<<

Vector for the Magnetic

Let me start with two two theorems of : Theorem 1: If a vector field has zero everywhere in space, then that field is a of some scalar field. Theorem 2: If a vector field has zero everywhere in space, then that field is a curl of some other vector field.

The first theorem allows us to introduce the for the static electric field,

E(x, y, z) = 0 x, y, z = E(x, y, z) = V (x, y, z) for some V (x, y, z), (1) ∇ × ∀ ⇒ −∇ while the second theorem allows us to introduce the for the magnetic field,

B(x, y, z) = 0 x, y, z = B(x, y, z) = A(x, y, z) for some A(x, y, z). (2) ∇· ∀ ⇒ ∇ ×

The (1) and (2) have many uses. In particular, they are needed for the Lagrangian or Hamiltonian description of a charged particle’s motion in ,

m L(r, v) = v2 qV (r) + qv A(r), (3) 2 − · 1 2 H(r, p) = p qA(r) + qV (r), (4) 2m −  or in ,

1 H = p A(r) + qV (r). (5) 2m −  b b b b I shall explain these issues — as well as Aharonov–Bohm effect and Dirac’s monopoles — in a couple of exta lectures, one next week, and one after the Thanksgiving break; see also my notes “Dynamics of a Charged Particle and Gauge Transforms” and “Aharonov–Bohm Effect and Dirac Monopoles”. But for the current set of notes, I would like to focus on using the vector potential A(x, y, z) to calculate the magentic field.

1 Let me start with some general properties of the vector potential. While the electrostatic field E(r) determines the scalar potential V (r) up to an overall constant term, the magnetic field B(r) determines the vector potential A(r) only up to a gradient of an arbitrary scalar field Λ(x, y, z). Indeed, the vector potentials A(x, y, z) and

′ A (x, y, z) = A(x, y, z) + Λ(x, y, z) (6) ∇

have the same curl everywhere, so they correspond to the same magnetic field,

′ ′ B (x, y, z) = A (x, y, z) = A(x, y, z) + Λ(x, y, z) = B(x, y, z)+0. (7) ∇ × ∇ × ∇×∇

The relations (6) between different vector potentials for the same magnetic field are called the gauge transforms.

Despite ambiguity of the vector potential itself, some of its properties are gauge invariant, i.e., the same for all potentials related by gauge transforms. For example, for any closed loop , the integral L A d~ℓ (8) · IL is gauge invariant; indeed,

′ A (r) dr A(r) dr = Λ(r) dr = dΛ(r) = 0. (9) · − · ∇ · IL IL IL IL

Physically, the integral (8) in the magnetic flux through the loop . Indeed, take any surface L spanning the loop ; by the Stokes’ theorem, S L

2 2 ΦB[through ] = B d A = ( A) d A = A d~ℓ. (10) S · ∇ × · · ZZS ZZS IL

We may use eq. (10) to easily find the vector potential for magnetic field which have some symmetries. For example, consider the uniform magnetic field B = Bˆz inside a long

2 solenoid. By the rotational and translational symmetries of the solenoid, we expect

A(s,φ,z) = A(s)φφφφφˆ, (11)

while the magnitude A(s) follows from eq. (10): Take a circle of radius s < Rsolenoid, then

A d~ℓ = A(s) 2πs, (12) · × circleI

while the magnetic flux through that circle is

2 ΦB[circle] = B πs , (13) ×

hence B πs2 A(s) = × = 1 Bs. (14) 2πs 2 In Cartesian coordinates, the vector potential becomes

A = 1 Bsφφφφφˆ = 1 B(xˆy yˆx), (15) 2 2 − which makes it easy to verify

A = 1 B(ˆx ˆy) 1 B(ˆy ˆx) = Bˆz = B. (16) ∇ × 2 × − 2 ×

Eq. (15) gives the vector potential inside the long solenoid. Outside the solenoid, the magnetic field is negligible, but the flux through a circle of radius s > Rsolenoid is non-zero due to the flux inside the solenoid. Thus,

2 ΦB[circle] = B πR (17) ×

and hence 2 2 BR 2πs A(s) = ΦB = πR B = A(s) = . (18) × ⇒ 2s

3 In vector notations,

BR2 φφφφφˆ BR2 xˆy yˆx BR2 A = = − = φ, (19) 2 s 2 x2 + y2 2 ∇

which agrees with zero magnetic field outside the solenoid,

BR2 B = A = φ = 0. (20) ∇ × 2 ∇×∇

Equations for the Vector Potential

A static magnetic field of steady currents obeys equations

B = 0, (21) ∇· B = µ0J. (22) ∇ ×

In terms of the vector potential A(x, y, z), the zero-divergence equation (21) is automatic: any B = A has zero divergence. On the other hand, the Ampere Law (22) becomes a ∇ × second-order differential equation

