- Electrostatics of Conductors
- Solitary Sphere
- Assembly of Conductors: Capacitance Coefficients
- Uncharged Conductors in External Field: Polarization Tensor
- Solitary Long Cylinder
- Forces on Conductors
- Electrical and Thermal Resistance Related to Capacitance
- Electrical Images
- Conformal Mapping
- References
Chapter 1: Fundamentals Free
-
Published:2021
John Lekner, "Fundamentals", Electrostatics of Conducting Cylinders and Spheres, John Lekner
Download citation file:
This introductory chapter summarizes fundamentals of electrostatics, in part to introduce concepts and notation. Some readers may wish to skim or skip it. Capacitance and polarization coefficients are defined for assemblies of conductors. Also considered are forces on conductors, electrical images, and conformal mapping. The appendices discuss prolate spheroids and a long cylinder of elliptical cross-section.
1.1 Electrostatics of Conductors
We start with Gauss's law and Maxwell's formulation of Faraday's law:
When charges are at rest (electrostatics) there is no magnetic field due to their motion; we also assume that any other magnetic field present is static. Hence the curl of E is zero, and the electric field can be written as the gradient of a potential,
It then follows from Gauss's law that the potential V satisfies Poisson's equation,
An electric field causes charges to move. In insulators the electrons and nuclei adjust locally; in conductors some of the electrons can move macroscopically. Hence in electrostatics the electric field inside a conductor is zero: the charges have moved to make it so. The potential is constant inside and on the conductor. It follows that any excess charge resides on the surface. (A metal contains nuclei and electrons, some of the latter being free to move. The bulk is neutral, with proton charges being cancelled by electron charges. In Gauss's law the charge density ρ is zero inside the conductor. It is understood that field and charge density have been averaged over a volume large enough to contain many nuclei and electrons, but still microscopic.)
Note that the electrostatic field is normal to a conducting surface: if there were a tangential component, charges would move. The surface charge density is proportional to the magnitude of the field, as shown in the following. Gauss's law in its integral form, namely that the integral of the scalar product of the field with an (outward-facing) surface element over a closed surface is equal to 4π times the charge enclosed by the surface, , applied to a cylindrical volume element with one end-face inside the conductor and the other just outside, gives the local surface charge density σ:
1.2 Solitary Sphere
We recall some basic formulas. A single charge Q at the origin has the electric potential
It generates a radial (or central) electric field
A dipole p has the potential and field
1.2.1 Polarizability
A conducting sphere in an external field experiences charge separation: it becomes “polarized.” We shall take the external field to be E0 = [0, 0, E0] (pointing along the positive z direction). Then positive charge accumulates at the north pole of the sphere, negative charge at the south. Far from the sphere the resulting field due to this charge redistribution is that of a dipole, of magnitude p. The polarizability α is defined by p = αE0. We shall see that the polarizability of an isolated conducting sphere of radius a is α = a3.
We wish to find the electric potential V(r) everywhere outside a sphere of radius a in the specified external field; it is constant on and inside the conducting sphere. The potential must be V = −E0z far from the sphere, which we take to be centered on the origin. There are no charges outside the sphere, apart from those at infinity, which give rise to the specified external field. Hence the Poisson equation to be satisfied by the potential V(r) becomes the Laplace equation. In spherical polar coordinates (r, θ, ϕ) this reads
Here we need only solutions independent of the azimuthal angle ϕ; these are of the form V(r, θ) = rnPn(cos θ), r−n−1Pn(cos θ) where Pn(cos θ) are the Legendre polynomials, of which the first four are
The external potential is Vext = −E0z = −E0r cos θ; its polar-angle dependence is purely in P1 = cos θ. Hence we look for a solution in the form V = −E0r cos θ + Ar−2 cos θ + V0. We can choose the constant potential on the sphere to be zero, in which case, for a conducting sphere of radius a,
The potential is zero on and inside the sphere, and correctly tends to the external potential at large distances. The term E0a3r−2 cos θ corresponds to that of a dipole at the center of the sphere, V(r) = p · r/r3, with p pointing in the direction of the external field, upwards in Fig. 1.1.
Electric potential (contours and shading) and electric field (arrows) of a conducting sphere in an external field, which points up in the figure. The field at the sphere is normal to it, and its magnitude determines the surface charge density. There is zero charge density on the equator, where the V = 0 equipotential surface (the plane z = 0) intersects the conducting sphere. The charge density is maximum (positive and negative) at the poles.
Electric potential (contours and shading) and electric field (arrows) of a conducting sphere in an external field, which points up in the figure. The field at the sphere is normal to it, and its magnitude determines the surface charge density. There is zero charge density on the equator, where the V = 0 equipotential surface (the plane z = 0) intersects the conducting sphere. The charge density is maximum (positive and negative) at the poles.
The normal component of the field at the surface of the sphere is found by taking the radial derivative of the potential function (1.11). It is
The electric field is three times larger than the applied external field E0 at each pole. The electric field magnitude at the surface of the sphere is maximum at the poles, zero at the equator. There is an accumulation of positive charge at the north pole, and negative charge at the south pole, so at a distance the polarized sphere acts as a dipole.
Comparison of (1.12) with (1.8) shows that the magnitude of the dipole is p = E0a3. The polarizability α is defined by p = αE0, so the polarizability of an isolated conducting sphere of radius a is a3.
