You are on page 1of 25

When action is not least

C. G. Graya兲
Guelph-Waterloo Physics Institute and Department of Physics, University of Guelph, Guelph, Ontario
N1G2W1, Canada
Edwin F. Taylorb兲
Department of Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139
共Received 17 February 2006; accepted 26 January 2007兲
We examine the nature of the stationary character of the Hamilton action S for a space-time
trajectory 共worldline兲 x共t兲 of a single particle moving in one dimension with a general
time-dependent potential energy function U共x , t兲. We show that the action is a local minimum for
sufficiently short worldlines for all potentials and for worldlines of any length in some potentials.
For long enough worldlines in most time-independent potentials U共x兲, the action is a saddle point,
that is, a minimum with respect to some nearby alternative curves and a maximum with respect to
others. The action is never a true maximum, that is, it is never greater along the actual worldline than
along every nearby alternative curve. We illustrate these results for the harmonic oscillator, two
different nonlinear oscillators, and a scattering system. We also briefly discuss two-dimensional
examples, the Maupertuis action, and newer action principles. © 2007 American Association of Physics
Teachers.
关DOI: 10.1119/1.2710480兴

I. INTRODUCTION that the particle will indeed follow. For example, the adjacent
curves need not conserve the total energy. The Hamilton ac-
Several authors1–12 have simplified and elaborated the ac- tion principle says that with respect to all nearby curves the
tion principle and recommended that it be introduced earlier action along the actual worldline is stationary, that is, it has
into the physics curriculum. Their work allows us to see in zero variation to first order; formally we write ␦S = 0.
outline how to empower students early in their studies with
Whether or not this stationary value of the action is a local
the fundamental yet simple extensions of Newton’s prin-
minimum is determined by examining ␦2S and higher order
ciples of motion made by Maupertuis, Euler, Lagrange, Ja-
cobi, Hamilton, and others. The simplicity of the action prin- variations of the action with respect to the nearby curves, as
ciple derives from its use of a scalar energy and time. Its we will discuss in this paper.
transparency comes from the use of numerical11 and Misconceptions concerning the stationary nature of the ac-
analytical12 methods of varying a trial worldline to find a tion abound in the literature. Even Lagrange wrote that the
stationary value of the action, skirting not only the equations value of the action can be maximum,13 a common error14 of
of motion but also the advanced formalism of Lagrange and which the authors of this paper have been guilty.12,15 Other
others characteristic of upper level mechanics texts. The goal authors use extremum or extremal,16 which incorrectly in-
of this paper is to discuss the conditions for which the sta- cludes a maximum and formally fails to include a saddle
tionary value of the action for an actual worldline is a mini- point. 关Mathematicians often use the 共correct兲 term critical
mum or a saddle point. instead of stationary, but because the former term has other
For single-particle motion in one dimension 共1D兲, the meanings in physics we use the latter.兴 A similar error mars
Hamilton action S is defined as the integral along an actual or treatments of Fermat’s principle of optics, which is errone-
trial space-time trajectory 共worldline兲 x共t兲 connecting two ously said to allow the travel time of a light ray between two
specified events P and R, points to be a maximum.17

S= 冕
P
R
L共x,ẋ,t兲dt, 共1兲
The present paper has three primary purposes: First it de-
scribes conditions under which the action is a minimum and
different conditions under which it is a saddle point. These
where L is the Lagrangian, x is the position, t is the time, and conditions involve second variations of the action. Some pio-
P共x P , t P兲 and R共xR , tR兲 are fixed initial and final space-time neers of the second variation theory of the calculus of varia-
events. A dot, as in ẋ, indicates the time derivative. The tions are Legendre, Jacobi, Weierstrass, Kelvin and Tait,
Lagrangian L共x , ẋ , t兲 depends on t implicitly through x共t兲 and Mayer, and Culverwell.18 Although inspired by the early
may also depend on t explicitly, for example, through a time- work of Culverwell,21 our derivation of these conditions is
dependent potential. For simplicity we use Cartesian coordi- new, simpler, and more rigorous; it is also simpler than mod-
nates throughout, but the methods and conclusions apply for ern treatments.24,25 Second, this paper explains the results
generalized coordinates. with qualitative heuristic descriptions of how a particle re-
The Hamilton action principle compares the numerical sponds to space-varying forces derived from the potential in
value of the action S along the actual worldline to its value which it moves. 共Those who prefer immediate immersion in
along every adjacent curve 共trial worldline兲 anchored to the the formalism can begin with Sec. IV.兲 Third, it clarifies
same initial and final events. These alternative curves are these results and illustrates the variety of their consequences
arbitrary as long as they are piecewise smooth, have the by applying them to the harmonic oscillator, two nonlinear
same end events, and are adjacent to 共near to兲 a worldline oscillators, and a scattering system. Criteria used to decide

434 Am. J. Phys. 75 共5兲, May 2007 http://aapt.org/ajp © 2007 American Association of Physics Teachers 434
the nature of the stationary value of the action are also useful
for other purposes in classical and semiclassical mechanics,28
but are not discussed in this paper.
Appendix A adapts the results to the important Maupertuis
action W. Appendix B gives examples of both Hamilton and
Maupertuis action for two-dimensional motion. Appendix C
discusses open questions on the stationary nature of action
for some newer action principles.

