Вы находитесь на странице: 1из 7

Classical applications of the KleinGordon equation

Pierre Gravela
Collge Militaire Royal du Canada, Dpartement de Mathmatiques et dInformatique, Kingston, Ontario, Canada K7K 5L0

Claude Gauthierb
Dpartement de Mathmatiques et de Statistique, Universit de Moncton, Moncton, Nouveau-Brunswick, Canada E1A 3E9

Received 9 September 2008; accepted 4 February 2011 The quantum mechanical origin of the KleinGordon equation hides its capability to model many classical systems. We consider three examples of vibrating systems whose mathematical descriptions lead to the KleinGordon equation. These examples are adapted to applications such as the motion of suspended cables and Inca rope suspension bridges. We also discuss the correspondence between the classical and quantum settings of this equation as a way to provide an explanation of the concept of mass. 2011 American Association of Physics Teachers. DOI: 10.1119/1.3559500 I. INTRODUCTION It often happens that a model put forward to describe a given phenomenon is replaced by a better model. It also happens that the former model can be used to solve problems that are sometimes very different from the one for which it was introduced. Such a situation is the case for the Klein Gordon equation, which rst appeared in the notebooks of Schrdinger in late 1925 as an attempt to describe the de Broglie waves of the electron in the hydrogen atom.1 Because this equation does not take into account the electrons spin, it led to predictions which did not match the experimental results and was thus put aside by Pauli.2 Between April and September 1926, the KleinGordon equation independently appeared in seven papers to describe relativistic massive particles without spin.3 But, in these early days of quantum mechanics, there arose insurmountable difculties related to the interpretation of the function representing the probability density. We now know that these difculties disappear in the quantum theory of elds. The aim of the present paper is to show how the Klein Gordon equation is useful in the description of some vibrating systems in classical mechanics. Also, we shall relate the classical and quantum settings of this equation to explain the concept of mass. Because of their simplicity, we shall focus our presentation on vibrating systems made of exible strings. Recall the usual assumptions for ordinary vibrating strings. Each string is assumed to be of constant linear mass density and to be under uniform tension T. The strings are assumed to be perfectly exible and free from any internal or external friction. The amplitude of the vibrations is assumed to be sufciently small for the tension in the strings to stay constant in time. To describe the vibrations of a string, we use a Cartesian coordinate system whose x-axis is collinear with the position of the string at rest. The equation describing the vibrations in a plane can be deduced from Hamiltons principle or Newtons second law applied to an innitesimal section of the string. If u x , t denotes the displacement from rest position of the point x at time t, it is easily shown that these vibrations are locally described by the linear wave equation
447 Am. J. Phys. 79 5 , May 2011 http://aapt.org/ajp
2 tu

= c2 2u, x

where the positive constant c = T / represents the speed at which traveling waves move along the string.

II. THE BRACED STRING Our rst example of a classical mechanics phenomenon that leads to the KleinGordon equation is not new.4 We discuss it briey because it is easy to visualize many features of all phenomena described by this equation. We assume that the plane in which the string vibrates offers resistance to the strings motion. This setting can be realized by embedding the string in a piece of elastic material, such as a thin at sheet of rubber. In addition to the restoring force due to the string tension, there is then a restoring force due to the elastic material surrounding the string. The new restoring force is given by Yu x , t dx, where Y is the Young modulus of the material surrounding the string. The equation describing the vibrations of this elastically braced string is
2 tu

= c 2 2u x

u,

where c has the same meaning as in Eq. 1 and = Y / . Equation 2 is the KleinGordon equation, originally used to describe a particle of spin zero and mass m = / c, where c is the speed of light and is Plancks constant divided by 2 . In this context the KleinGordon equation is not deduced from a mechanical model as is done here for the strings. Instead it is obtained by considering the relativistically invariant energy equation, E 2 = c 2 p 2 + m 2c 4 , 3

and and applying the usual operator substitutions p i Ei / t , where p is the momentum of the particle. The resulting differential operator equation acts on a wave function .5 If the string vibrates between two rigid supports separated by a distance L, we can show that the allowed frequencies are
2011 American Association of Physics Teachers 447

n=

nc L

n = 1,2, . . . .

