F’
s −f 0 f
Draw the Central Ray for a Converging Lens : Incoming rays which go through
the center of the lens are undeflected. Draw a solid line (with arrow) from the
object point through the center of the lens.
CC
F F’
s −f 0 f
Draw the Focal Ray for a Converging Lens : Incoming rays which go “through”
the primary focal point, F , are refracted parallel to the axis.
• Draw a guide line Choose a point on the object (object point) to analyze.
Draw a faint line through the object point and the primary focal point, F ,
taking care to extend the line to a point on the lens (lens point).
• Draw the incoming ray Using the guide line, draw a solid line (with arrow)
from the object point to the lens point.
• Draw the line of sight Draw a faint line parallel to the axis which extends
through the lens point.
• Draw the outgoing (refracted) ray Draw a solid line (with arrow) along the
line of sight from the lens point away from the object.
• Draw the apparent source direction Draw a dashed line along the line of
sight, from the lens point towards the object.
FF
F F’
s −f 0 f
• and the general direction of the light after refraction (arrowhead direction)
Locate Landmarks for a Diverging Lens: A concave lens has two interfaces, the
centers of curvature (C1 and C2 ) are located r1 < 0 and r2 > 0 from their respective
interfaces. The secondary focal point F ′ is located a focal length, f , from the
centerline of the lens. The primary focal point is located a distance −f from the
center of the lens.
diverging
ni nt
C1 F’ F C2
s r1 f 0 −f r2
Draw the Parallel Ray for a Diverging Lens: Incoming rays that are parallel to
the axis are refracted “through” the secondary focal point, F ′ .
• Draw the incoming ray Choose a point on the object (object point) to analyze.
Draw a solid line (with arrow) that is parallel to the axis from this object point
to the centerline of the lens (lens point).
• Draw the line of sight Draw a faint line through the lens point and the
secondary focal point.
• Draw the outgoing (refracted) ray Draw a solid line (with arrow) along the
line of sight that starts at the lens point and extends away from the object.
• Draw the apparent source direction Draw a dashed line along the line of
sight, which extends the outgoing ray in the opposite direction.
F’ F
s f 0 −f
Draw the Central Ray for a Diverging Lens: Incoming rays which go through the
center of the lens are undeflected. Draw a solid line (with arrow) from the object
point through the center of the lens.
CC
F’ F
s f 0 −f
Draw the Focal Ray for a Diverging Lens: Incoming rays which go “through” the
primary focal point, F , are refracted parallel to the axis.
• Draw a guide line Choose a point on the object (object point) to analyze.
Draw a faint line through the object point and the primary focal point, F .
• Draw the incoming ray Using the guide line, draw a solid line (with arrow)
from the object point to the lens point.
• Draw the line of sight Draw a faint line parallel to the axis which extends
through the lens point.
• Draw the outgoing (refracted) ray Draw a solid line (with arrow) along the
line of sight from the lens point away from the object.
• Draw the apparent source direction Draw a dashed line along the line of
sight, from the lens point towards the object.
FF
F’ F
s f 0 −f
Optical Systems
Our final chapter investigates the human eye as an optical detector and some common instruments used
to enlarge objects for human viewing.
Example 42.1 Compute the Image Distance Formed by Two Optical Elements (Mirrors
and Lenses)
Problem: An object is 4.0cm from the center of a converging lens with a focal length of 3.0cm. It is in
front of a concave mirror with a focal length of 1.875cm. The lens and the mirror are separated by 15cm of
air. Compute the final image distance from the mirror and the total magnification.
Solution
467
42.1. OPTICAL SYSTEMS CHAPTER 42. OPTICAL SYSTEMS
final image
Mirror (b)
Lens (a)
object
intermediate image
optical coordinate axis (a)
sa 0 fa sa’
optical coordinate axis (b)
sb’ sb fb 0
D
Definitions
sa ≡ Distance of object from element a
sb ≡ Distance of image from element b
sa ≡ Distance of intermediate image from element a
sb ≡ Distance of intermediate object from element b
fa ≡ Focal length of element a
fb ≡ Focal length of element b
D ≡ Separation between the optical elements
Strategy: Compute the intermediate image distance of the first lens and then use this as the object for the
second lens.
(a) Draw the Optical System: Draw the real optical axis and the optical elements on the real optical axis.
