Ampère's force law

In magnetostatics, the force of attraction or repulsion between two current-carrying wires (see Figure 1) is often called Ampère's force law. The physical origin of this force is that each wire generates a magnetic field, as defined by the Biot-Savart law, and the other wire experiences a magnetic force as a consequence, as defined by the Lorentz force.

The best-known and simplest example of Ampère's force law, which underlies the definition of the ampere, the SI unit of current, states that the force per unit length between two straight parallel conductors is

 F_m = 2 k_A \frac {I_1 I_2 } {r},

where kA is the magnetic force constant, r is the separation of the wires, and I1, I2 are the direct currents carried by the wires. This is a good approximation for finite lengths if the distance between the wires is small compared to their lengths, but large compared to their diameters. The value of kA depends upon the system of units chosen, and the value of kA decides how large the unit of current will be. In the SI system,[1] [2]

 k_A \ \overset{\underset{\mathrm{def}}{}}{=}\   \frac {\mu_0}{ 4 \pi} \

with μ0 the magnetic constant, defined in SI units as[3][4]

 \mu_0  \ \overset{\underset{\mathrm{def}}{}}{=}\  4 \pi \times 10^{-7} \  newtons / (ampere)2.

Thus, in vacuum, the force per meter of length between two parallel conductors – spaced apart by 1 m and each carrying a current of 1 A – is exactly

 \displaystyle 2 \times 10^{-7} N/m.

The general formulation of the magnetic force for arbitrary geometries is based on line integrals and combines the Biot-Savart law and Lorentz force in one equation as shown below. [5] [6][7]:

 \vec{F}_{12} = \frac {\mu_0} {4 \pi} \int_{L_1} \int_{L_2} \frac {I_1 d \vec{\ell}_1\ \mathbf{ \times} \ (I_2 d  \vec{\ell}_2 \ \mathbf{ \times } \ \hat{\mathbf{r}}_{21} )} {|r|^2},

where

To determine the force between wires in a material medium, the magnetic constant is replaced by the actual permeability of the medium.

Contents

Historical Background

The form of Ampere's force law commonly given was derived by Maxwell and is one of several expressions consistent with the original experiments of Ampere and Gauss. The x-component of the force between two linear currents I and I’, as depicted in the diagram to the right, was given by Ampere in 1825 and Gauss in 1833 as follows[8] :

 dF_x = k I I' ds' \int ds \frac {cos(xds) cos(rds') - cos(rx)cos(dsds')} {r^2} .

Following Ampere, a number of scientists, including Wilhelm Weber, Rudolf Clausius, James Clerk Maxwell, Bernhard Riemann and Walter Ritz, developed this expression to find a fundamental expression of the force. Through differentiation, it can be shown that:

 \frac{cos(xds) cos(rds')} {r^2} = -cos(rx) \frac{(cos\epsilon - 3 cos\phi cos\phi')} {r^2} .

and also the identity:

 \frac{cos(rx) cos(dsds')} {r^2} = \frac{cos(rx) cos\epsilon} {r^2} .

With these expressions, Ampere's force law can be expressed as:

 dF_x = k I I' ds'\int ds' cos(rx) \frac{2cos\epsilon - 3cos\phi cos\phi'} {r^2} .

Using the identities:

 \frac{\partial r} {\partial s} = cos\phi,  \frac{\partial r} {\partial s'} = -cos\phi'    .

and

 \frac{\partial^2 r} {\partial s \partial s'} = \frac{-cos\epsilon %2B cos\phi cos\phi'} {r} .

Ampere's results can be expressed in the form:

 d^2 F = \frac{k I I' ds ds'} {r^2} \left( \frac{\partial r}{\partial s} \frac{\partial r}{\partial s'} - 2r \frac{\partial^2 r}{\partial s \partial s'}\right)  .