2 t v1

= c 2 2v 1 . x

6b

These frequencies are larger than the corresponding ones for an ordinary string for which = 0. The waves traveling along a semi-innite elastically braced string driven from its nite end by a transverse alternating force are also different from those of an ordinary string. If is the driving frequency, then the phase velocity of the waves along the braced string is larger than c by the factor 1 / 1 / 2. The additional elastic force thus increases the speed of the traveling waves along the string. However, the energy of these waves is transported according to their group velocity, which is smaller than c by the factor 1 / 2. Because these velocities depend on the frequency, a wave shape composed of waves of different frequencies will change from its original shape. Conversely, this dependence of the velocities on the frequencies means that non-localized disturbances along the string may temporarily form localized packets. These properties apply to all solutions of the KleinGordon equation with 0. To illustrate this case for a simple model, consider a cable under tension linked at regular intervals to a beam by suspension rods of length a and negligible mass. Each element of the cable behaves as the bob of a pendulum attached to the beam by a suspension rod. The cable corresponds to the string of our model and the gravitational force plays the role of the elastic material which braces the string. In addition to the tension in the cable, there is a restoring force on the cable elements due to the pendulum conguration. If is the angle of the lateral motion of a suspension rod with respect to its equilibrium position and 1, this force is proportional to g sin g , where g is the acceleration of gravity at Earths surface. It is easy to show that the horizontal motion of the suspended cable obeys the KleinGordon equation with = g / a.

The variable v0 represents the crosswise stretching of the elastic strip between the strings and describes vibrations resulting from forces without a net transverse component such that in pure modes, the center of mass at x does not move. In contrast, v1 describes the transverse motion of the elastic strip as a whole or, if preferred, the motion of the curve made up of the points always at mid-distance between the two vibrating strings. Its vibrations result from forces with a nonzero net transverse component. These vibrations correspond to two kinds of transverse waves propagating along the twostring structure, with only those of the second kind do so at the speed equal to that of waves along each of the two free strings. IV. THREE PARALLEL STRINGS This example is the most complex of the three discussed in this paper. Its use is illustrated by rope bridges, the classical example of which are the three-rope Inca suspension bridges. These bridges were a part of the Inca road system and a good example of Inca innovations in engineering.6 Such a rope bridge is till in use today near Cuzco, Peru, and crosses the 67 m of the Apurimac Canyon and is 36 m above the river. To build a three-rope bridge, a stone anchor is needed on each side of the canyon to be crossed. One massive cable made of woven grass is rst installed over the canyon to link these anchors. Two additional cables acting as guard rails are then added above and on both sides of the rst cable. They are then linked to the rst cable with smaller gauge secondary suspension cables all of the same length. Given that the rst massive cable is the footpath, the secondary suspension cables must be at an angle to the vertical sufcient to allow easy and secure crossing of the bridge. This angle must not be too large, otherwise the guard rails would be too low to be useful. Contrary to the four-cable suspension bridges which have a footpath made with plaited branches that can be used by livestock, these three-cable bridges are designed for the exclusive use of humans. Miniature versions of these bridges can sometimes be seen in playgrounds. Under normal conditions, only small relative displacements between the three main cables take place, and the secondary suspension cables are under tension because they support a part of the footpaths weight. The analysis of this case is an extension of the analysis in Sec. III except that there are three strings to consider, and we have to consider two-dimensional motion of the cable elements because each of them will move in a plane perpendicular to the main cables. Another extension is that we now have to take gravity into account. Because gravity is a constant force, it can be incorporated by a redenition of the zero of their vertical components. For this reason the differential equations are not affected by such a change. The analysis given in Appendix B shows that the overall motion decomposes into six independent quantities, each of them obeying its own differential equation separately from the others. Four of these quantities obey the ordinary wave equation, and the two others obey the KleinGordon equation. If the members in pairs of secondary cables form right angles when at rest, the quantities obeying the ordinary wave equation are the displacements of each guard rail perpendicular to the plane containing it and the footpath when at rest.
P. Gravel and C. Gauthier 448

III. TWO PARALLEL STRINGS Consider the coupling of transverse waves propagating along two parallel taut strings when the motion is restricted to the plane dened by the strings at rest. Examples corresponding to this model are the dividing nets used in the games of badminton and table tennis. In these examples the waves are vertical. For simplicity, we shall assume the coupling between the strings to be through a massless elastic material with Young modulus Y. The equations describing the motions of the coupled strings are derived in Appendix A. If u0 and u1 denote the displacements from their rest positions of the points on strings with string number 0 and 1, respectively, then
2 t u0 2 t u1