II. KINETIC FOCUS


This section introduces the concept of kinetic focus due to
Jacobi,22 which plays a central role in determining the nature
of the stationary action. We start with an analogous example
taken verbatim from Whittaker,36 an analysis of the relative
length along different paths. Whittaker employs the Mauper-
tuis action principle 共discussed in our Appendix A兲, which
requires fixed total energy along trial paths, not fixed travel
time as with the Hamilton action principle. In force-free sys-
tems the value of the Maupertuis action is proportional to the
path length. The term kinetic focus is defined formally later
in this section. Figures 1共a兲 and 1共b兲 illustrate this example.
Whittaker wrote that “A simple example illustrative of the
results obtained in this article is furnished by the motion of a
particle confined to a smooth sphere under no forces. The
trajectories are great-circles on the sphere and the 关Mauper-
tuis兴 action taken along any path 共whether actual or trial兲 is
proportional to the length of the path. The kinetic focus of
any point A is the diametrically opposite point A⬘ on the
sphere, because any two great circles through A intersect
again 共for the first time兲 at A⬘. The theorems of this article
amount, therefore in this case to the statement that an arc of
a great circle joining any two points A and B on the sphere is
the shortest distance from A to B when 共and only when兲 the
point A⬘ diametrically opposite to A does not lie on the arc,
that is, when the arc in question is less than half a great-
circle.”
The elaboration of this analogy is discussed in the captions
of Figs. 1共a兲 and 1共b兲 using equilibrium lengths of a rubber Fig. 1. 共a兲 On a sphere the great circle line ABC starting from the north pole
band on a slippery spherical surface. at A is the shortest distance between two points as long as it does not reach
For a contrasting example, we apply a similar analysis to the south pole at A⬘. On a slippery sphere a rubber band stretched between
free-particle motion in a flat plane. In this case the length of A and C will snap back if displaced either locally, as at D, or by pulling the
the straight path connecting two points is a minimum no entire line aside, as along AEC. The point A⬘ is called the antipode of A or
matter how far apart the endpoints. A rubber band stretched in general the kinetic focus of A. We say that if a great circle path terminates
before the kinetic focus of its initial point, the length of the great circle path
between the endpoints on a slippery surface will always snap
is a minimum. 共b兲 If the great circle ABA⬘G passes through antipode A⬘ of
back when deflected in any manner and released. An alterna- the initial point A, then the resulting line has a minimum length only when
tive straight path that deviates slightly in direction at the compared with some alternative lines. For example on a slippery sphere a
initial point A continues to diverge and does not cross the rubber band stretched along this path will still snap back from local distor-
original path again. Therefore no kinetic focus of the initial tion, as at D. However, if the entire rubber band is pulled to one side, as
point A exists for the original path. along AFG, then it will not snap back, but rather slide over to the portion
Note that on both the sphere and the flat plane there is no AHG of a great circle down the backside of the sphere. With respect to paths
like AFG, the length of the great circle line ABA⬘G is a maximum. With
path of true maximum length between any two separated respect to all possible variations we say that the length of path ABA⬘G is a
points. The length of any path can be increased by adding saddle point. If a great circle path terminates beyond the kinetic focus of its
wiggles. initial point, the length of the great circle path is a saddle point.
How do we find the kinetic focus? In Fig. 1共a兲 we place
the terminal point C at different points along a great circle
path between A and A⬘. When C lies between A and A⬘,
every nearby alternative path such as AEC is not a true path coalesce with the original path ABA⬘. The kinetic focus is
共a path of minimum length兲, because it does not lie along a defined by the existence of this coalescing alternative true
great circle. When terminal point C reaches A⬘, there is sud- path. As the final point C moves away from the initial point
denly more than one alternative great circle path connecting A, the kinetic focus A⬘ is defined as the earliest terminal
A and A⬘. 共In this special case an infinite number of alterna- point at which two true paths can coalesce.
tive great circle paths connect A and A⬘.兲 Any alternative The term kinetic focus in mechanics derives from an
great circle path between A and A⬘ can be moved sideways to analogy37 to the focus in optics, that point A⬘ at which rays