Draw the optical coordinate axis for both elements. The optical coordinate axis for a lens points to the
right. The optical coordinate axis for the mirror points to the left. The axes have difference origins, which
are marked on each axis.
(b) Compute the Image Distance as Formed by Element a: Use Focal length version of the thin-lens
equation. The intermediate image distance is, from the thin lens equation:
1 1 1
+ ′ = .
sa sa fa
−1 −1
1 1 1 1
s′a = − = − = 12cm
fa S 3.0cm 4.0cm
(c) Compute the Magnification of Lens (a): The magnification is given by
s′a 12cm
ma = − =− = −3
sa 4.0cm
(d) Calculate the Object Distance for Element b: The object distance for the mirror is
The intermediate image is drawn in gray and is between the lens and mirror.
(e) Compute the Image Distance as Formed by Element b: The focal length of the concave mirror is
on the positive part of the mirror’s optical axis The image distance from the mirror found from the mirror
equation is
1 1 1
+ ′ =
sb sb fb
−1 −1
1 1 1 1
s′b = − = − = 5cm
fb sb 1.875cm 3.0cm
(f) Compute the Magnification of Mirror b: The magnification of element (b) is
s′b 5cm
mb = − =− = −1.67
sb 3cm
(g) Compute the Total Magnification of the System: The total magnification is the product of the
magnifications,
mT = ma mb = (−3)(−1.67) = 5
The final image distance, s′b , is positive, so the image is real. The magnification |mT | > 1 so the object is
enlarged and mT > 0 for an upright image.
Solution
(a) Compute the Intermediate Image: The thin lens equation locates the image of the first lens,
1 1 1 1 1 1
= + ′ ⇒ = +
f1 s1 s1 10cm 30cm s′1
The intermediate image distance is s′1 = 15cm. The magnification of the first lens is
s′1 15cm 1
m1 = − =− =−
s1 30cm 2
The image forms to the right, on the transmitted side, of the second lens. It forms a virtual object for this
lens.
F1 F2’ F1’ F2
final image
f2 0 s2 −f2 s2’
(b) Calculate the Object Distance for Lens 2: The object distance for the second lens is s2 = d − s′1 =
12cm − 15cm = −3cm.
(c) Calculate the Final Image Location: The thin lens equation locates the image of the second lens,
1 1 1 1 1 1
= + ′ ⇒ = +
f2 s2 s2 −4cm −3cm s′2
The final image distance is s′2 = 12cm. The magnification of the second lens is
s′2 12cm
m2 = − =− = +4
s2 −3cm
(d) Describe the Image: The total magnification is the product of the magnifications of the individual
lenses, mT = m1 m2 = (− 12 )(4) = −2. The final image is real (s′2 > 0), inverted (mT < 0), and enlarged
(|mT | > 1).
optic
aqueous humor nerve
crystalline lens
Focal Length of Eye: The human eye has an effective focal length of 1.7cm. This
treats the combination of the cornea, crystalline lens, and humors as a simple thin
lens in air. The actual size of the eye is 2.4cm and the final image is formed in the
vitreous humor.
To change the focal length of the eye, muscles in the eye must exert a force on the crystalline lens. When
these muscles are relaxed the eye is said to be relaxed. A well-functioning eye will focus incoming parallel
rays on the retina when relaxed with focal length 1.7cm. Since many of you have glasses, not all eyes are
well-functioning. The point where an object must be placed for the image to be focussed on the retina by a
relaxed eye is called the far point of the eye. For a well-functioning eye, the far point is at infinity.
The Far Point: The far point
of the human eye is the point Far Point of Normal Eye is at Infinity
that is in focus for a relaxed
eye. For a normal eye, the far
point is at infinity.
As an object is brought in from infinity the eye changes the focal length of the crystalline lens to keep
the image of object focussed on the retina. This shift in focal length is called accommodation. As the object
gets close to the eye, the eye can no longer change its focal length to focus the image on the retina. The
closest point the eye can focus is called the near point.
The Near Point of the Eye: The near point of the human eye is the closest point
where an object can be placed so that its image is focussed on the retina. The near
point is about 7cm for a teenager, 12cm for a young adult, 28cm − 40cm in middle
age, and 100cm at age 60. The near point of a normal eye is taken to be ℓnp = 25cm
for optical calculations.