= c 2 2u 0 x = c 2 2u 1 x

u0 u1 u0 + u1 ,

5a 5b

where c and have the same meanings as in Eq. 2 . The functions u0 and u1 appear in Eqs. 5a and 5b . These equations decouple by introducing the variables v0 = u0 u1 and v 1 = u 0 + u 1:
2 t v0

= c 2 2v 0 2 x

v0

6a

448

Am. J. Phys., Vol. 79, No. 5, May 2011

Also obeying this equation in this case are the displacements of the curve dened by the centers of mass of all the pairs, each of which are made of one guard rail element and the corresponding footpath element in the same plane perpendicular to the main cables. The two quantities obeying the KleinGordon equation are the distances between the footpath and each of the guard rails. V. SYMMETRY BREAKING AND MASS The study of the symmetries of the system formed by the two parallel strings and the elastic strip between them allows us to explain the origin of the inertial mass associated with this system. We rst consider each of the two strings separately. Each string can be seen as made of innitesimal length elements joined by elastic links. If the elasticity coefcient of these links is zero, then their motions are independent of one another. This independence means that the motion of each of these elements is subject to no restriction in the three-dimensional space whose origin coincides with the string element at rest. We say that the element has three degrees of freedom, and that its motions tangent space is three-dimensional. In this case the inertial mass of each of the elements is a physically irrelevant parameter because it can be eliminated from its equation of motion. If the elasticity coefcient of the links between innitesimal elements of each string is nonzero, these elements lose their degree of freedom in the direction dened by the string because the tension is uniform within the string. Because we limit ourselves to linear elastic forces, the innitesimal elements will move only in directions perpendicular to the straight line of the string at rest. The tangent space, that is, the space of all possible velocities of the particle, will now be two-dimensional. The motions are described by Eq. 1 , where the inertial mass of the string elements partially determines the value of the parameter c. The mass parameter cannot be removed from the equations any longer, and its presence can be seen to depend on the existence of the elastic links between the string elements. The value of the mass parameter is thus a measure of the elements resistance to follow the trajectory imposed by the elastic links. The original three-dimensional isotropy in the freedom of each strings element is reduced or broken by the elastic links to a two-dimensional isotropic freedom. This symmetry breaking implies that a mass must be associated with these elements. We have seen that the vibrations of two coupled parallel strings decouple into two types of waves moving along them. The waves of one of these types correspond to motion in which the two strings move in unison and are described by Eq. 6b . This equation is the same as Eq. 1 and describes waves propagating along the system of two strings and the elastic strip. These waves are similar to the waves in each of the strings alone because the motion of the corresponding pairs of elements on the two strings has the same freedom as the elements of each string, in this case with respect to the straight line equidistant to the two strings at rest. Because the two strings have the same linear density and a combined tension that is twice that of each string, the speed at which waves move along this system is the same as the speed in each string. The other type of waves in the two coupled strings is given by Eq. 6a and corresponds to vibrations of the two strings moving in opposition. The symmetry breaking generated by the elastic strip between the strings deter449 Am. J. Phys., Vol. 79, No. 5, May 2011