1.2.2 Capacitance
Capacitance is a measure of the ability to store charge. Usually it is applied to two or more conductors; the general formulation is due to Maxwell and will be discussed in Sec. 1.3. For two conductors, with charge separation leading to positive and negative charges +Q, −Q on them, the capacitance C(Q, −Q) is defined in terms of Q and the potential difference V between them:
One can also ask, what is the capacitance of two conductors kept at the same potential V? (They may be connected by a thin wire, or to a common voltage source, for example.) The total charge on the two conductors is then Q′, say, and the capacitance is
For a single isolated sphere with total charge Q on it, we can regard the second conductor as being at infinity, with charge −Q. Field lines diverge radially from the charged sphere. Outside the conducting sphere (for r > a where a is its radius), the potential is given by (1.6), so the potential on the sphere is V = Q/a. Hence the capacitance of an isolated sphere is equal to its radius:
1.2.3 Inequalities
Let C = C(Q, −Q) be the capacitance of a conducting body of maximum diameter D, surface area A, volume V, and integral of the mean curvature over the surface M. Maxwell shows that C ≤ D/2 (Maxwell, 1891 vol. 1, p 117). The other conductor is understood to be at infinity, as in the previous paragraph. It is known that A3 ≥ 36πV2 for all solid bodies, and that M2 ≥ 4πA and 2πD ≥ M ≥ 4πC for convex bodies (Szegö, 1945). Define the radii
Then RM ≥ RA ≥ RV, the first inequality holding for convex bodies, the second for all bodies. Combined with Maxwell's inequality and the Poincaré theorem that the capacitance is greater or equal to RV, we have the bounds
All these lengths, and RA, are equal in the case of a spherical conductor.
1.3 Assembly of Conductors: Capacitance Coefficients
The electrostatics of conductors is simple in that each is at a fixed potential: if the electric potential V(r) varied on the conductor, its gradient would cause an electric field , and the mobile charges (which characterize a conductor) would flow until their redistribution gave a constant potential.
The energy of a system of N conductors with charges Qi and potentials Vi is (Maxwell, 1891, Sec. 84; Landau and Lifshitz, 1960, Sec. 2; Jackson, 1975, Sec. 1.11)
The capacitance coefficients Cij of a system of conductors are defined by the equations (Maxwell, 1891, Sec. 87)
If we know the capacitance coefficients, we can calculate the electrostatic energy for specified potentials Vi on the conductors:
Alternatively, we may specify the potentials, and invert the matrix C formed by the coefficients Cij in Eq. (1.19), which we may write in vector form as Q = CV. Its inverse is V = C−1Q, so the energy may be written as W = Q · V = (CV) · V = Q · C−1Q. In either case the energy will be a function of the relative positions of the conductors (through the position dependence of the capacitance coefficients), and by differentiations with respect to the coordinates we can find the forces acting on the conductors.
The work required to change the charges by amounts δQi is the work required to bring these charge increments from infinity (where the potential is zero) to the conductors:
The change in energy may also be expressed in terms of the change in potentials of the conductors (Landau and Lifshitz, 1960, Sec. 2):
Hence
The reverse order of differentiation gives Cij for the same quantity. Hence the capacitance matrix is symmetric:
In the case of two conductors a, b we need three coefficients: Caa, Cab, Cbb. From (1.19) the charges on the conductors are
When the conductors carry equal and opposite charges Q, −Q the potentials on them are
The capacitance of the two-conductor system is then the magnitude of the charge on each divided by the potential difference
Alternatively, the conductors may be kept at the same potential V. They then acquire charges
The total charge on the pair is then Qa + Qb = (Caa + 2Cab + Cbb)V. The capacitance of the pair is then
The capacitances C(Q, −Q) and C(V, V) are distinct physically; confusion between the two has led to errors, as discussed in Sec. 5 of Lekner (2011). We shall see in Chap. 4 that, for spheres in close approach, C(V, V) tends to a limit at contact (2a ln 2 in the case of equal radii), while C(Q, −Q) diverges logarithmically with the separation between the spheres. For spheres the capacitance coefficients depend on their radii a, b and on the distance between them. In general, the capacitance coefficients depend on the conductor geometric shapes together with the associated sizes, their separation, and their mutual orientation.
The electrostatic energy of a two-conductor system, with charges Qa, Qb, is, from (1.20),
Suppose now that the two conductors come close enough for an electrical short to occur, for example by electron tunneling at some local surface roughness. Charge will flow until the conductors are at the same potential V = (Qa + Qb)(Caa + 2Cab + Cbb)−1, from (1.30). The new charges are, from (1.19),
The energy after contact is, on substituting the new charges into (1.31) and simplifying,
We expect to be lower than the energy W before contact is made, otherwise charge would not flow. The electrostatic energy difference is
The factors in the denominator are both positive: if they were not, the physical capacitances C(Q, −Q) and C(V, V) defined in (1.28) and (1.30) would be negative. Hence
The capacitance inequalities in (1.35) are clearly true at large separation between the conductors, since then Cab → 0 and Caa, Cbb are positive. We shall verify the capacitance inequalities for two spheres, at any separation, in Sec. 4.2.
The equality of occurs when the charges on the conductors are in the special ratio
Since charge is conserved, this implies that the energies before and after contact are equal when (and only when) and . The distribution of charge before and after contact can be very different. Nevertheless, the energy W can approach when the capacitance coefficients are singular, as we shall see in Sec. 4.6.