2 µ0J = ( A) = ( A) A. (23) ∇ × ∇ × ∇ ∇· − ∇

Moreover, for any solution A(x, y, z) of this equation for any given cuurent density J(x, y, z), there is a whole family of other solutions related to each other by the gauge transforms

′ A (x, y, z) = A(x, y, z) + Λ(x, y, z), any Λ(x, y, z). (6) ∇

To avoid this redundancy, it is often convenient to impose an extra gauge-fixing condition on the vector potential besides A = B. In magnetostatics, the most commonly used ∇ ×

4 condition is the transverse gauge A = 0. Note that any vector potential can be gauge- ∇· transformed to a potential which obeys the transversality condition. Indeed, suppose A0 = ∇· 6 0, then for ′ 1 ( A0)(r ) ′ Λ(r) = ∇· d3Vol (24) 4π r r′ ZZZ | − | we have

2 Λ(r) = A0(r) (25) ∇ −∇ · and therefore A = A0 + Λ — which is gauge-equivalent to the A0 — has zero divergence, ∇

2 A = A0 + Λ = 0. (26) ∇· ∇· ∇

In the transverse gauge, B becomes simply the (minus) Laplacian of the vector ∇ × potential,

B = ( A) = ( A) 2A 2A, (27) ∇ × ∇ × ∇ × ∇ ∇· − ∇ −→ −∇

so the Ampere Law equation (23) becomes the Poisson equation for the vector potential,

2 A(x, y, z) = µ0 J(x, y, z). (28) ∇ −

Component by component, it looks exactly like the Poisson equation for the scalar potential of the ,

− 2V (x, y, z) = ǫ 1 ρ(x, y, z), (29) ∇ 0

so its solution has a similar Coulomb-like form

′ µ0 J(r ) ′ ′ ′ A(r) = dx dy dz . (30) 4π r r′ ZZZ | − |

As written, this formula is for the volume current J(x, y, z); for a surface current K, it

5 becomes ′ µ0 K(r ) A(r) = d2A, (31) 4π r r′ surfaceZZ | − |

while for a current in a thin wire we have

′ µ0 Idr A(r) = . (32) 4π r r′ wireZ | − |

These Coulomb-like equations for the vector potential lead to the appropriate Biot– Savart–Laplace equations for the magnetic field B(x, y, z) by simply taking the curl of both sides. For example, for the volume current J(r′),

′ µ0 J(r ) ′ B(r) = A[from eq. (30)] = d3Vol ∇ × ∇ × 4π r r′  ZZZ | − |  ′ µ0 J(r ) 3 ′ µ0 1 ′ 3 ′ = r d Vol = r J(r ) d Vol 4π ∇ × r r′ 4π ∇ r r′ × ZZZ | − | ZZZ  | − | ′ ′ µ0 (r r ) ′ ′ µ0 ′ r r ′ = − − J(r ) d3Vol = J(r ) − d3Vol . 4π r r′ 3 × 4π × r r′ 3 ZZZ | − | ZZZ | − | (33) However, for practical calculations of the magnetic field, it is often easier to first evaluate the Coulomb-like integrals (30)–(32) for the vector potential and then take its curl, instead of directly evaluating the appropriate Biot–Savart–Laplace integral.

Example: Rotating Charged Sphere

Consider a uniformly charged spherical shell of radius R and charge density σ. Let’s make this sphere around its axis with angular velocity ωωωωω. Consequently, a point P on this sphere with radius-vector r′ (counted from the sphere’s center) moves with linear velocity v = ωωωωω r′, which makes for the surface current density ×

′ ′ K(r ) = σv = σωωωωω r . (34) ×

Let’s find the magnetic field of this current, both inside and outside the sphere.