mines a new inertial mass whose value is a function of the strips Young modulus. By gradually decreasing the width of the elastic strip, the waves of the latter type disappear, as well as the new mass associated with the Young modulus of the strip. The new mass is thus directly related to the existence of the elastic strip and is a measure of the resistance to crosswise stretching of the elastic strip associated with the opposed motion of the two strings. These observations suggest that inertial mass is associated with a reduction of spatial symmetry due to an interaction, where the mass is associated with the quantity that is losing symmetry. VI. LINK BETWEEN CLASSICAL AND QUANTUM APPLICATIONS We note that in the description of the coupling between vibrating strings the parameter in the KleinGordon equation is related to the Young modulus Y of the medium carrying out the coupling. In quantum mechanics this parameter is related to the mass of the spin zero particle. To see how these notions relate, we return to the system of the two elastically linked parallel strings. If the Young modulus of the elastic strip is zero, there is no connection between the strings. In this case, each energy eigenstate of the system is unique and invariant under rotations of the strings around the straight line which is coplanar and equidistant from the two strings at rest. This property applies to the systems lowest energy state. If the strings are connected and form a coherent system, then each string loses its freedom to oscillate separately, but the two strings can still form an oscillating system. Assuming that the central line around which the two strings can rotate is collinear with the x-axis in three-dimensional space, then the plane containing the two strings and the elastic strip can form any angle with the x-y plane. Any one of these angles corresponds to the systems lowest energy state, which is thus degenerate. The system must select one of these angles to x its lowest energy state. As soon as this choice is made, the system is no longer invariant under rotations of the two strings. The introduction of the medium with a nonzero Young modulus between the strings thus breaks the rotational symmetry of the systems lowest energy state. In quantum eld theory this kind of symmetry breaking is called spontaneous, because it does not require an explicit mass term in the Lagrangian to violate symmetry.7 No potential is necessary to describe the motion of innitesimal elements of string when they are not linked to form a string under tension. If the elastic link between these elements is introduced, the description of the string as a whole requires a potential with a unique minimum, such as V y,z = a y 2 + z2 + b y 2 + z2 2 , 7 with a , b 0. If we take into account the elastic link between the two parallel strings, then the description of the motion of the whole physical system needs a potential with an innity of states of lowest energy, such as Eq. 7 with a 0 and b 0. This potential expresses the fact that the system of two strings has a stable equilibrium conguration where both strings tend to stay away from the straight line that is equidistant from each strings rest position. The potential is repulsive when the strings are close to one another and attractive when they are far apart. To describe the position r x , t of one of these two strings with respect to its position at rest, we assume that in the
P. Gravel and C. Gauthier 449

vicinity of the minimum of the potential energy at r = r0, that is, when r x , t y x , t 2 + z x , t 2 r0, the elastic force of the link between the coupled strings is linear in the radial distance between r and r0. Due to the tensions in the strings and elastic strip, the system will be in its state of lowest energy when the strings are parallel at the distance r = r0 from the x-axis. In this case y x , t = r0 cos and z x , t = r0 sin for a constant angle between the elastic strip and the x-y plane. The symmetry breaking corresponding to a specic choice by the system of its actual lowest energy conguration is linked to the choice of a value for . Assume that this symmetry breaks for = 0. We use u and v to represent the position of the string element at x in a Cartesian coordinate system centered at the position of this global-symmetry breaking minimum. It follows that y = r0 + u and z = v. The potential energy part of the Lagrangian of the system then takes the form V= Y 2 Yr2 0 2 r0 + u, v r0,0
2

lus, it is possible to generate a massive scalar quantum of vibration. The new mass stems from the coupling between linear waves. Hence the notion of mass becomes related to the more fundamental notion of coupling between linear waves. In this context, no mass exists unless there is a coupling, and the stronger the coupling, the larger the mass.

APPENDIX A: TWO COUPLED VIBRATING STRINGS To obtain the equations describing the motion of the two coupled strings, the innite strings are replaced by stretched strings of nite length L xed at both ends. The two strings are assumed to have constant linear density and to be under tension T. We also assume that the strings are connected along their length by a strip of massless material having an elasticity characterized by Young modulus Y. The amplitude of waves is assumed to be sufciently small for the tension in the strings to remain uniform and constant in time. The equations describing the vibrations result from the replacement of each of the two string segments with N material points placed at intervals of length = L / N + 1 , and each point has mass = . These material points are connected together and to the xed ends of the strings by linear and massless springs which are assumed to be perfectly elastic. The tension in these small massless springs is equal to T. The motion of the material points is described using a Cartesian coordinate system whose x-axis is collinear with one of the strings when at rest. The origin coincides with the left xed point of this string. The other string is placed parallel to the x-axis at y = y 1, where y 1 is a constant. The positions of the two sets of material points at rest are given by xij = i e1 + jy je2 , A1

8a
2

1+

2ur0 + u2 + v2 r2 0

1
2

8b

Yr2 1 2ur0 + u2 + v2 0 1+ 1 2 2 r2 0 =

8c

Y 4r2u2 + 4r0u u2 + v2 + u2 + v2 0 8r2 0

8d

Near its minimum r0 , 0 , this potential is Y 2 Y Y u + u u2 + v2 + 2 u2 + v2 2 . 2 2r0 8r0 9