A relaxed eye has focal length fr = 1.7cm which focusses parallel rays on the retina. The near point of a
normal eye is about ℓnp = 25cm. An object placed at the near point is focussed on the retina by an eye at
its minimum focal length. We can calculate the focal length of the normal eye at maximum accommodation
using the thin lens equation,
1 1 1 1 1 1
= + ⇒ = +
fnp ℓnp fr fnp 25cm 1.7cm
Solving for the focal length, gives the focal length at maximum accommodation of fnp = 1.59cm.
We can understand common eye problems in terms of the near and far point of the eye.
Nearsightedness: An eye is
nearsighted or myopic if par- Parallel Rays Focussed in Front of Eye
allel light rays are focussed
to a point in front of the
retina. Distant objects are
blurry. For a nearsighted eye,
the far point is not at infinity.
Farsightedness: Hyperopia
or farsightedness happens Parallel Rays Focussed Behind of Eye
when parallel rays are fo-
cussed beyond the retina.
The eye must accommodate
to bring far objects into fo-
cus. There is a limit to the
amount of accommodation so
the near point of the eye is Near Point of Farsighted Eye
much larger than the normal
near point of 25cm. There-
fore, objects near the eye are
blurry.
The magnification m we have been calculating is of limited use when evaluating the function of optical
instruments. For example, no matter how you cut it, the image of a star formed in a telescope is smaller
than the star itself.
image
near point αu αs
Solution
(a) Draw the System: The important distances are the distance of the eye to the near point ℓnp , the
distance of the lens to the eye ℓe = f , and the distance of the image to the eye ℓi .
αs
Near Point F’
convex lens
s’ s 0 f
i
np
(b) Solve for the image location: The location of the image can be found by the thin lens equation as
usual
1 1 1 1 1 1
= + ′ ⇒ = +
f s s 30cm 10cm s′
Solving for the image distance yields, s′ = −15cm and a magnification of m = −s′ /s = −(−15cm)/(10cm) =
3
2.
(c) Solve for the Magnifying Power: The magnifying power M P of the lens is the ratio of the angle αs
the image using the optical system makes on the retina to the angle αu on the retina with only the unaided
eye. The angle αu is calculated with the object at the near point, the closest it can be brought and still be in
focus, αu = h0 /ℓnp where h0 is the height of the object and I continue to use the small angle approximation
to turn tan α into α. With the lens, the angle on the retina is αs = hi /ℓi where ℓi is the distance of the
image to the eye and hi is image height. The magnifying power is defined as
αs hi ℓnp
MP = =
αu ho ℓi
The ratio of the image height to the object height is the magnification m = hi /ho = 23 . The distance of the
image to the eye is, observing the diagram, ℓi = ℓe − s′ = 30cm − (−15cm) = 45cm. I will use the standard
value of ℓnp = 25cm for my near point, so the magnifying power of the lens for this ℓe and s is
hi ℓnp 3 25cm
MP = = = 0.833
ho ℓi 2 45cm
For this choice, the object actually appears smaller than it would if observed without the glass at the near
point. In what follows, we investigate this system in general to find magnifying powers greater than 1. Note
the difference between the magnification and the magnifying power.
If we use the distances set up in the example, ℓnp the near point distance, ℓi the distance of the image to
the eye, and ℓe the distance of the lens to the eye we can develop a general expression for magnifying power
and investigate some useful case. As in the example, the magnifying power is
αs hi ℓnp
MP = =
αu ho ℓi
and the magnification is defined as
s′ hi
m=− =
s ho
so the magnifying power can be written
s′ ℓnp
MP = −
s ℓi
If we re-arrange the lens equation,
1 1 1 s′ s′
= + ′ ⇒ = −1
f s s s f
so the magnifying power can be written,
s′
ℓnp
MP = 1−
f ℓi
The image distance is related to the distance of the image to the eye by s′ = ℓe − ℓi , since the image distance
is negative.
ℓi − ℓe ℓnp
MP = 1 +
f ℓi
Magnifying Power of a Magnifying Glass: The magnifying power of a single lens
of focal length f is
ℓi − ℓe ℓnp
MP = 1 +
f ℓi
where ℓi is the distance from the image to the eye, ℓe is the distance from the lens
to the eye, and ℓnp is the distance from the eye to the near point.