1.4 Uncharged Conductors in External Field: Polarization Tensor
We saw in Sec. 1.2 that a solitary sphere of radius a has polarizability α = a3: its dipole moment in external field Eext is p = a3Eext. Here we wish to consider the electric potential of a localized assembly of conductors carrying zero net charge. Instead of p = αEext with a scalar polarizability and dipole parallel to the external field, the polarizability becomes a tensor:
The polarization tensor is symmetric, αij = αji (Landau and Lifshitz, 1960, Sec. 11). Any non-spherical body has tensor polarizability, and so a localized assembly will also have tensor polarizability. The electrostatic energy of a neutral localized assembly of conductors is W = −p · Eext, as an extension of the argument of Landau and Lifshitz (1960) Sec. 2, for a single conductor shows:
A uniform field may be represented as due to a charge Q∞ at infinity, and the energy of the system is W = Q∞V∞, where V∞ is the potential at the remote charge due to the uncharged conductors. The external field Eext causes the conductors to acquire a dipole moment p, the potential of which is p · r/r3. Hence W = Q∞p · r/r3. However, −Q∞r/r3 is just the field at the conductor due to Q∞. Hence W = −p · Eext.
The dipole moment of the assembly is specified by the polarizability tensor αij, in (1.37). Hence the electrostatic energy of the assembly in the external field Eext is
Only the dipole moment components of an uncharged system are needed to find the system energy in a uniform external field; the quadrupole and higher-order moments do not contribute. The higher multipoles are superpositions of equal and opposite pairs of dipoles, and thus have cancelling positive and negative terms.
The polarization tensor is real and symmetric and can thus be diagonalized. For simple systems it can be put in diagonal form by choice of coordinates. Suppose for example that for a two-sphere system we chose the z axis to coincide with the line through the sphere centers. Then the polarization tensor is
The transverse part is αxx = αyy = αT (response to a field perpendicular to the line of centers) and the longitudinal part (response to a field along the line of centers) is αzz = αL. Suppose the angle between the external field and the line of centers is θ,
Then the electrostatic energy (1.38) is independent of the azimuthal angle ϕ:
There is no net force on the polarized two-sphere system, but there is a torque because of the angle-dependence. The magnitude of the torque is
The polarizability tensor of the two-sphere system will be discussed in detail in Secs. 3.3–3.5. We shall see that αL > αT and that the torque always acts to align the spheres with the field.
1.5 Solitary Long Cylinder
In cylindrical polar coordinates is the distance from the z axis. We shall consider a circular cylinder of radius a, with its axis coinciding with the z axis. The length 2b of the cylinder is taken to be much larger than a. The results of this section apply only to regions far (in comparison with a) from the ends of the cylinder, that is for |z| ≪ b, and also when ρ ≪ b. The dependence on the variable z can then be neglected. With this understood, we can derive a useful result for the polarizability per unit length. The capacitance per unit length is not defined, as we shall see. The difficult problems associated with a finite cylinder are left to Chap. 8.
The Laplace equation in cylindrical polar coordinates reads
If we neglect the dependence on z, as discussed above, the solutions are
First, we shall discuss the polarizability per unit length of the cylinder, because this is easier. The external field is taken to be perpendicular to the cylinder, say Eext = [E0, 0, 0]. The potential associated with the external field is −E0 ρ cos ϕ, and the total potential, taken to be zero on the cylinder, is
To obtain the dipole moment and thus the polarizability, we need the concept of a two-dimensional dipole. This can be constructed from two parallel line charges, with charges per unit length ±q, as in Lekner (2013). From Gauss's law the field at distance ρ from a line charge +q is E = 2q/ρ, so the potential is V = constant −2q ln ρ. If the lines with charges ±q are at [±δx, 0], and , the potential due to the line charge pair is 2q ln (ρ−/ρ+). The line pair separation is 2δx and the two-dimensional dipole moment per unit length is p = 2qδx. We take the limit as δx → 0 of the potential, keeping p fixed:
Comparison with the potential (1.45) shows that the cylinder of radius a in an external field of strength E0 has dipole moment p = E0a2 and thus polarizability α per unit length
Figure 1.2 shows how an external field is modified by a long conducting cylinder (the external field is perpendicular to the cylinder axis).
Electric potential (contours and shading) and electric field (arrows) of a long conducting cylinder in an external field, which points to the right in the figure. The field at the cylinder surface is maximum (and twice the external field) at the left and right, and zero at the top and bottom of the cross-section shown in the figure.
Electric potential (contours and shading) and electric field (arrows) of a long conducting cylinder in an external field, which points to the right in the figure. The field at the cylinder surface is maximum (and twice the external field) at the left and right, and zero at the top and bottom of the cross-section shown in the figure.
Now we try to find the capacitance per unit length: let a cylinder of radius a carry a charge per unit length q. Because of the logarithmic nature of the potential the second conductor cannot be at infinity. Let it be a coaxial cylinder at some finite radius R, carrying a charge per unit length −q. The potential in a ≤ ρ < R is V(ρ) = 2q ln R/ρ; at the inner cylinder the potential is V0 = 2q ln R/a, dependent on the radius of the large cylinder. The outer cylinder may be taken to be earthed, with potential zero. The capacitance per unit length of the cylinder pair, , varies with R: there is no physical limit as R is increased to infinity. There is, however, a well-defined capacitance per unit length for a pair of parallel cylinders, as we have just seen for two coaxial cylinders, and as we shall find in Chap. 7 for the general case.
For the capacitance associated with an isolated cylinder, we need to treat the finite cylinder, as we shall in Chap. 8. Maxwell (1877) showed that for a cylinder of diameter 2a and length 2b the capacitance has the form, when b ≫ a,
The capacitance per unit length C/2b depends on the length to diameter ratio, logarithmically. The seemingly reasonable assumption of a constant charge per unit length is wrong, although it provides a first approximation when the cylinder is long. The charge density increases toward the ends of the cylinder, and this non-uniformity cannot be ignored (Griffiths and Ye 1995, Jackson 2000, 2001).