Any motion in the radial u direction near the bottom of the potential corresponds to widthwise stretching or compression of the elastic strip and confers a KleinGordon mass to the u variable through the Yu2 / 2 term. Motion restricted to the circular bottom of the potential corresponds to transverse movements of the elastic strip. These motions do not confer any mass to the variable v which obeys the ordinary wave equation. No energy is transferred to the elastic strip by such movements, because the elastic strip does not resist transverse deformations. The other terms in the potential function correspond to the interaction terms between the elds u and v. This situation is similar to the potential-induced global spontaneous symmetry breaking of the Goldstone process described in Appendix C. One way to interpret the repulsive part of V with a 0 and b 0 is by assigning a negative sign to the inertial mass associated with the elastic strip through the KleinGordon equation. If we dene the inertial mass of a system as the one associated with its global potential, this mass is positive because the potential V with a 0 and b 0 is attractive at large distances. VII. CONCLUDING REMARKS The quantization of linear waves in an innite vibrating string or along a set of uncoupled vibrating strings leads to a massless quantum of vibration. We have seen that when the medium between these strings has a nonzero Young modu450 Am. J. Phys., Vol. 79, No. 5, May 2011

where for the rst string y j = y 0 = 0 and where i = 0 , 1 , . . . , N + 1, j = 0 , 1, and e1 and e2 are the unit vectors along the xand y-axis, respectively. The points x0j and xN+1,j, j = 0 , 1, correspond to the xed end points of the two strings. The material point at xij is denoted by i , j . Let uij t be the displacement of the point i , j parallel to the y-axis relative to xij at time t. The functions uij t are assumed to be twice differentiable. We assume that the coupling can be expressed by a link between members of pairs of material points having the same value of their rst index. Hence, we assume that the points i , 0 and i , 1 , i = 1 , 2 , . . . , N, are connected by a massless spring with a given stiffness constant. This constant represents the resistance to transverse displacement due to the medium between the string segments of length and is denoted by k . We can determine the equation of motion for the material point 1,0 from the forces acting on it. These forces result from the displacement from their equilibrium positions of 1,0 and its neighboring points, that is, the points 0,0 , 2,0 , and 1,1 see Fig. 1 . For small displacements the transverse force on the point 1,0 due to the xed point 0,0 is given by Tu10 / . The force due to the point 2,0 is equal to T u20 u10 / . The force due to the point 1,1 equals k u10 + u11 . The equation of motion of the point 1,0 follows from Newtons second law. We nd
P. Gravel and C. Gauthier 450

Fig. 1. Schematic of the system of two strings with an elastic band between them.

u10 =

2u10 + u20 + k

u10 + u11 ,

A2

where u10 = d2u10 t / dt2. The equations of motion of the other material points are obtained similarly. We thus arrive at the following system of differential-difference equations, for j = 0 , 1, u1j = T 2u1j + u2j + 1
jk

Fig. 2. Motion in a cross section of a three-rope Inca suspension bridge. The white dot inside the large black dot indicates that the x-axis is perpendicular to the plane of the sheet.

= u10 + u11
jk

A3a +

4 4

u0 + u1
t

k 4 k 4

u0 + u1
x

u0 u1

u0 u1

Y u0 u1 2

u2j =

u1j 2u2j + u3j + 1

u20 + u21 A3b

A4b

] uN1,j = T uN2,j 2uN1,j


jk

APPENDIX B: THREE COUPLED VIBRATING STRINGS As before, we set up a Cartesian coordinate system see Fig. 2 . Assuming the cables to be straight and parallel when at rest, we will use the horizontal x-axis to represent the equilibrium position of the footpath. The y-axis is also horizontal. The z-axis is vertical and the x z-plane passes through the line that is equidistant to the two guard rails at rest. We denote by j = 0 , 1 , 2 the footpath and the two guard rail cables, respectively. The two-component vectors ui;j = uyi;j , uzi;j , j = 0 , 1 , 2, denote small displacements from their equilibrium positions of the cable elements at x = i . These elements are assumed not to be immediate neighbors of the anchoring points, that is, i 0 and i N + 1. For simplicity, the absence of the rst index and the semicolon in the indices of a quantity will be understood to mean that we are using the value of the quantity that corresponds to the plane at x = i for any i = 1 , 2 , . . . , N. We shall use two constant vectors B1 = 1 , 2 and B2 = 1 , 2 of length B in each x = i plane parallel to the y z-plane, and b = by , bz to designate the corresponding unit vectors. The differences between the equilibrium and the displaced position of guard rail k = 1 , 2 and the footpath are Bk = Bk + uk u0 Bk + uk u0
yk , zk