So let’s play with this expression a little and investigate some importance cases:
• Case I: ℓe = 0: Suppose we’re holding the glass very close to the eye so that ℓe = 0. In this case the
magnifying power becomes
ℓi − ℓe ℓnp 1 1
MP = 1 + = + ℓnp
f ℓi ℓi f
In this case, the magnifying power increases as the distance of the image to the eye decreases. The
closest the image can be brought to the eye and still be in focus is the near point, so the maximum
magnifying power in this case is
1 1 ℓnp
M Pmax,ℓe =0 = + ℓnp = 1 +
ℓnp f f
For my daughter’s lens used in the example, the maximum magnifying power in this case is
ℓnp 25cm
M Pmax,ℓe =0 = 1 + =1+ = 1.833
f 30cm
• Case II: ℓe = f : I can also hold the lens so the eye is at its focal point. In this case,
ℓi − ℓe ℓnp ℓi − f ℓnp ℓnp
M Pℓe =f = 1 + = 1+ =
f ℓi f ℓi f
This was the case used in the example, because it was easy to draw, and we reproduce the magnifying
power found in the example M P = 25cm/30cm = 0.833.
• Case III: ℓi = ∞ If we place the object a focal length’s distance from the lens, the image forms at
negative ∞ and the magnifying power is
ℓi − ℓe ℓnp 1 ℓi − ℓe 1 ∞ − ℓe ℓnp
MP = 1 + = + ℓnp = + ℓnp =
f ℓi ℓi f ℓi ∞ f∞ f
So for my daughter’s lens, M Pℓi =∞ = 0.833 It is this case, that is used to determine the power of a
lens.
intermediate Image αs
s −fo fo
0 s’
fo L fe
The distance L between the focal points of the objective and eyepiece is called the tube length. Normally,
the tube length is chose to be L = 16cm as used in the figure. The microscope drawn uses an objective with
focal length fo = 2cm and an eyepiece with focal length fe = 4cm. We already know how to calculate the
magnifying power of an eyepiece, a simple magnifier, where the object is at the focal point leaving the image
at infinity, M Peyepiece = ℓnp /fe = 25cm/4cm = 6.25. Therefore the eyepiece is a 6x eyepiece. The function
of the objective is to make the object larger. The magnification of the objective is m = −s′ /s. The image
distance is fo + L, so using the thin lens equation
1 1 1 s′ s′
+ ′ = ⇒ +1=
s s fo s fo
s′ s′ fo + L L
m=− =1− =1− =−
s fo fo fo
The total magnifying power of the microscope is the product of the magnification of the objective and the
magnifying power of the eyepiece,
L ℓnp
M P = m · M Peyepiece = −
fo fe
Magnifying Power of Microscope: The magnifying power of a compound micro-
scope is given by
L ℓnp
MP = − = M Pobjective · M Peyepiece
fo fe
where L is the tube length, the distance between the two focal points, fo is the focal
length of the objective, and fe is the focal length of the eyepiece.
Magnifying Power of the Objective: The magnifying power of a microscope ob-
jective is
L
M Pobjective = −
fo
If L = 16cm and fo = 1.6cm, then M P = 10 and the objective is a 10x objective.
For the microscope drawn in above the magnifying power of the objective is M Pobjective = −16cm/2cm =
−8 so the objective is an 8x objective. The total magnifying power of the microscope M P = M Pobjective M Peyepiece =
(−8)(6.25) = −50 or 50x.
Solution
(a) Compute
the Focal Length of the Objective: The magnifying power of the objective is M Pobjective =
L
10 = fo so the focal length of the objective is fo = L/10 = 1.6cm.
(b) Compute
the Focal Length of the Eyepiece : The magnifying power of the eyepiece is M Peyepiece =
ℓnp
4 = fe so the focal length of the eyepiece is fe = ℓnp /4 = 6.25cm.
(c) Compute the Distance Between the Lenses: Reviewing the diagram at the beginning of the section,
the distance between the lenses is fo + L + fe = 1.6cm + 16cm + 6.25cm = 23.85cm
(d) Compute the Total Power: The total power of the microscope is the product of the magnifying power
of the objective and the magnifying power of the eyepiece, M P = 10 × 4 = 40.