6 Instead of using the Biot–Savart–Laplace equation, let’s start by calculating the vector potential from eq. (31):

′ ′ ′ µ0 K(r )= σωωωωωωωωωω r ′ µ0σ r ′ A(r) = × d2A = ωωωωω d2A . (35) 4π r r′ 4π × r r′ sphereZZ | − | sphereZZ | − |

Note: I use vector notations for the angular velocity ωωωωω instead of the spherical coordinates based on the spin axis because evaluating the integral on the RHS of eq. (35) is easier in a different system of spherical coordinates. Indeed, once I pull the ωωωωω factor outside the × integral, the remaining integral ′ r ′ d2A (36) r r′ sphereZZ | − | depends only on the r and on the sphere’s radius R, hence by spherical symmetry the vector obtaining from the integral (35) must point in the direction of r — from the center of the sphere towards the point where we evaluate the vector potential. Consequently,

µ0σ A(r) = (ωωωωω ˆr) (37) 4π × I where is the magnitude of the integral in eq. (35), or equivalently its projection on the ˆr I axis, thus ′ r ˆr ′ = · d2A . (38) I r r′ sphereZZ | − |

To take this integral, let’s use the spherical coordinates where the “north pole” θ′ = 0 points in the direction of r so that the θ′ coordinate of some point r′ on the sphere is the angle between the vectors r′ and E. Consequently,

′ ′ ′ ′ ′ ′ r ˆr = R cos θ , r r 2 = r2 + r 2 2r r = r2 + R2 2Rr cos θ , (39) · | − | − · − and hence ′ R cos θ ′ ′ ′ = R2 sin θ dθ dφ . (40) I √r2 + R2 2Rr cos θ′ × sphereZZ −

The integral over dφ′ here is trivial and yields 2π, while in the integral over dθ′ it’s convenient

7 to change the integration variable to c = cos θ′. Thus

π +1 cos θ′ sin θ′ dθ′ cdc = 2πR3 = 2πR3 . (41) I √r2 + R2 2Rr cos θ′ √r2 + R2 2Rrc Z0 − −Z1 −

To evaluate the remaining integral, we expand the denominator into Legendre polynomials in c,

Rℓ Pℓ(c) for r > R (measuring A outside the sphere), rℓ+1 × 1  ℓ =  X √ 2 2  ℓ r + R 2Rrc  r −  Pℓ(c) for r < R (measuring A inside the sphere), Rℓ+1 × ℓ  X  (42)  then note that in the numerator c = P1(c) and therefore

+1 2 2 3 for ℓ =1 Pℓ(c) cdc = δℓ,1 = (43) × 2ℓ +1 × ( 0 for any other ℓ. −Z1

Consequently, R outside the sphere, 2 r2 = 2πR3  (44) I × 3 ×  r  inside the sphere, R2  and plugging this result into eq. (37), we finally arrive at the vector potential:

4 µ0σR ωωωωω ˆr Outside the sphere, A = × , (45) 3 r2

µ0σR Inside the sphere, A = r(ωωωωω ˆr). (46) 3 ×

Now that we finally got the vector potential, the magnetic field obtains by taking its

8 curl. By the double vector product formula,

(rωωωωωωωωωω ˆr) = (ωωωωω r) = ωωωωω( r) (ωωωωω )r = ωωωωω(3) ωωωωω = 2ωωωωω , (47) ∇ × × ∇ × × ∇· − ·∇ − hence

µ0σR 2µ0σR inside the sphere, B = A = r(ωωωωω ˆr) = ωωωωω. (48) ∇ × 3 × 3   Note uniformilty of this field inside the sphere!

As the outside the sphere, by the Leibniz rule

ωωωωω ˆr ωωωωω r 1 1 × = × = (ωωωωω r) + (ωωωωω r) ∇ × r2 ∇ × r3 ∇r3 × × r3 ∇ × ×       3ˆr 1 1 = − (ωωωωω r) + (2ωωωωω) = 2ωωωωω 3ˆr (ωωωωω ˆr) (49) r4 × × r3 r3 − × × where   ˆr (ωωωωω ˆr) = ωωωωω(ˆr ˆr) ˆr(ωωωωω ˆr) = ωωωωω ˆr(ωωωωω ˆr), (50) × × · − · − · 2ωωωωω 3ˆr (ωωωωω ˆr) = 3ˆr(ˆr ωωωωω) ωωωωω. (51) − × × · −

Altogether, ωωωωω ˆr 1 × = 3ˆr(ˆr ωωωωω) ωωωωω (52) ∇ × r2 r3 · −    and therefore

4 µ0σR ωωωωω ˆr µ0σR outside the sphere, B = A = × = 3ˆr(ˆr ωωωωω) ωωωωω . (53) ∇× 3 r2 3r3 · −     Curiously, this looks like the field of a pure magnetic with dipole

2 4π QnetR m = R4σωωωωωωωωωω = ωωωωω. (54) 3 2

I shall explain the magnetic , quadrupoles, etc., later in these notes.