+ uN,j + 1 T

uN1,0 + uN1,1
jk

A3c

uN,j =

uN1,j 2uN,j + 1

uN,0 + uN,1 . A3d

Equations A3a and A3d differ from the others because they include the xed end points of the two strings. Now we let the number of material points approach innity and the common distance between successive points of the same string approach zero on both strings, such that / = and k / are constant. The value of the constant k / can be obtained by observing that it corresponds to Y when = 1 so that k / = Y. In this limit each relation of Eq. A3 , other than the rst and the last, transforms into one of the two partial differential Eqs. 5a and 5b . The dependent variables u0 and u1 are the functions associated with ui0 and ui1, respectively, for i = 1 , 2 , . . . , N. Observe that Eq. 5 can also be deduced from the Lagrangian L=
2 tu 0 + tu 1 2

Bk + uk u0 B .

B1

k 2

2 xu 0 +

xu 1

We let uk u0 = k = order as A4a Bk = Bk +


k

and approximate Eq. B1 to rst B

Y u0 u1 2

B2a
k

Bk +

451

Am. J. Phys., Vol. 79, No. 5, May 2011

P. Gravel and C. Gauthier

451

= Bk + +

1 1
1 yk

ui;1 =
2 zk 2 2

ui1;1 2ui;1 + ui+1;1 +

u0 u1 b b B3b

1 k2

+2
2 1

2 yk

2 1/2 zk

B2b

Bk +

1
k

1 yk + 2 2 1+ 2

2 zk

B2c

ui;2 =

ui1;2 2ui;2 + ui+1;2 +

u0 u2 b+ b+ . B3c

What is left after a rst-order approximation in the s is the projection of k parallel to Bk, namely k bk, where b1 = b and b2 = b+ are the components of the stretching vector of the distance between guard rail k and the footpath. The equations for the forces on each of the three cables are then found to be k T ui1;0 2ui;0 + ui+1;0 ui;0 =
0 0

u0 u1 b b B3a

+ u0 u2 b+ b+

Here and T are the tensions in the guard rails and the footpath, respectively, while and 0 are the linear densities, and 0, multiplied by in the guard rails and the footpath respectively. The terms in and T are the rst-order approximations of the total restoring forces due to the tensions in the three main cables, whose form for the rst guard rail is

ui1;1 2ui;1 + ui+1;1


2

+ uyi1;1 2uyi;1 + uyi+1;1 2 + uzi1;1 2uzi;1 + uzi+1;1

1/2 .

B4

The limits of the differential-difference Eq. B3 as goes to zero while keeping constant / l, 0 / l, and k / lead to a system of three coupled partial differential equations:
2 t u0

L u1 u0 +

A2 u2 u0 + A2

=0

B6c

= C 2 2u 0 A 2 x

u0 u1 b b + u0 u2 b+ b+ B5a

L u2 u0 L u1 u0 L u2 u0
+

A2 u1 u0 2 + A2
2 2

=0

B6d B6e B6f

2 t u1 2 t u2

= =

c 2 2u 1 x c 2 2u 2 x

= =

u1 u0 b b u2 u0 b+ b+ ,

B5b B5c

+ A2 u1 u0

+ A2 u2 u0 + ,

where c = / , C = T / 0, = Y / and A = Y / 0. If the three cables are made of the same material, it is reasonable to assume that the constants c = / and C = T / 0 are very close to one another. In these square roots the tensions serve to keep the bridge more or less straight against gravity, which means that they must be proportional to the linear density to obtain the same shape. At the same time, the linear density increases proportionally to the area of the cross section of the cable. The numbers c and C should thus be approximately the same. Here we shall assume that they are identical. Note that there is no reason why the constants and A should behave the same way. We now write the displacement vectors ui in terms of their components in the b , b+ basis of the planes orthogonal to the three main cables to obtain the following system of six scalar differential equations L u 1+ = 0 L u 2 = 0
452 Am. J. Phys., Vol. 79, No. 5, May 2011

where L = 2 c2 2 is the wave equation operator with speed t x c. These equations are decoupled in terms of the six different scalar arguments of L. As we can see, the quantities u1 u0 and u2 u0 +, which represent the stretching of the secondary cables parallel to their rest orientation, do not obey wave equations, but instead obey KleinGordon equations.