435 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 435
Fig. 2. From the common initial event P we draw a true worldline 0 and a
second true worldline 1 that terminates at some event R on the original Fig. 3. Several true harmonic oscillator worldlines with initial event P
worldline 0. The event nearest to P at which worldline 1 coalesces with = 共0 , 0兲 and initial velocity v0 ⬎ 0. Starting at the initial fixed event P at the
worldline 0 is the kinetic focus Q. origin, all worldlines pass through the same event Q. That is, Q is the kinetic
focus for all worldlines of the family starting at the initial event P = 共0 , 0兲.
Worldlines 1 and 0 differ infinitesimally; worldlines 2 and 0 differ by a finite
amount. This oscillator is discussed in detail in Sec. VIII.
emitted from an initial point A converge under some condi-
tions, such as interception by a converging lens.
This paper deals mainly with the action principle for the A more typical case is the quartic oscillator 共see Fig. 4兲,
Hamilton action S, which determines worldlines in space- which is described by U共x兲 = Cx4. In this case alternative
time for fixed end-events 共that is, end-positions and travel worldlines starting from the initial event P can cross any-
time兲 rather than the action principle for the Maupertuis ac- where along the original worldline 共some crossing events are
tion W 共see Appendix A兲, which determines spatial orbits 共as indicated by little squares in Fig. 4兲. When the alternative
well as space-time worldlines兲 for fixed end-positions and worldline coalesces with the original worldline, the crossing
total energy. The kinetic focus for the Hamilton action has a point has reached the kinetic focus Q.
use similar to that for the examples of the Maupertuis action Another typical case is the piecewise-linear oscillator
in Figs. 1共a兲 and 1共b兲. We will show that a worldline has a shown in Fig. 5. This oscillator has the potential energy
minimum action S if it terminates before reaching the kinetic U共x兲 = C兩x兩. For the piecewise-linear oscillator, as for the
focus of its initial event. In contrast, a worldline that termi- quartic oscillator, alternative worldlines starting from P can
nates beyond the kinetic focus of the initial event P has an cross at various events along the original worldline. Note
action that is a saddle point. that for the piecewise-linear oscillator an alternative world-
We use the label P for the initial event on the worldline 0 line that crosses the original worldline before its kinetic fo-
共see Fig. 2兲, Q for the kinetic focus of P on the worldline, cus lies below the original worldline instead of above it 共as
and R for a fixed but arbitrary event on the worldline that for the quartic oscillator兲. We can equally well use an alter-
terminates on worldline 0 and also terminates another true native worldline that crosses from below or one that crosses
worldline 共#1 in Fig. 2兲 connecting P to R. For the Hamilton from above to define the coalescing worldline and kinetic
action S our definition of the kinetic focus of a worldline is focus.
the following. The kinetic focus Q of an earlier event P on a Notice the gray line labeled caustic in Figs. 4 and 5, and
true worldline is the event closest to P at which a second true also in Fig. 6, which shows the worldlines for a repulsive
worldline, with slightly different velocity at P, intersects the
first worldline, in the limit for which the two worldlines coa-
lesce as their initial velocities at P are made equal.
The kinetic focus is central to the understanding of the
stationary nature of the action S, but its definition may seem
obscure. To preview the consequences of this definition, we
briefly discuss some examples that we will discuss later in
the paper. Figure 3 shows the true worldlines of the harmonic
oscillator, whose potential energy has the form U共x兲
= 共 21 兲kx2. The harmonic oscillator is the single 1D case of the
definition of space-time kinetic focus with the following ex-
ceptional characteristic: every worldline originating at P in
Fig. 3 passes through the same recrossing point. The 2D
spatial paths on the sphere in Figs. 1共a兲 and 1共b兲 show the
same characteristic: Every great circle path starting at A
passes through the antipode at A⬘. In both cases we can find
the kinetic focus without taking the limit for which the ve- Fig. 4. Schematic space-time diagram of a family of true worldlines for the
locities at the initial point are equal and the two worldlines quartic oscillator 关U共x兲 = Cx4兴 starting at P = 共0 , 0兲 with v0 ⬎ 0. The kinetic
focus occurs at a fraction 0.646 of the half-period T0 / 2, illustrated here for
coalesce, but we can take that limit. For the harmonic oscil- worldline 0. The kinetic foci of all worldlines of this family lie along the
lator this limit occurs when the amplitudes are made equal. heavy gray line, the caustic. Squares indicate recrossing events of worldline
The harmonic oscillator will turn out to be the only excep- 0 with the other two worldlines. This oscillator is discussed in detail in Sec.
tion to many of the rules for the action. IX.