1.6 Forces on Conductors
The surface of a conductor experiences an outward force per unit area (a negative pressure), equal in magnitude to the energy density of the field just outside the conductor (Landau and Lifshitz, 1960, Sec. 5; Jackson, 1975, Sec. 1.11; Griffiths, 1981, Sec. 2.5.3). The energy density is u = (8π)−1E2 in the immediate neighborhood of the surface. A small outward displacement Δr of an element of area ΔA produces an electrostatic energy decrease ΔU = −uΔrΔA. The outward force per unit area thus has magnitude u:
We illustrate with a simple but perhaps surprising example, based on the conducting sphere in an external field E0 of Sec. 1.2. Suppose, as in Problem 3 of Chap. 1 of Landau and Lifshitz (1960), that the sphere is cut in half at the equator, the external field pointing south to north as in Fig. 1.1. The separated hemispheres retain their positive and negative surface charge densities (it is assumed that the separation is small on the scale of the sphere radius a). By Gauss's law the surface charge density is related to the field component normal to the surface by σ = En/4π. Using the expression for the normal field component given in (1.12), namely En(θ) = 3E0 cos θ, the outward force per unit area is
The net force acting on the northern hemisphere is thus, using dA = 2πa2 sin θ dθ,
The force acts northward (upward). An equal and opposite force acts on the southern hemisphere.
But wait, you say, surely there is an attractive force between the northern positive charges and the southern negative charges? Yes, but these charges also experience the pull of the external field. What we have calculated are the forces on the conductors, the two separated hemispheres.
Suppose now that the hemispheres move apart, the northern one moving north, the southern one moving south. The charges redistribute progressively, but each retains its total charge, the one on the northern hemisphere being
When the hemispheres are far apart (on the scale of the radius a) the force on the northern one tends to
This force is larger than the force (1.51) experienced when the hemispheres had barely separated, by the factor 4/3.
1.7 Electrical and Thermal Resistance Related to Capacitance
From a mathematical point of view, all of this book is devoted to solving Laplace's equation, with the boundary conditions that the solution (the electric potential) be constant on some surfaces, usually cylinders or spheres. The flow of heat and the flow of charge (in uniform media with constant conductivity) also obey Laplace's equation: let J denote the electric current density, or the heat flow vector, and K the electrical or thermal conductivity. For steady (time-independent) flow , and or . Thus, if K is constant, or , where V and T are the electric potential and the temperature.
Suppose that we have solved the electrostatic problem of two oppositely charged conductors a and b, at potentials Va, Vb, and found the capacitance C(Q, −Q) of (1.28), C(Q, −Q) = Q/(Va − Vb). Then we have also solved the problem of steady flow between these two conductors (of electricity or heat), provided their conductivity is much larger than that of the medium surrounding them: The electric flux through a surface surrounding the conductor a with charge Q is, from Gauss's law, . By direct analogy, the total electric current through a surface surrounding the conductor a with potential Va, which ends up going into the conductor b with potential Vb, is .
Since , the current is given by the same capacitance, I = 4πKeC(Q, −Q) (Va − Vb). In the heat flow case, the total heat current is H = 4πKtC(Q, −Q)(Ta − Tb). The electrical or thermal resistances to the flow are thus
The larger the capacitance, the smaller the resistance. In Gaussian units the capacitance has the dimension of length. Roughly speaking, we expect the resistance to be proportional to the distance between the conductors, and inversely proportional to the effective area of flow. Equation (1.54) gives the precise dependence.
As a simple example of (1.54), consider the flow from a sphere of radius a to the sphere at infinity. The capacitance is a from Sec. 1.2, so the resistance to flow from the (highly conducting) sphere to infinity through a medium of conductivity K is (4πKa)−1.
In problems involving long cylinders, the appropriate capacitance per unit length is C(q, −q) where ±q are charges per unit length on the two cylindrical surfaces, and the flows and resistances are then per unit length values also. In such cases, C(q, −q) is dimensionless. One example we shall meet shortly in Appendix 1B: two long confocal elliptic cylinders. The equipotentials are confocal ellipses, the field (and flow) lines lie on confocal hyperbolas. The capacitance per unit length is given in (1B.9). Another cylinder example will be considered in Secs. 1.8 and 7.2: two parallel circular cylinders of radii a, b with distance c between their axes. The capacitance per unit length is, from (7.7) and (7.15),
The first expression applies when the cylinders are exterior to each other, the second when one encloses the other. The formulas (1.54), with the capacitances (1.55) substituted, now give the resistance per unit length of the cylinders. Feynman et al. (1964), in Chap. 12, “Electrical analogs,” discusses flow and other problems that are equivalent to electrostatic ones.
1.8 Electrical Images
The young William Thomson (later Baron Kelvin of Largs, often referred to as Kelvin) introduced the idea of fictitious charges arranged to make a chosen surface an equipotential (Thomson, 1872). In the simplest case of a charge near a conducting plane surface, the plane is made to be at zero potential by an equal and opposite charge placed equidistant on the normal to the plane at the mirror image point of the original charge; hence the name “image charge.” As in all examples to follow, what actually happens is that charge moves on the conductor to make it an equipotential. However, knowing the location and magnitude of the fictitious charges enables calculation of all electrostatic properties.