APPENDIX C: THE GOLDSTONE MODEL In the Goldstone model we start with a complex wave function x = 1 x + i 2 x / 2 and the Lagrangian density L x =h x x V , C1

B6a B6b

where V is a potential, x is the complex conjugate of x , and there is a summation over repeated indices using the Minkowski metric h = h = diag 1 , 1 , 1 , 1 . In the simplest case the self-interaction potential of the eld is the quartic function
P. Gravel and C. Gauthier 452

2 2

C2

Lx =

are arbitrary real parameters. The Lawhere 0 and grangian density given by Eq. C1 is invariant under global rotations about the origin of the plane containing the values of x x x x = x = x exp i x exp i
2

1 2 + 1 2 1 4

x x
2

1 2 r2 0 2

x
2

x r0 x
2

x +

x C6

C3a , C3b

x +

0, the minimum of V in which is a real constant. If is unique and located at 0 = 0. This minimum position remains unchanged under the phase transformations of x x in Eq. C3 . The existence of this minimum does and not break the original rotational symmetry of the Lagrangian. The mass associated with the eld x is given by Eq. C2 as the coefcient 2 of the quadratic term. If 2 0, the potential V assumes the shape of a sombrero whose minimum corresponds to the set of complex values 0 of x satisfying
0

r0 2

C4

where r0 = 2 / 2 . This lowest state of energy is innitely degenerate. Any system described by L x will settle around one of these energetically equivalent minima, none of which is invariant under a rotation around the origin in the plane containing the values of . The symmetry of the Lagrangian is said to be spontaneously broken when the system is in one of these minima. If we select the real positive minimum position 0 = r0 / 2, and expand x around 0, we can write x = 1 2 r0 + x +i x , C5

where we have omitted a constant term which is of no consequence. The rst three terms of Eq. C6 correspond to the free Lagrangian density, and the last two terms are associated with the interactions of the elds x and x . We deduce that x corresponds to a massless eld and that x is a eld of the KleinGordon type with mass 2 r2. The eld 0 x describes radial displacements close to the bottom of the valley of the potential, and the eld x represents displacements restricted to the circular bottom of the valley. The free part of the Lagrangian shows that, when quantized, the eld x yields a massless boson, and x yields a boson of mass 2 r2. This process is the Goldstone phenomenon: A 0 continuous global symmetry implies the existence of a massless eld. The similarity between the pairs 1 , 2 and u , v of Sec. VI is clear, as is that of the process through which u and 1 acquire mass while v and 2 remain massless.
a

where x and x are real elds expressing the deviation of x from its state of lowest energy 0. In terms of these new elds, Eq. C1 becomes

Electronic mail: gravel-p@rmc.ca Electronic mail: claude.gauthier@umoncton.ca H. Kragh, Erwin Schrdinger and the wave equation: The crucial phase, Centaurus 26, 154197 1982 . 2 Wolfrang Pauli, in Wissenschaftlicher Briefwechsel, edited by A. Hermann, K. v. Meyenn, and V. F. Weisskopt Springer, New York, 1979 , Vol. I, p. 356. 3 H. Kragh, Equation with the many fathers. The Klein-Gordon equation in 1926, Am. J. Phys. 52, 10241033 1984 . 4 P. M. Morse and H. Feshbach, Methods of Theoretical Physics McGrawHill, New York, 1953 , Vol. I, p. 138. 5 F. Mandl and G. Shaw, Quantum Field Theory John Wiley & Sons, New York, 1984 , p. 43. 6 D. W. Gade, Bridge types in the central Andes, Ann. Assoc. Am. Geogr. 62, 94109 1972 . 7 K. Moriyasu, An Elementary Primer for Gauge Theory World Scientic, Singapore, 1983 , p. 96.
1

453

Am. J. Phys., Vol. 79, No. 5, May 2011

P. Gravel and C. Gauthier

453

Вам также может понравиться