436 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 436
Fig. 7. For the Maupertuis action, the heavy line envelope 共the “parabola of
safety”兲 is the locus of spatial kinetic foci 共xQ , y Q兲, or spatial caustic, of the
family of parabolic orbits of energy E originating from the origin
O共x P , y P兲 = 共0 , 0兲 with various directions of the initial velocity v0. The po-
tential is U共x , y兲 = mgy. The horizontal and vertical axes are x and y, respec-
tively, and the caustic/envelope equation is y = v20 / 2g − gx2 / 2v20, found by
Johann Bernoulli in 1692. The caustic divides space. Each final point
共xR , y R兲 inside the caustic can be reached from initial point 共x P , y P兲 by two
Fig. 5. Schematic space-time diagram of a family of true worldlines for a
orbits of the family, each final point on the caustic by one orbit of the family,
piecewise-linear oscillator 关U共x兲 = C 兩 x 兩 兴, with initial event P = 共0 , 0兲 and ini-
4 and each point outside the caustic by no orbit of the family. Y is the vertex
tial velocity v0 ⬎ 0. The kinetic focus Q0 of worldline 0 occurs at 3 of its
共highest reachable point y = v20 / 2g兲 of the caustic and X1, X2 denote the
half-period T0 / 2. Similarly, small circles Q1 and Q2 are the kinetic foci of
maximum range points 共x = ± v20 / g兲. This system is discussed in detail in
worldlines 1 and 2, respectively. The heavy gray curve is the caustic, the
Appendix B. 共Figure adapted from Ref. 43.兲
locus of all kinetic foci of different worldlines of this family 共originating at
the origin with positive initial velocity兲. Squares indicate events at which the
other worldlines recross worldline 0. This oscillator is discussed in detail in
Sec. IX. linear gravitational potential. The word caustic is derived
from optics37 共along with the word focus兲. When a cup of
coffee is illuminated at an angle, a bright curved line with a
potential. The caustic is the line along which the kinetic foci cusp appears on the surface of the coffee 共see Fig. 8兲. Each
lie for a particular family of worldlines 共such as the family of point on this spatial optical caustic or ray envelope is the
worldlines that start from P with positive initial velocity in focus of light rays reflected from a small portion of the cir-
Figs. 4 and 5兲. A caustic is also an envelope to which all cular inner surface of the cup.
worldlines of a given family are tangent. The caustics in In Figs. 4–7 the caustic for a family of worldlines 共or
Figs. 5 and 6 are space-time caustics, envelopes for space- paths兲 represents a limit for those worldlines 共or paths兲. No
time trajectories 共worldlines兲. Figure 7 shows a purely spatial
caustic/envelope for a family of parabolic paths 共orbits兲 in a

Fig. 6. Schematic space-time diagram for the repulsive inverse square po-
tential 关U共x兲 = C / x2兴, with a family of worldlines starting at P共x P , 0兲 with
various initial velocities. Intersections are events where two worldlines
cross. The heavy gray straight line xQ = 共␥ / x P兲tQ, where ␥ = 共2C / m兲1/2, is the
caustic, the locus of kinetic foci Q 共open circles兲 and envelope of the indi-
rect worldlines. Worldline 2, with zero initial velocity, is asymptotic to the
caustic, with kinetic focus Q at infinite space and time coordinates. The
caustic divides space-time: each final event above the caustic can be reached
by two worldlines of this family of worldlines, each final event on the
caustic by one worldline of the family, and each final event below the
caustic by no worldline of the family. This system is discussed in detail in Fig. 8. The coffee-cup optical caustic. The caustic shape in panel 共b兲 共a
Sec. X. nephroid兲 was derived by Johann Bernoulli in 1692 共Ref. 44兲.