1.8.1 The Capacitance of Oppositely Charged Parallel Cylinders
A two-dimensional problem, similar to the equal and opposite point charges above, is that of two parallel line charges of opposite sign. From Gauss's law, or from Sec. 1.5, the field due to a line charge (charge q per unit length) is equal to 2q/r at distance r from the line. The potential due to line charges q and −q at a point distant by ra and rb from the two line charges is 2q ln rb/ra. The potential is constant when rb/ra is constant. As we shall see, this is on circles in planes perpendicular to the parallel line charges. The potential is thus constant on cylinders with axes parallel to the line charges. We take the line charges to intersect the [x, y] plane at [±ℓ, 0]; then
For constant rb/ra = eu, we find
Suppose the cylinders have radii a, b, and their axes are separated by distance c. The cylinder surfaces are at ua and −ub, where sinh ua = ℓ/a, sinh ub = ℓ/b; their centers are at . Hence the distance between the cylinder axes is
The outer relation gives the value of ℓ in terms of a, b, c:
Figure 1.3 illustrates the two-cylinder geometry.
Lengths in the two-cylinder problem: a, b are the radii of the two cylinders, and c is the distance between the cylinder axes. The figure is drawn for b = 2a, c = 4a. The bipolar coordinate lengths ra, rb are distances to a field point from the points [±ℓ, 0]. The cylinder surfaces correspond to constant u = ln rb/ra.
Lengths in the two-cylinder problem: a, b are the radii of the two cylinders, and c is the distance between the cylinder axes. The figure is drawn for b = 2a, c = 4a. The bipolar coordinate lengths ra, rb are distances to a field point from the points [±ℓ, 0]. The cylinder surfaces correspond to constant u = ln rb/ra.
Since rb/ra = eu, V = 2qu. The cylinder surfaces are at ua and −ub; the potential difference is therefore Va − Vb = 2q(ua + ub) = 2q(arcsinh ℓ/a + arcsinh ℓ/b). Thus the capacitance per unit length of the cylinder pair is
This problem is discussed in detail in Sec. 7.2; the solution given there is in bicylindrical coordinates u, v, one of which was used here to simplify the algebra. The first of the expressions given in (1.55) is equivalent to (1.60), as shown in Sec. 7.2.
A related problem is that of a charged cylinder alongside an earthed conducting plane. Suppose that cylinder of radius a is near a conducting surface at x = 0. Line charges ±q at [±ℓ, 0] will give zero potential on the x = 0 plane, and potential Va = 2qua = 2q arcsinh ℓ/a on the cylinder. Let d be the distance of the cylinder axis from the conducting plane (d = c/2 in our previous notation). The image of our cylinder is another cylinder of radius a with axis at [−d, 0], so ℓ2 = d2 − a2 from (1.58). The capacitance per unit length of the cylinder-plane combination is therefore
When the cylinder is far from the plane on the scale of its radius, the capacitance tends to
The opposite limit is that of a cylinder close to the conducting plane. We set d = a + s and expand in powers of s, the closest cylinder-plane separation, to find
1.8.2 Images for Spherical Conductors
Suppose a positive charge Q is placed at distance d from the center of a conducting sphere of radius a, carrying zero charge. The external charge will polarize the sphere: negative charges will go to the side facing Q, leaving positive charge on the opposite side, and the sphere surface an equipotential. A charge Q′ = −(a/d)Q placed at distance a2/d from the center of the sphere on the line toward the external charge will make the sphere surface an equipotential also: if r, r′ are the distances from Q, Q′ to a point on the surface, the potential there is a constant plus . But it follows from the lengths given that r′/r = a/d, so the potential is constant on the sphere. Just as for cylinders, discussed above, r′/r = constant corresponds to the surface of a sphere, in this case the sphere with radius a and center at the origin. [See Eq. (1.56), to which we can add a z2 term in three dimensions.]
The same construction works if the charge Q is internal to the sphere, d < a.
1.8.3 The Capacitance of a Pair of Spheres Intersecting at Right Angles
The case of two spheres cutting orthogonally is the subject of Sec. 168 of Maxwell's Treatise (Maxwell, 1891) (see also Smythe, 1950, Sec. 5.103). Figure 1.4 gives the geometry.
Orthogonally intersecting spheres, of radii a, b and distance between centers . A section through the sphere centers is shown; the circle of intersection lies in the z = 0 plane, with center at the origin, and radius ab/c.
Orthogonally intersecting spheres, of radii a, b and distance between centers . A section through the sphere centers is shown; the circle of intersection lies in the z = 0 plane, with center at the origin, and radius ab/c.
We first note that the radius of the circle of intersection is : the area of the right-angled triangle in the figure is ℓc = ab. Hence the center of the a circle is at za = a2/c, and the center of the b circle is at −zb, where zb = b2/c. It follows that the origin is the image point of the center of sphere a with respect to sphere b, and also the image point of the center of sphere b with respect to sphere a. (These relations are special to two spheres intersecting at right angles.) If charges Qa = qa, Qb = qb (that is, charges proportional to the sphere radii) are placed at the sphere centers, and Q′ = −q ab/c is placed at the origin, both spheres will be equipotential surfaces, at potential q. Note q is a charge per unit length, the same dimension as electric potential. The capacitance of the outer conducting surface formed by two spheres intersecting at right angles is thus
The equipotentials may be plotted as those of the three charges Qa, Qb, Q′. Figure 1.5 shows the equipotentials drawn for the configuration of Fig. 1.4, with a : b : c = 3 : 4 : 5.
Equipotentials around two charged overlapping spheres, intersecting at right angles. Shown is a section through the sphere centers.
Equipotentials around two charged overlapping spheres, intersecting at right angles. Shown is a section through the sphere centers.