(e) Compute the Object Location: The object must be placed to produce an image at s′ = L + fo . Using
the thin lens equation,
1 1 1 1 1 1 1 1 1
+ = ⇒ + = ⇒ + =
s s′ fo s L + fo fo s 17.6cm 1.6cm
So the object distance, s = 1.76cm.
fo fe
Unfortunately this diagram is completely useless, because it appears since the telescope moves parallel
rays together, that it makes objects smaller. A telescope views an angle αu of a distant object. Since the
length of the telescope is negligible when compared to the object distance, αu is the angle the image of the
object makes on the retina without the telescope. Trace this ray through the telescope.
objective eyepiece
αs
αu F
fo fe e
The magnifying power is the ratio M P = −αs /αu . The negative sign is needed because the image is
inverted. If the ray hits the eyepiece a distance h from the axis and ℓe is the distance from the eyepiece to
the eye, then using the small angle approximation, αu = h/(fo + fe ) and αs = h/ℓe . The magnifying power
is then
αs h/ℓe fo + fe
MP = − =− =−
αu h/(fo + fe ) ℓe
We can find ℓe by noting that since the ray passes through the center of the objective, it behaves as if it
came from a point object at the objective, so we can use the thin lens equal to determine where it again
intersects the axis at the eye. The object distance is the distance between the lenses, s = fo + fe , and the
image distance is s′ = ℓi ,
1 1 1 1 1 1
+ ′ = ⇒ + =
s s fe fo + fe ℓi fe
Therefore,
1 1 1 fo
= − =
ℓi fe fo + fe fe (fo + fe )
and the magnifying power
fo + fe fo fo
MP = − = −(fo + fe ) =−
ℓe fe (fo + fe ) fe
A refracting telescope works the same way, with the focal point of the eyepiece placed at the same point
as the focal point of the mirror. We have an additional problem, we have to get the light out of the telescope
somehow. Most refracting telescopes you see in stores use a plane mirror to bend the rays from the objective
out the side of the tube to the eyepiece.
F F
Solution
Appendix
43.1 Units
A list of unit conversions follows:
kgm
1N = 1
s2
Joule(J): The joule is unit for energy,
kgm2
1J = 1 = 1Nm
s2
Ampere(A): The Ampere is the unit used to measure current
C
1A = 1
s
Volt(V): The volt is the unit of potential difference and emf ,
Nm
1V = 1
C
Ohms (Ω): The unit for resistance is the Ohm,
V Nms Js
1Ω = 1 =1 2 =1 2
A C C
Tesla(T): A tesla is the unit used for a magnetic field,
Ns N
1T = 1 =1
Cm Am
A related unit is the Gauss, 1T = 1 × 104 G.
1Wb = 1Tm2
Henry(H): The Henry is the unit for inductance,
Wb Tm2
1H = 1 =1
A A
481
43.2. CONSTANTS CHAPTER 43. APPENDIX
43.2 Constants
The Speed of Light(c) c = 3 × 108 ms
2
C
Permittivity of Free Space(ε0 ) ε0 = 8.85 × 10−12 Nm 2.
• Diagram if Requested
• Symbolic Formulas
Solution
kq1 q2
F = Formula
d2
r
kq1 q2
d= Symbolic solution
F
s
2
(8.99 × 109 Nm
C2 )(3nC)(5nC)
d= Substitution with units
(100N )
d = 2.12 × 10−5 m An answer with units
Note, since this is a simple application of a formula no English description is required for a good solution
and no diagram was requested.
The above is pretty minimal. If any steps were left out of the above example, it would be an example of
a poor solution and might not score all the points it deserved on a homework or a test. A great solution will
also have
• Explanation of what is being done in words. In this case, Use Coulomb’s for the electric force and
solve for the separation, d.
• Guiding English, section headings, formula names, etc. In this case the phrase Apply Coulomb’s Law.
Solve for separation.
• Vital reasoning explained in complete sentences. None was needed in the example.
Solution
Strategy: Use Coulomb’s law and then solve for the separation.
(a) Use Coulomb’s Law for the electric force,: Since both
charges are positive the force is repulsive.
kq1 q2
F =
d2
F
3nC 5nC
Index
The index is on the next page. I’m still playing with this part of text.
484