437 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 437
worldline of that family exists for final events outside the ity, that is, with a slightly different slope of the worldline.
caustic. At least one worldline can pass through any event Worldline number 1 is also a true worldline. In the limiting
inside the caustic. Exactly one worldline can pass through an case of a vanishing difference in the initial velocities at event
event on the caustic, and this event is the worldline’s kinetic P 共vanishing angle between the initial slopes兲, the two
focus. This observation is consistent with the definition of worldlines will cross again, and the two particles collide at
the kinetic focus as an event at which two separate world- the kinetic focus event Q.
lines coalesce. We can convert this practical 共actually heuristic兲 idea into
At the kinetic focus the worldline is tangent to the caustic. an analytical method, often easily applied when we have an
When two curves touch but do not cross and have equal analytic expression for the worldline. Let the original world-
slope at the point where they touch, the curves are said to line be described by the function x共t , v0兲, where v0 is the
osculate or kiss, which leads to a summary preview of the initial velocity. Then the second worldline is the same func-
results of this paper: tion with incrementally increased initial velocity x共t , v0
+ ⌬v0兲. We form the expansion in ⌬v0
When a worldline terminates before it kisses the
⳵x
caustic, the action is minimum; when the worldline x共t, v0 + ⌬v0兲 = x共t, v0兲 + ⌬v0 + O共⌬v20兲. 共2兲
terminates after it kisses the caustic, the action is a ⳵v0
saddle point. At an intersection point R we have x共tR , v0 + ⌬v0兲 = x共tR , v0兲.
For intersection point R near Q we therefore have
One consequence is that when we use a computer to plot a
family of worldlines 共by whatever means兲, we can eyeball ⳵x
⌬v0 + O共⌬v20兲 = 0, 共3兲
the envelope/caustic and locate the kinetic focus of each ⳵v0
worldline visually.
which implies that for R → Q when ⌬v0 → 0, we have
This summary covers every case but one, because when
no kinetic focus exists, there is no caustic so a worldline of ⳵x
= 0. 共4兲
any length has the minimum action. The one case not cov- ⳵v0
ered by this rule is the harmonic oscillator. For the harmonic
oscillator and also for the sphere geodesics of Fig. 1, the Equation 共4兲 is an analytic condition for the incrementally
caustic collapses to a single point at the kinetic focus. In this different worldline that crosses the original worldline at the
case there is no caustic curve; only one caustic point 共a focal kinetic focus, and hence it locates the kinetic focus Q. Sec-
point兲 exists. A corresponding optical case is a concave re- tions VIII and IX discuss applications of this method.
flecting parabolic surface of revolution illuminated with in-
coming light rays parallel to its axis; the optical caustic col- III. WHY WORLDLINES CROSS
lapses into a single point at the focus 共focal point兲 of the
parabolic mirror. When the optical caustic reduces to a point Section II defines the kinetic focus in terms of recrossing
for a lens or mirror system, the resulting images have mini- worldlines. The burden of this paper is to show that when a
mum distortion 共minimum aberration兲. kinetic focus exists, the action along a worldline is a mini-
For the quartic and the piecewise-linear oscillators 共and mum if it terminates before the kinetic focus Q of the initial
the harmonic oscillator兲 subsequent crossing points exist at event P, whereas the action is a saddle point when the world-
which two worldlines can coalesce. We have defined the ki- line terminates beyond the kinetic focus. In this section we
netic focus as the first of these, the one nearest the initial consider only actual worldlines and describe qualitatively
event P. The procedure for locating the subsequent kinetic why two worldlines originating at the same initial event
foci is identical to that for locating the first one and is dis- cross again at a later event. We also examine the special
cussed briefly in the examples of Secs. VIII and IX. For 1D initial conditions at P under which the coalescing worldlines
potentials subsequent kinetic foci45 exist for the bound determine the position of the kinetic focus Q. The key pa-
worldlines but not for the scattering worldlines, for example, rameter turns out to be the second spatial derivative U⬙
those in Fig. 6. We shall not be concerned with subsequent ⬅ ⳵2U / ⳵x2 of the potential energy function U. Sufficiently
kinetic foci; when we refer to the kinetic focus we mean the long worldlines can cross again only if they traverse a space
first one, as we have defined it. We shall show in what fol- in which U⬙ ⬎ 0. For simplicity we restrict our discussion
lows that for a few potentials 关for example, U共x兲 = C and here to time-independent potentials U共x兲, but continue to use
U共x兲 = Cx兴 kinetic foci do not exist, because true worldlines partial derivatives of U with respect to x to remind ourselves
beginning at a common initial event P do not cross again. of this restriction. Features that can arise for time-dependent
The definition of kinetic focus in terms of coalescing potentials U共x , t兲 are discussed in Sec. XI.
worldlines provides a practical way to find the kinetic focus. Think of two identical particles that leave initial event P
In Fig. 2 note the slopes of nearby worldlines 1 and 0 at the with different velocities and hence different slopes of their
initial event P. The initial slope of curve 1 is only slightly space-time worldlines, so that their worldlines diverge. The
different from that of worldline 0; as that difference ap- following description is valid whether the difference in the
proaches zero, the crossing event approaches the kinetic fo- initial slopes is small or large. Figures 3–5 illustrate the fol-
cus Q. The slope of a worldline at any point measures the lowing narrative. At every event on its worldline each par-
velocity of the particle at that point. This coalescence of two ticle experiences the force F = −U⬘ = −⳵U / ⳵x evaluated at that
worldliness as their initial velocities approach each other location. For a short time after the two particles leave P they
leads to a method for finding the kinetic focus: Launch an are at essentially the same displacement x, so they feel nearly
identical second particle from event P 共simultaneously with the same force −U⬘. Hence the space-time curvature of the
the original launch兲 but with a slightly different initial veloc- two worldlines 共the acceleration兲 is nearly the same. There-