The method of images is simplest in the case given, when the spheres intersect at right angles. One image charge Q′ and two charges Qa, Qb at the centers of the two spheres together make the composite body an equipotential. The method of images also works for angles of intersection π/n, n = 2, 3, 4, ….(n = 1 corresponds to the smaller sphere entirely inside the bigger sphere, and n → ∞ to touching spheres, in tangential contact). For finite n a finite number of images is needed (Tong, 1996; Palaniappan, 1999 and Palaniappan and Felderhof, 1999).
1.9 Conformal Mapping
Two-dimensional electrostatic problems may often be most easily solved by complex analysis. We give the main results here, and shall illustrate some of the methods by considering a long conducting cylinder of elliptical cross-section in Appendix 1B.
Any differentiable function w of the complex variable z = x + iy satisfies Laplace's equation: let primes denote differentiation with respect to z, then
Cauchy–Riemann equations. Differentiation of w = u + iv with respect to x and y gives
The Cauchy–Riemann equations follow on equating the real and imaginary parts of the two expressions for w′ in (1.66):
That the equations also hold follows by differentiation of (1.67).
The curves along which u is constant are orthogonal to the curves along which v is constant. Constant u implies dx ∂xu + dy ∂yu = 0; the slope is thus dy/dx = −∂xu/∂yu. Likewise, the slope of a v = constants curve is −∂xv/∂yv, and the product of the slopes equals −1, from the Cauchy–Riemann equations, proving that the curves cross at right angles.
If the dimensionless function v(x, y) is proportional to the potential, the curves of constant u(x, y) will give the electric field lines. This is because the electric field lines are given by dx/Ex = dy/Ey and E is proportional to the gradient of v, so dx/∂xv = dy/∂yv, giving the slope dy/dx = ∂yv/∂xv = −∂xu/∂yu, namely that of curves of constant u. Equipotentials and field lines are orthogonal, and orthogonality is preserved in the mapping from x, y to u, v. This is true of the angle of intersection of any pair of corresponding curves in the two spaces; hence the name “conformal mapping.”
There is a nice way of getting the electric field from the complex potential w = u + iv (see for example Smythe, 1950 Sec. 4.11): with z = x + iy as before,
The numerator on the right is equal to dx + idy times ∂xu − i ∂yu, or to dx + idy times ∂yv + i ∂xv, from the Cauchy–Riemann equations (1.67). Hence
Suppose v is proportional to the potential. Then the vector is proportional to the electric field, with the same proportionality constant, ex = −∂xv, ey = −∂yv. Hence the electric field and its magnitude are given by
This is a remarkable result: the electric field is determined once the conformal mapping has been found (within the region covered by the mapping, and in the absence of external fields). We just need w(z) and its derivative, not the explicit forms of v(x, y) and of its derivatives. Incidentally, the results (1.70) hold irrespective of whether v or u is chosen to be the potential function.
From Gauss's law the field at distance ρ from a line with charge q per unit length has magnitude E = 2q/ρ, so the potential is V = constant − 2q ln ρ. The complex variable z = x + iy = ρeiϕ where , and is the azimuthal angle. Hence W = −2q ln z = −2q{ln ρ + iϕ} is a suitable complex function to represent a line charge. The electric field components are given by Ex − iEy = i dW/dz = −2iq/z, or Ex = 2qx/ρ2, Ey = 2qy/ρ2, which is the radial electric field we started with. The advantage of the complex potential is that we can add contributions from line charges qn at points zn = xn + iyn: the function W becomes , with derivative . With equal and opposite line charges we regain the two-cylinder calculation of Sec. 1.8.
In special cases, the coordinates of any point on a conductor can be expressed as real functions of one real variable, whose range covers the conductor. Then a conformal mapping, and hence a solution of Laplace's equation, may be written down in terms of the unicursal (one-variable) characterization. Let the parametrizing variable be p, with x = X(p), y = Y(p). Then the relation between w and z which makes v = 0 on the conductor is
(Putting v = 0 gives x + iy = X(u) + iY(u), the real and imaginary parts of which give the parametric form of the surface of the conductor.) Examples of this method may be found in Jeans (1920), Secs. 320 and 321, in Smythe (1950), Sec. 4.16, and in Appendix 1B.
Appendix 1A: Electrostatics of prolate spheroids
This book is almost exclusively about conducting spheres and cylinders, but it may help the reader to have another object to compare and contrast with these simple shapes. One feature to compare is field enhancement in the presence of an external field. The redistribution of charge on a conducting object placed in an external electric field changes the field in its neighborhood. For example, the field close to a conducting sphere varies from three times the external field to zero (Sec. 1.2). Elongated objects can have much larger field amplification. A representative shape that can be made as pointed as we wish is the prolate spheroid. This is an ellipsoid with semiaxes a, a, b (b > a) formed by rotating an ellipse about its long axis,
The conductor surface given by (1A.1) corresponds to a fixed value of ξ in prolate spheroidal coordinates (ξ, η). These are related to the cylindrical coordinates by
For a, b variable but f fixed, sections through the z axis give confocal ellipses, with semiaxes a, b and focal points at ρ = 0, z = ±f. The prolate spheroidal coordinates have the ranges 1 < ξ < ∞, −1 ≤ η ≤ 1. The conductor surface (1A.1) corresponds to ξ = ξ0, where
(e is the eccentricity). The Laplacian of Eq. (1.43) in cylindrical coordinates (ρ, ϕ, z) transforms, in prolate spheroidal coordinates, to (ξ2 − η2)−1 times
The Laplace equation is separable in both coordinate systems. Products of the Legendre functions Pn, Qn are solutions when azimuthal dependence is absent.