As Maxwell noted, terms can be added to this expression, which are derivatives of a function Q(r) and, when integrated, cancel each other out. Thus, Maxwell gave "the most general form consistent with the experimental facts" for the force on ds arising from the action of ds'[9] :

 d^2 F_x = k I I' ds ds' \left[ \left( \frac{1} {r^2} \left( \frac{\partial r}{\partial s} \frac{\partial r}{\partial s'} - 2r \frac{\partial^2 r}{\partial s \partial s'}\right) %2B r \frac{\partial^2 Q}{\partial s \partial s'}\right) cos(rx) %2B \frac{\partial Q} {\partial s'} cos(xds) - \frac{\partial Q} {\partial s} cos(xds') \right]   .

Q is a function of r, according to Maxwell, which "cannot be determined, without assumptions of some kind, from experiments in which the active current forms a closed circuit." Taking the function Q(r) to be of the form:

 Q = - \frac{(1%2Bk)} {2r}

We obtain the general expression for the force exerted on ds by ds:

 \mathbf{d^2F} = -\frac{k I I'} {2r^2} \left[ (3-k)\hat{\mathbf{r_1}} (\mathbf{dsds'}) - 3(1-k) \hat{\mathbf{r_1}} (\mathbf{\hat{r_1} ds}) (\mathbf{\hat{r_1} ds'}) - (1%2Bk)\mathbf{ds} (\mathbf{\hat{r_1} ds'})-(1%2Bk) \mathbf{d's} (\mathbf{\hat{r_1} ds}) \right] .

Integrating around s' eliminates k and the original expression given by Ampere and Gauss is obtained. Thus, as far as the original Ampere experiments are concerned, the value of k has no significance. Ampere took k=-1; Gauss took k=+1, as did Grassmann and Clausius, although Clausius omitted the S component. In the non-etheral electron theories, Weber took k=-1 and Riemann took k=+1. Ritz left k undetermined in his theory. If we take k = -1, we obtain the Ampere expression:

 \mathbf{d^2F} = -\frac{k I I'} {r^3} \left[ 2 \mathbf{r} (\mathbf{dsds'}) - 3\mathbf{r} (\mathbf{r ds}) (\mathbf{r ds'}) \right]

If we take k=+1, we obtain

 \mathbf{d^2F} = -\frac{k I I'} {r^3} \left[ \mathbf{r} (\mathbf{dsds'}) - \mathbf{ds (r ds')} -\mathbf{ds'(r ds)} \right]

Using the vector identity for the triple cross product, we may express this result as

 \mathbf{d^2F} = \frac{k I I'} {r^3} \left[ \left(\mathbf{ds}\times\mathbf{ds'}\times\mathbf{r}\right)%2B \mathbf{ds'(r ds)} \right]

When integrated around ds' the second term is zero, and thus we find the form of Ampere's force law given by Maxwell:

 \mathbf{F} = k I I' \int \int \frac{\mathbf{ds}\times(\mathbf{ds'}\times\mathbf{r})} {|r|^3}

Derivation of parallel straight wire case from general formula

Start from the general formula:

 \vec{F}_{12} = \frac {\mu_0} {4 \pi} \int_{L_1} \int_{L_2} \frac {I_1 d \vec{\ell}_1\ \mathbf{ \times} \ (I_2 d  \vec{\ell}_2 \ \mathbf{ \times } \ \hat{\mathbf{r}}_{21} )} {|r|^2},

Assume wire 2 is along the x-axis, and wire 1 is at y=D, z=0, parallel to the x-axis. Let x_1,x_2 be the x-coordinate of the differential element of wire 1 and wire 2, respectively. In other words, the differential element of wire 1 is at (x_1,D,0) and the differential element of wire 2 is at (x_2,D,0). By properties of line integrals, d\vec{\ell}_1=(dx_1,0,0) and d\vec{\ell}_2=(dx_2,0,0). Also,

\hat{\mathbf{r}}_{21} = \frac{1}{\sqrt{(x_1-x_2)^2%2BD^2}}(x_1-x_2,D,0)

and

|r| = \sqrt{(x_1-x_2)^2%2BD^2}

Therefore the integral is

 \vec{F}_{12} = \frac {\mu_0 I_1 I_2} {4 \pi} \int_{L_1} \int_{L_2} \frac {(dx_1,0,0)\ \mathbf{ \times} \ ((dx_2,0,0) \ \mathbf{ \times } \ (x_1-x_2,D,0) )} {|(x_1-x_2)^2%2BD^2|^{3/2}}.