438 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 438
fore the two worldlines will initially curve in concert while
their initial relative velocity carries them apart; at the begin-
ning their worldlines steadily diverge from one another. As
time goes by, this divergence carries one particle, call it II,
into a region in which the second spatial derivative U⬙ is 共let
us say兲 positive. Then particle II feels more force than par-
ticle I 共but still in the same x direction as the force on it兲. As
a result the worldline of II will head back toward particle I,
leading to converging worldlines. As the two particles draw
near again, they are once more in a region of almost equal U⬘
and therefore experience nearly equal acceleration, so their Fig. 9. An original true worldline, labeled 0, starts at initial event P. We
relative velocity remains nearly constant until the worldlines draw an arbitrary adjacent curve, labeled 1, anchored at two ends on P and
intersect, at which event the two particles collide. a later event R on the original worldline. The variational function ␣␾ is
Note the crucial role played by the positive value of U⬙ in chosen to vanish at the two ends P and R.
the relative space-time curvatures of worldlines I and II nec-
essary for them to recross. Suppose instead that U⬙ ⬍ 0. Then
as II moves away from I, it enters a region of smaller slope point. This later crossing point is defined as the kinetic focus
U⬘ and hence smaller force than that on particle I. Hence the for the Maupertuis action W applied to spatial orbits 共see
two worldlines will diverge even more than they did origi- Appendix B兲. This reconvergence has important conse-
nally; the more they separate, the greater will be their rate of quences for the stability of orbits and the continuing survival
divergence. As long as both particles move in a region where of life on Earth as our planet experiences small nudges from
U⬙ ⬍ 0, the two worldlines will never recross. 共If U⬙ = 0, the the solar wind, meteor impacts, and shifting gravitational
two worldlines continue indefinitely to diverge at the initial forces from other planets.
rate.兲
As a special case let the relative velocity of the two par-
ticles at launch be only incrementally different from one an- IV. VARIATION OF ACTION FOR AN ADJACENT
other 共Fig. 2兲 for motion in potentials with U⬙ ⬎ 0, and let CURVE
this difference of initial velocities approach zero. In this limit
the particles will by definition collide at the kinetic focus Q “…another feature in classical mechanics that
of the initial event P. It may seem strange that an incremen- seemed to be taboo in the discussion of the varia-
tal relative velocity at P results in a recrossing at Q at a tional principle of classical mechanics by physi-
significant distance along the worldline from P. We might cists: the second variation…” Martin Gutzwiller46
think that as this difference in slope increases from zero, the
recrossing event would start at P and move smoothly away The action principle says that the worldline that a particle
from it along worldline I, not “snap” all the way to Q. The follows between two given fixed events P and R has a sta-
source of the snap lies in the first and second spatial deriva- tionary action with respect to every possible alternative ad-
tives of U. When both particles start from the same initial jacent curve between those two events 共Fig. 9兲. Thus the
event, the first derivative at essentially the same displace- action principle employs not only actual worldlines, but also
ment leads to nearly the same force −U⬘ on both particles, so freely imagined and constructed curves adjacent to the origi-
that any difference in the initial velocity, no matter how nal worldline, curves that are not necessarily worldlines
small, continues increasing the separation. It is only with themselves. In this paper the word worldline 共or for empha-
greater relative displacement over time that the difference in sis true worldline兲 refers to a space-time trajectory that a
these forces, quantified by U⬙ ⬎ 0, deflects the two world- particle might follow in a given potential. The word curve
lines back toward one another, leading to eventual recross- means an arbitrarily constructed trajectory that may or may
ing. No alternative true worldline starting at P and with neg- not be a worldline. To study the action we need curves as
ligibly different initial velocity crosses the original worldline well as worldlines. 共In the literature the terms actual, true,
earlier than its kinetic focus 共though widely divergent world- and real trajectory are used synonymously with our term
lines may cross sooner, as shown in Figs. 4 and 5兲. One worldline; the terms virtual and trial trajectory are used for
consequence of this result is that a worldline terminating our use of the term curve.兲
before its kinetic focus has minimum action, as shown ana- In this section we investigate the variational characteristics
lytically in Sec. VI. of the action S of a worldline in order to determine whether
In other words, potentials with U⬙共x兲 ⬎ 0 are stabilizing, S is a local minimum or a saddle point with respect to arbi-
that is, they bring together neighboring trajectories that ini- trary nearby curves between the same fixed events. In Fig. 9
tially slightly diverge. Potentials with U⬙共x兲 ⬍ 0 are destabi- a true worldline labeled 0 and described by the function x0共t兲
lizing, that is, they push further apart neighboring trajectories starts at initial event P. We construct a closely adjacent ar-
that initially slightly diverge. It is thus not surprising that bitrary curve, labeled 1 and described by the function x共t兲,
trajectory stability is closely related to the character of the which starts at the same initial event P and terminates at a
stationary trajectory action 共saddle point or minimum兲.23,29–31 later event R on the original worldline. To compare the ac-
Planetary orbits also exhibit crossing points distant from tion along P0R on the worldline x0共t兲 with the action along
the location of a disturbance; an incremental change in the P1R on the arbitrary adjacent curve x共t兲, let
velocity at one point in the orbit leads to initial and contin- x共t兲 = x0共t兲 + ␣␾共t兲, 共5兲
ued divergence of the two orbits which, for certain potential
functions, reverses to bring them together again at a distant and take the time derivative,

439 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 439
d 3L 3⳵ U
3
ẋ共t兲 = ẋ0共t兲 + ␣␾
˙ 共t兲. 共6兲 = − ␾ . 共16兲
d␣3 ⳵x3
In Eqs. 共5兲 and 共6兲 ␣ is a real numerical constant of small
absolute value and ␾共t兲 is an arbitrary real function of time The expansion 共7兲 for L defined by Eq. 共11兲 now becomes