As an example, consider a charged conducting spheroid. The potential is V0, a constant, on the surface ξ = ξ0. The distance r from the origin is related to the spheroidal coordinates by
At large r the spheroidal coordinates have the asymptotic forms ξ → r/f, η → z/r = cos θ, θ being the polar angle in spherical coordinates. A suitable solution has the form
Figure 1.6 shows equipotentials around a charged prolate spheroid, for which the major to minor axis ratio is b/a = 13/5; the eccentricity is e = 12/13, the focal points are shown.
On the left is a section of a prolate spheroid, with length ratios b : f : a = 13 : 12 : 5; the focal points are shown as diamonds. The right-hand side shows equipotentials around the charged spheroid, in a section through the long axis.
On the left is a section of a prolate spheroid, with length ratios b : f : a = 13 : 12 : 5; the focal points are shown as diamonds. The right-hand side shows equipotentials around the charged spheroid, in a section through the long axis.
If the total charge on the spheroid is Q, its capacitance is C = Q/V0, and the potential at large distance is Q/r. Hence
The last expression is the leading asymptotic term for b ≫ a, a long thin spheroid. This is similar to the capacitance of a cylinder of circular cross-section (length 2b, diameter 2a), with b ≫ a leading term (1.48), namely C → b.
Next we look at the prolate spheroid in an external field. In the longitudinal configuration, when the external electric field E0 points along the long (z) axis of the spheroid, the potential tends to −E0z = −E0 fξη far from the ellipsoid. The potential due to the polarized spheroid tends to that of a dipole, pz/r3. Possible solutions thus involve products of P1(ξ) = ξ or Q1(ξ) = = ξ arccoth ξ − 1, with P1(η) = η. We assume zero potential on the spheroid ξ = ξ0. Then
The electric field at the surface is given by
At either pole (z = ±b, ξ = ξ0, η = ±1), the field magnification is maximum,
The field amplification factor increases without limit as ξ0 tends to 1 from above (b much larger than a) and the spheroid becomes needle-shaped. In the spherical limit (b tending to a, ξ0 tending to infinity), the field amplification factor tends to three. Since arccoth and , Eq. (1A.10) can be rewritten as
Figure 1.6 showed a spheroid with semiaxis ratio a/b = 5/13 and eccentricity e = 12/13; in this case the maximum amplification E/E0 is nearly 8.
The polarizability tensor and the torque
We obtain the longitudinal polarizability αL by comparing the potential (1A.8) due to the polarized spheroid with that of a dipole directed along the z direction, αLE0z/r3. We find
The last expression gives the leading term of αL for a highly elongated spheroid, when b ≫ a ( is the eccentricity).
For the transverse polarizability, we take the external field to be along the x axis; far from the spheroid E → [E0, 0, 0], and V → −E0x = −E0 ρ cos ϕ = −E0r sin θ cos ϕ. Possible solutions thus involve , multiplied by . We again set the potential to zero on the spheroid ξ = ξ0. Then
The first term gives the potential due to the external field, the second that due to the polarized spheroid. At large r (and thus large ξ), , so the potential due to the spheroid tends to px/r3, with p = αTE0 and
The last expression gives the leading term of αT for a highly elongated spheroid, when b ≫ a, to be compared with αL in (1A.12).
The polarizability tensor is , as in the two-sphere example of Sec. 1.4; we have taken the spheroid long axis along the z axis, and the external field in the zx plane:
From (1.38) and (1.39) the longitudinal and transverse components αL and αT determine the system energy:
The torque is obtained by differentiation of the energy with respect to the angle θ between the spheroid long axis and the field: its magnitude is, as in (1.42),
The longitudinal and transverse polarizability components both tend to b3 in the spherical limit (eccentricity e tending to zero, a tending to b). For elongated spheroids, the longitudinal polarizability dominates, as may be expected. Equation (1A.17) gives the torque: proportional to the square of the field, zero when the spheroid is aligned with the field or perpendicular to the field, maximum at 45° to the field. The torque always acts to align the spheroid with the field. Note that one can obtain the torque from the polarizability tensor components: there is no need to find and integrate the force per unit area E2/8π (Sec. 1.6).
Figure 1.7 shows equipotentials near a prolate spheroid inclined at 30° to an external field. For external field inclined at angle θ to the z axis (which is the long axis of the prolate spheroid), the potential is V = VL cos θ + VT sin θ. If the external field lies in the in the zx plane, and we view the equipotentials in the zx plane, cos ϕ = 1 in (1A.13). The zero equipotential surface meets the spheroid surface at right angles. The curve of intersection of the two surfaces is the locus of zero electric field, and thus also of zero surface charge density on the spheroid.
Prolate spheroid inclined at 30° to an external field, seen in the zx plane. The figure shows equipotentials near the spheroid, and the direction of the external field (dashed white line). The torque on the spheroid is such as to align it with the field. The length ratios are as in Fig. 1.6, b : f : a = 13 : 12 : 5.
Prolate spheroid inclined at 30° to an external field, seen in the zx plane. The figure shows equipotentials near the spheroid, and the direction of the external field (dashed white line). The torque on the spheroid is such as to align it with the field. The length ratios are as in Fig. 1.6, b : f : a = 13 : 12 : 5.
Aspects of the conducting prolate spheroid are covered in Smythe (1950), Sec. 5.281, Morse and Feshbach (1953) pp. 1284–1288, and Landau and Lifshitz (1960) Sec. 4. Application to electroporation in cancer therapy is discussed in Lekner (2014).