9 Example: Flat Current Sheet

For our next example, consider a flat current sheet in the xy plane with uniform current density K in the ˆy direction. In terms of the 3D current density,

J(x, y, z) = Kδ(z) ˆy. (55)

Consequently, the Poisson equation for the vector potential of the current sheet is

2 A = µ0Kδ(z)ˆy. (56) ∇ −

Thanks to the symmetries of this equation, we may look for a solution of the form

A(x, y, z) = A(z only) ˆy (57) where A(z) obey the 1D Poisson equation

d2A = µ0Kδ(z). (58) dz2 −

Despite the delta function on the RHS, the solution of this differential equation is continuous at z = 0, namely

1 ⋆ A(z) = µ0K z , (59) − 2 ×| |

although its derivative has a discontinuity,

dA disc = µ0K. (60) dz −  

1 ⋆ A general solution of eq. (58) is A(z) = 2 µ0K z + αz + β for arbitrary constants α and β, but the upside-down symmetry z z of the− current× sheet | | requires α = 0, while β is physically irrelevant. →−

10 This is general behavior of the vector potential for all kinds of 2D current sheets, flat or curved, with uniform or non-uniform 2D currents: The vector potential is continuous across the current sheet, but its normal derivative has a discontinuity,

∂A disc = µ0K. (61) ∂x −  normal  Consequently, the magnetic field has a discontinuity

disc(B) = µ0K n (62) ×

where n is the unit vector to the current sheet. ⊥ Multipole Expansion for the Vector Potential

Suppose I flows through a closed wire loop of some complicated shape, and we want to find its magnetic field far away from the wire. Let’s work through the vector potential according to the Coulomb-like formula

′ µ0 Idr A(r) = . (63) 4π r r′ wireI | − |

Far away from the wire, we may expand the denominator here into a power series in (r′/r), thus ∞ 1 r′ℓ = Pℓ(cos α) (64) r r′ rℓ+1 × ℓ=0 | − | X where α is the angle between the vectors r and r′,

′ cos α = ˆr ˆr . (65) ·

Plugging the expansion (64) into eq. (63) for the vector potential, we obtain

∞ µ0I 1 ′ℓ ′ ′ A(r) = r Pℓ(ˆr ˆr ) dr (66) 4π rℓ+1 · ℓ=0 X wireI — the expansion of the vector potential into magnetic multipole terms. Let me write down

11 more explicit formulae for the three leading terms,

1 ′ dr monopole r hh ii  1 I ′ ′  + 2 (ˆr r ) dr dipole µ0I  r · hh ii  A(r) =  I  . (67) 4π  1 3 ′ 2 1 ′2 ′   + ( (ˆr r ) r ) dr quadrupole   r3 2 · − 2 hh ii   I     + higher multipoles   ··· hh ii      Naively, the leading term in this expansion is the monopole term for ℓ = 0 (the top line in eq. (67)), but it vanishes for any closed current loop,

′ dr =0 (68) I Thus, the magnetic multipole expansion starts with the dipole term — which dominates the magnetic field at large distances from the wire loop. (Except when the dipole moment happens to vanish.)

Let’s simplify the dipole term in (67) using a bit of vector calculus, Let c be some constant vector. Then

′ ′ ′ ′ c (ˆr r ) dr = (ˆr r ) c dr · · · · I I by Stokes’ theorem (69) hh ii ′ 2 = r′ (ˆr r ) c d A ∇ × · · ZZ   where

′ ′ r′ (ˆr r ) c = r′ (ˆr r ) c = ˆr c, (70) ∇ × · ∇ · × ×   and hence

′ ′ 2 2 c (ˆr r ) dr = (ˆr c) d A = (ˆr c) d A · · × · × · I ZZ I = (ˆr c) a where a is the vector area of the loop (71) × · hh ii = (a ˆr) c. × ·

12 Since c here can be any constant vector, it follows that

′ ′ (ˆr r ) dr = a ˆr. (72) · × I

Finally, plugging this integral into the dipole term in the expansion (67), we arrive at

µ0 Ia ˆr µ0 m ˆr A (r) = × = × (73) dipole 4π r2 4π r2

where m = Ia is the magnetic dipole moment of the current loop.