冉 冊 冉 冊
that goes to zero at both P and R. The action principle says
⳵U ˙ + ␣ − ␾2 ⳵ U + m␾
2 2
that action in Eq. 共1兲 along x0共t兲 is stationary with respect to
L = L0 + ␣ − ␾ + mẋ0␾ ˙2
the action along x共t兲 for small values of ␣. To simplify the ⳵x 2 ⳵x2
analysis we restrict ␾共t兲 to be a continuous function with at
most a finite number of discontinuities of the first derivative;
that is, all curves x共t兲 are assumed to be at least piecewise
+
␣3
6
冉 ⳵ 3U
− ␾3 3 + ¯ ,
⳵x
冊 共17兲

smooth.47 Within this limitation x共t兲 represents all possible where the U derivatives are evaluated along x0共t兲. If we sub-
curves adjacent to x0共t兲, not only any actual nearby world- stitute Eq. 共17兲 into the action integral 共1兲, we obtain an
lines. From Eqs. 共5兲 and 共6兲 the Lagrangian L共x , ẋ , t兲 can be expansion of S in powers of ␣:
regarded as a function of ␣ and hence expanded in powers of
␣ for small ␣, S = S 0 + ␦ S 0 + ␦ 2S 0 + ␦ 3S 0 + ¯ . 共18兲
48
dL ␣2 d2L ␣3 d3L The standard result of the action principle is that along a
L = L0 + ␣ + + + ¯ , 共7兲 true worldline the action is stationary; that is, the term ␦S0 in
d␣ 2 d␣2 6 d␣3 Eq. 共18兲, called the first order variation 共or simply the first
where L0 = L共x0 , ẋ0 , t兲 and the derivatives are evaluated at ␣ variation兲, is zero for all variations around an actual world-
= 0, that is, along the original worldline x0共t兲. We apply Eqs. line x0共t兲. 关This condition is necessary and sufficient for the
共5兲 and 共6兲 to the first derivative in Eq. 共7兲, validity of Lagrange’s equation of motion for x0共t兲.兴 We will
need the higher-order variations ␦2S0 and ␦3S0 for an actual
dL ⳵L dx ⳵L dẋ ⳵L ˙ ⳵L worldline. From Eqs. 共17兲 and 共18兲 they take the forms
=␾ +␾ 共8兲

冕冉 冊
= + ,
d␣ ⳵x d␣ ⳵ẋ d␣ ⳵x ⳵ẋ ␣2 R
⳵ 2U
␦ 2S = − ␾2 + m␾
˙ 2 dt, 共19兲
so that we can write in operator form 2 P ⳵x2
d ⳵ ˙ ⳵ and
=␾ +␾ , 共9兲
d␣ ⳵x

⳵ẋ
␣3 R
⳵ 3U
␦ 3S = − ␾3 dt, 共20兲
and apply it twice in succession to yield 6 ⳵x3

冉 冊冉 冊
P
d 2L ⳵ ˙ ⳵ ⳵L ˙ ⳵L
= ␾ +␾ ␾ +␾ where the derivatives of U are evaluated along x0共t兲. In Eqs.
d␣2 ⳵x ⳵ẋ ⳵x ⳵ẋ 共19兲 and 共20兲 and for most of what follows we use the com-
pact standard notations ␦2S = ␦2S0 and ␦3S = ␦3S0 共as well as
⳵ 2L ˙ ⳵ L +␾˙ 2⳵ L .
2 2
= ␾2 2 + 2␾␾ 共10兲 ␦S = ␦S0兲.
⳵x ⳵x⳵ẋ ⳵ẋ2 In the remainder of this article we use Eqs. 共19兲 and 共20兲
We consider the most common case in which the Lagrang- to determine when the action is greater or less for a particular
ian L is equal to the difference between the kinetic and po- adjacent curve than for the original worldline, paying pri-
tential energy: mary attention to the second order variation ␦2S. When the
action is greater for all adjacent curves than for the world-
L = K − U = 21 mẋ2 − U共x,t兲, 共11兲 line, ␦2S ⬎ 0 and the action along the worldline is a true
minimum. The phrase “for all adjacent curves” means that
where U may be time dependent. Then L has the partial the value of ␦2S in Eq. 共19兲 is positive for all possible varia-
derivatives tions ␣␾共t兲. Equation 共19兲 shows immediately that when
⳵L ⳵U ⳵L ⳵2U / ⳵x2 is zero or negative along the entire worldline, then
=− , = mẋ, 共12兲 the integrand is everywhere positive, leading to ␦2S ⬎ 0.
⳵x ⳵x ⳵ẋ Hence, if ⳵2U / ⳵x2 艋 0 everywhere, a worldline of any length
has minimum action. This result was previewed in the quali-
⳵ 2L ⳵ 2U ⳵ 2L ⳵ 2L tative argument of Sec. III.
= − , = 0, = m, 共13兲 The outcome is more complicated when ⳵2U / ⳵x2 is neither
⳵x2 ⳵x2 ⳵x⳵ẋ ⳵ẋ2
zero nor negative everywhere along the worldline. We show
in Sec. V that even in this case we have ␦2S ⬎ 0 for suffi-
⳵ 3L ⳵ 3U ⳵ 3L
= − , = 0. 共14兲 ciently short worldlines, so that the action is still a minimum.
⳵x3 ⳵x3 ⳵ẋ3 Later sections show that “sufficiently short” means a world-
line terminated before the kinetic focus. For a worldline ter-
Hence the second ␣ derivative of L reduces to minated beyond the kinetic focus, the action is smaller
共␦2S ⬍ 0兲 for at least one adjacent curve and greater 共␦2S
d 2L 2⳵ U
2
= − ␾ + m␾
˙ 2. 共15兲 ⬎ 0兲 for all other adjacent curves, a condition called a saddle
d␣2 ⳵x2
point in the action. When the action is a saddle point, the
We apply Eq. 共9兲 to Eq. 共15兲 to obtain the third derivative value of ␦2S in Eq. 共19兲 is negative for at least one variation
of L for this case: ␣␾共t兲 and positive for all other variations ␣␾共t兲.