Appendix 1B: Long elliptical cylinder
We shall illustrate some of the conformal mapping methods outlined in Sec. 1.9 by considering a long conducting cylinder of elliptical cross-section, with semiaxes a, b.
We saw in Sec. 1.9 that any differentiable function w of the complex variable x + iy satisfies Laplace's equation; in fact both the real and imaginary parts u(x, y), v(x, y) of w(x + iy) = u + iv satisfy Laplace's equation. We saw also that if the coordinates of any point on a conductor can be expressed as real functions of one real parameter, a conformal mapping may be written down in terms of the one-parameter characterization. Our example is a cylinder of elliptic cross-section with axis along the z direction; if the semiaxes are a, b the parametrization is x = a cos ϕ, y = b sin ϕ. We define real functions u(x, y), v(x, y) by
We expand and equate real and imaginary parts, to obtain
The equipotentials v = constant (and thus a′, b′ = constant) lie on cylinders with elliptic cross sections (x/a′)2 + (y/b′)2 = 1, confocal with the a, b ellipse (x/a)2 + (y/b)2 = 1, since their focal distance f ′ is given by
Suppose we put a conducting surface at some value of v, corresponding to one of the equipotential surfaces. Since v = 0 is our original cylinder with semiaxes a, b, and we wish our original cylinder to carry a positive charge q per unit length, we take the outer conducting surface to be at v = −v0, with semiaxes
The field lines are found by eliminating the variable v corresponding to the potential, using cosh2 v − sinh2 v = 1; this gives us confocal hyperbolas,
The equipotentials and field lines of such a system are illustrated in Fig. 1.8. In the related problem of hyperbolic conductors, the role of equipotentials and field lines is interchanged, and the equipotentials are confocal hyperbolas, the field lines lie on confocal ellipses (Lekner 2004).
Equipotentials and field lines (dashed) between a long cylinder of elliptic cross-section (black, semiaxes a, b with a = 2b), and a coaxial cylinder (red). The equipotentials are confocal ellipses, the field lines lie on confocal hyperbolas.
Equipotentials and field lines (dashed) between a long cylinder of elliptic cross-section (black, semiaxes a, b with a = 2b), and a coaxial cylinder (red). The equipotentials are confocal ellipses, the field lines lie on confocal hyperbolas.
The capacitance per unit length of a two-conductor system such as depicted in Fig. 1.8 is found by calculating the ratio of the charge per unit length (±q on the two conductors) to the potential difference between them. The surface charge density is obtained from the electric field at the surface, σ = En/4π by Gauss's law.
From Eq. (1.70), the square of the electric field is , if one chooses v(x, y) to represent the potential. This holds for any differentiable relation z = F(w) between the complex variables z = x + iy and w = u + iv. Specializing to (1B.1) which reads z = a cos w + ib sin w, we find
We need the magnitude e0 of the electric field on the inner cylinder, on which v = 0. This is
The surface charge density σ = En/4π is to be integrated over the surface of the inner elliptic cylinder. The arc length is given by, from x = a cos u, y = b sin u,
The element of arc length is, in this case, proportional to the reciprocal of the field strength, so the integration of the charge density over u gives 2π/4π = 1/2. The charge per unit length of the inner cylinder divided by the potential difference between the inner and outer cylinders (on which v = 0 and v = −v0, respectively) gives the capacitance per unit length:
[We have set a′ = a0, b′ = b0, v = −v0 in (1B.5), to find .] In the circular cross-section limit, b → a, b0 → a0 = R, we regain the capacitance per unit length of Sec. 1.5, C = (2 ln R/a)−1.
The result (1B.9) was obtained in Morales et al. (2015) by the use of elliptic cylinder coordinates,
As before, fixed v′ correspond to ellipses, fixed u to hyperbolas, with foci at x = ±f, y = 0:
Comparison of (1B.15) with (1B.5) shows that v′ = v − arctanh b/a. The relation between x, y and u, v′ is again conformal, so the Laplacian transforms to . Possible solutions are linear combinations of u and v′. To make connection with the problem of two confocal conducting elliptical cylinders, the inner one of semiaxes a, b and at zero potential, and the outer one at potential V0 and with semiaxes a0, b0, we note that a solution of a = f cosh v′, b = f sinh v′ is . Hence a solution of Laplace's equation which is linear in v′ gives a potential difference between the inner and outer cylinders of , as found above.
Finally, we wish to find the longitudinal and transverse polarizabilities of an elliptical cylinder. We saw in Sec. 1.5 that a circular cylinder of radius a has the polarizability per unit length α = a2/2. In the longitudinal configuration the external electric field is along the x direction, with potential −E0x = −E0 cos u(a cosh v − b sinh v). We take the potential to be zero on the cylinder with semiaxes a, b (v = 0), and the full potential function to be
At large distance from the cylinder axis, the variable v is large and negative, and
Also tan ϕ = y/x → tan u. Thus the term due to the cylinder in (1B.12), E0a cos u ev → E0a cos ϕ . This is to be compared with the field of a two-dimensional dipole (see Sec. 1.5), 2p ρ−1 cos ϕ. The dipole moment and longitudinal polarizability are thus
In the transverse configuration, the external electric field is along the y direction, perpendicular to the long axis of the cylinder. The external potential is now −E0y = −E0 sin u(b cosh v − a sinh v). We again take the total potential to be zero on the cylinder:
The field of a two-dimensional dipole is now 2p ρ−1 sin ϕ. Hence
The more general case of a dielectric elliptic cylinder in an external field is discussed by Morse and Feshbach (1953), p. 1199.