I am going to skip over the higher multipoles in these notes. Instead, let me consider replacing a single wire loop with a circuit of several connected wires. In this case, we may use the Kirchhoff Law to express the whole circuit as several overlapping loops with independent currents; if a wire belongs to several loops, the current in that wire is the algebraic sum of the appropriate loop currents. By the superposition principle, the vector potential of the whole circuit is the sum of vector potentials of the individual loops, and as long as the whole circuit occupies small volume of size r, we may expand each loop’s A into multipoles, ≪ exactly as we did it for a single loop. In general, the leading contribution is the net dipole term,

loops µ0 mi ˆr µ0 mnet ˆr A (r) = × = × (74) dipole 4π r2 4π r2 i X where loops loops mnet = mi = Ii ai (75) i i X X is the net dipole moment of the whole circuit.

Now suppose instead of a circuit of thin wires we have some current density J(r′) flowing through the volume of some thick conductor. However, the conductor’s size is much smaller than the distance r to where we want to calculate the vector potential and the magnetic field. In this case, we may use the multipole expansion, but the algebra is a bit different

13 from what we had for a thin wire:

∞ µ0 J(r) 3 ′ µ0 1 ′ℓ ′ ′ 3 ′ A(r) = d Vol = r Pℓ(ˆr ˆr ) J(r ) d Vol , (76) 4π r r′ 4π rℓ+1 · ℓ=0 ZZZ | − | X ZZZ

or in a more explicit form

1 ′ ′ J(r ) d3Vol monopole r hh ii  ZZZ  1 ′ ′ 3 ′ + 2 (ˆr r ) J(r ) d Vol dipole µ0I  r · hh ii  A(r) =  ZZZ  . (77) 4π  1 3 ′ 2 1 ′2 ′ 3 ′   + ( (ˆr r ) r ) J(r ) d Vol quadrupole   r3 2 · − 2 hh ii   ZZZ     + higher multipoles   ··· hh ii      The monopole term here vanishes just as it did for the wire loop, albeit in a less obvious way. To see how this works, pick a constant vector c and take the divergence

′ ′ r′ (c r )J(r ) = c J + (c r)( J), (78) ∇ · · · · ∇·  where the second term on the RHS vanishes for a steady — and hence divergence-less — current. Consequently,

′ ′ ′ ′ c J(r )d3Vol = (c J(r )) d3Vol · · ZZZV ZZZV by eq. (78) hh ii ′ ′ 3 ′ = r′ (c r )J(r ) d Vol ∇ · · (79) ZZZV  by Gauss theorem hh ii ′ ′ 2 ′ = (c r )J(r ) d A · · ZZS 

where is the surface of the volume . That volume must include the whole conductor, S V but we may also make it a bit bigger, which would put the surface outside the conductor. S

14 But then there would be no current along or across , so the integral on the bottom line S of (79) must vanish. Consequently, the top line of eq. (79) must vanish too, and since c is an arbitrary constant vector, this means zero monopole moment,

′ ′ J(r )d3Vol = 0. (80) ZZZV

Next, consider the dipole term in (77) and try to rewrite it in the form (73) for some dipole moment vector m. This time, the algebra is a bit more complicated. For an arbitrary but constant vector c, we have

′ ′ ′ c ˆr (J r ) = (c J)(ˆr r ) (c r )(ˆr J), (81) · × × · · − · · ′ ′ ′ ′ ′ ′ ′ r′ (c r )( ˆr r ) J(r ) = (c J)(ˆr r ) +(c r )(ˆr J)+ (c r )(ˆr r ) ( J),(82) ∇ · · · · · · · · ∇·  where the last term vanishes for a steady current. hh ii which has J =0 hh ∇· ii and hence ′ 1 ′ 1 ′ ′ ′ (c J)(ˆr r ) = c ˆr (J r ) + r′ (c r )(ˆr r ) J(r ) . (83) · · 2 · × × 2 ∇ · ·   Consequently, dotting c with the dipole integral, we obtain

′ ′ ′ ′ ′ c (ˆr r ) J(r ) d3Vol = (c J)(ˆr r ) d3Vol · · · · ZZZV ZZZV 1 ′ ′ = c ˆr (J r ) d3Vol 2 · × × ZZZV   1 ′ ′ ′ 3 ′ + r′ (c r )(ˆr r ) J(r ) d Vol 2 ∇ · · (84) ZZZV  

1 ′ ′ ′ = c ˆr (J(r ) r ) d3Vol 2 ·  × ×  ZZZV 1 ′ ′ ′ 2 ′ + (c r )(ˆr r ) J(r ) d A 2 · · · ZZS