Evaluating the cross-product:

 \vec{F}_{12} = \frac {\mu_0 I_1 I_2} {4 \pi} \int_{L_1} \int_{L_2} dx_1 dx_2 \frac {(0,-D,0)} {|(x_1-x_2)^2%2BD^2|^{3/2}}.

Next, we integrate x_2 from -\infty to %2B\infty:

 \vec{F}_{12} = \frac {\mu_0 I_1 I_2} {4 \pi}\frac{2}{D}(0,-1,0) \int_{L_1} dx_1 .

If wire 1 is also infinite, the integral diverges, because the total attractive force between two infinite parallel wires is infinity. In fact, we want to know the attractive force per unit length of wire 1. Therefore, assume wire 1 has a large but finite length L_1. Then the force vector felt by wire 1 is:

 \vec{F}_{12} = \frac {\mu_0 I_1 I_2} {4 \pi}\frac{2}{D}(0,-1,0) L_1 .

As expected, the force that the wire feels is proportional to its length. The force per unit length is:

 \frac{\vec{F}_{12}}{L_1} = \frac {\mu_0 I_1 I_2} {2 \pi D}(0,-1,0) .

The direction of the force is along the y-axis, representing wire 1 getting pulled towards wire 2 if the currents are parallel, as expected. The magnitude of the force per unit length agrees with the expression for F_m shown above.

References and notes

  1. ^ Raymond A Serway & Jewett JW (2006). Serway's principles of physics: a calculus based text (Fourth Edition ed.). Belmont, CA: Thompson Brooks/Cole. p. 746. ISBN 053449143X. http://books.google.com/books?id=1DZz341Pp50C&pg=RA1-PA746&dq=wire+%22magnetic+force%22&lr=&as_brr=0&sig=4vMV_CH6Nm8ZkgjtDJFlupekYoA#PRA1-PA746,M1. 
  2. ^ Paul M. S. Monk (2004). Physical chemistry: understanding our chemical world. New York: Chichester: Wiley. p. 16. ISBN 0471491810. http://books.google.com/books?vid=ISBN0471491802&id=LupAi35QjhoC&pg=PA16&lpg=PA16&ots=IMiGyIL-67&dq=ampere+definition+si&sig=9Y0k0wgvymmLNYFMcXodwJZwvAM. 
  3. ^ BIPM definition
  4. ^ "Magnetic constant". 2006 CODATA recommended values. NIST. http://physics.nist.gov/cgi-bin/cuu/Value?mu0. Retrieved 2007-08-08. 
  5. ^ The integrand of this expression appears in the official documentation regarding definition of the ampere BIPM SI Units brochure, 8th Edition, p. 105
  6. ^ Tai L. Chow (2006). Introduction to electromagnetic theory: a modern perspective. Boston: Jones and Bartlett. p. 153. ISBN 0763738271. http://books.google.com/books?id=dpnpMhw1zo8C&pg=PA153&lpg=PA153&dq=%22ampere's+law+of+force%22&source=web&ots=uZOFz9dWv7&sig=NJp3UQvbCOvcVm7eJN4IUdlC9bs. 
  7. ^ Ampère's Force Law Scroll to section "Integral Equation" for formula.
  8. ^ O'Rahilly, Alfred (1965). Electromagnetic Theory. Dover. pp. 104. 
  9. ^ Maxwell, James Clerk (1904). Treatise on Electricity and Magnetism. Oxford. pp. 173. 

Further reading

See also