Similar to what we did for the monopole term, let’s take the integration volume a bit V larger that the whole conductor, so its surface is completely outside the conductor. Then S

15 on the last line of eq. (84) the current J vanishes everywhere on the surface, which kills the surface integral. This veaves us with

′ ′ ′ 1 ′ ′ ′ c (ˆr r ) J(r ) d3Vol = c ˆr (J(r ) r ) d3Vol , (85) · · 2 ·  × ×  conductor+ZZZ conductor+ZZZ   and since c is an arbitrary constant vector,

′ ′ ′ ˆr ′ ′ ′ (ˆr r ) J(r ) d3Vol = J(r ) r d3Vol · 2 × × conductor+ZZZ conductor+ZZZ (86) 1 ′ ′ ′ = r J(r ) d3Vol ˆr. 2 ×  × conductor+ZZZ   Plugging this formula into the dipole term in the vector potential (77), we arrive at

µ0 ′ ′ 3 ′ µ0 m ˆr Adipole(r) = (ˆr r ) J(r ) d Vol = × (87) 4πr2 · 4π r2 conductor+ZZZ

— exactly as in eq. (73) for the current loop — for the magnetic dipole moment

1 ′ ′ ′ m = r J(r ) d3Vol . (88) 2 × conductor+ZZZ

Let me conclude this section with the dipole term in the magnetic field,

µ0 3(m ˆr)ˆr m Bdipole(r) = Adipole(r) = · − , (89) ∇ × 4π r3 where the algebra of taking the curl is exactly as in eq. (52) earlier in these notes. In spherical coordinates centered at the dipole and aligned with the dipole moment,

µ0m sin θ A = φφφφφˆ, (90) 4π r2 µ0m 2 cos θ ˆr + sin θθθθθθˆ B = . (91) 4π r3

16 Force and Torque on a Magnetic Dipole

The B field (89) of a magnetic dipole looks exactly like the E field of an electric dipole:

µ0 3(m ˆr)ˆr m 1 3(p ˆr)ˆr p Bdipole = · 3 − , Edipole = · 3 − . (92) 4π r 4πǫ0 r Likewise, the force and the torque on a pure magnetic dipole in an external magnetic field have exactly similar form to the force and the torque on an electric dipole in an external electric field (see my notes on electric dipoles for details). In particular, in a uniform external B field, there is no net force on a dipole but there is a net torque,

Fnet = 0, τττττ net = m B. (93) × In fact, in a uniform B field these formulae are exact for any closed current loop rather than just a pure dipole. For the net force, this is trivial,

Fnet = Id~ℓ B = I d~ℓ = 0 B = 0. (94) × × I I  For the net torque, we need some algebra first:

d r (r B) = dr (r B) + r (dr B), (95) × × × × × × B (dr r) = dr (r B) r (B dr) by the Jacobi identity × × − × × − × × hh ii = dr (r B) + r (dr B), (96) − × × × × hence r (dr B) = 1 d r (r B) + 1 B (dr r). (97) × × 2 × × 2 × ×  Consequently,

τττττ net = r dF = r (Idr B) × × × Z I in light of eq. (97) hh ii I I = d r (r B) + B (dr r) 2 × × 2 × × I I (98) where the first of a total differential is zero hh ii I = 0+ B dHr r × 2 × I I = r dr B. 2 × ×  I 

17 But for any closed loop 1 r dr is the vector area a of the surface spanning that loop, 2 × hence H I r dr = Ia = m (99) 2 × I — the of the lopp, — so the net torque is indeed

~τnet = m B. (100) ×

When the external magnetic field is non-uniform the net force on a current loop does not vanish. For a small loop, the net force is related to the magnetic moment as

Fnet = (m B) (101) ∇ · where the gradient acts only on the components of B and not on the m. Let me skip the proof of this formula and simply say that it is completely similar to the force on an electric dipole in a non-uniform electric field,

Fnet = (p )E = (p E). (102) ·∇ ∇ ·

(Note that (p E) (p )E = p ( E) = 0 since E = 0.) ∇ · − ·∇ × ∇ × ∇ × For atoms and molecules, the magnetic dipole moment is fixed by the quantum effects. Consequently, the magnetic force (101) on an atom or a molecules acts as a potential force, with a

U(x, y, z) = m B(x, y, z). (103) − ·

The same potential energy — or rather its variation when the magnetic moment m changes its direction — is also responsible for magnetic the torque τττττ = m B. ×

18