440 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 440
When ␦2S = 0 for one or more adjacent curves, as happens
at a kinetic focus,49 we need to examine the higher-order
variations to see whether S − S0 is positive, negative, or zero
␾共t兲 = 冕 P
t
␾˙ 共t⬘兲dt⬘ 艋 共t − t P兲␾˙ max ⬍ T␾˙ max , 共22兲

for these particular adjacent curves.


There is no worldline whose action is a true maximum, where T = tR − t P, and ␾
˙ max is the maximum value between P
that is, for which ␦2S ⬍ 0 or more generally for which S and R. With this substitution the magnitude of the first inte-
− S0 ⬍ 0 for every adjacent curve. The following intuitive gral in Eq. 共21兲 for ␦2S can be bounded:
proof by contradiction was given briefly by Jacobi22 and in
more detail by Morin50 for the Lagrangian L = K − U with K
positive as in Eq. 共11兲. Consider an actual worldline for
冏冕 P
R
共− ␾2U⬙兲dt ⬍T3␾ 冏
˙ 2 兩U⬙ 兩.
max max 共23兲

which it is claimed that S in Eq. 共1兲 is a true maximum. Now The second integral in Eq. 共21兲 can be rewritten as
modify this worldline by adding wiggles somewhere in the
middle. These wiggles are to be of very high frequency and
very small amplitude so that they increase the kinetic energy 冕P
R
m␾
˙ 2 dt = mT具␾
˙ 2典, 共24兲
K compared to that along the original worldline with only a
small change in the corresponding potential energy U. The where 具␾
˙ 2典 is the mean square of ␾
˙ over the time interval T.
Lagrangian L = K − U for the region of wiggles is larger for
the new curve and so is the overall time integral S. The new Compare Eqs. 共23兲 and 共24兲 and note that ␾˙ 2 and 具␾˙ 2典 have
max
worldline has greater action than the original worldline, the same order of magnitude for all values of T; the reader
which we claimed to have maximum action. Therefore S can check the special case ␾共t兲 = A sin共n␲共t − t P兲 / T兲, where n
cannot be a true maximum for any actual worldline. is any nonzero integer. Here we assume for simplicity that
␾共t兲 is smooth and is nonzero for all times t in the range T
except possibly at discrete points; a similar argument can be
given if this condition is violated. Also note that 兩Umax ⬙ 兩 will
V. WHEN THE ACTION IS A MINIMUM not increase as R becomes closer to P. Thus if the range is
sufficiently small, the most important term in ␦2S is the one
We now employ the formalism of Sec. IV to analyze the
action along a worldline that begins at initial event P and that contains ␾˙ . In this limit Eq. 共21兲 reduces to
terminates at various final events R that lie along the world-
line farther and farther from P. In this section we show that
the action is a minimum for a sufficiently short worldline PR
␦ 2S → ␣ 2m
1
2
冕 P
R
␾˙ 2dt

in all potentials, and we give a rough estimate of what suf- ⬎ 0 for sufficiently short worldlines. 共25兲
ficiently short means. 共We showed in Sec. IV that the action
is a minimum for all worldlines in some potentials.兲 In Sec. This quantity is positive because m, ␣2, and ␾ ˙ 2 are all
VI we show that sufficiently short means before the terminal positive.51 Therefore ␦2S adds to the action, which demon-
event reaches the event R at which ␦2S first vanishes for a strates that the action is always a true minimum along a
particular, unique variation. We also will show that this R is sufficiently short worldline. We shall use this result repeat-
Q, the kinetic focus of the worldline. In Sec. VII we show edly in the remainder of this paper.
that conversely ␦2S must vanish at the kinetic focus, and that We can give a rough estimate55 of the largest possible
when final event R is beyond Q, the action along PR is a value of T such that ␦2S ⬎ 0 for all variations 共the exact value
saddle point. is given in Sec. VI兲. If we use Eqs. 共23兲 and 共24兲 in Eq. 共21兲,
In considering different locations of the terminal event R we see that
along the worldline, it is important to recognize that the set
of incremental functions ␾ that go to zero at P and at R will ␣2 ˙ 2 兩U⬙ 兩T3兴,
␦ 2S ⬎ 关mT具␾
˙ 2典 − ␾
max max 共26兲
be different for each terminal position R. Particular functions 2
may have similar forms for all R; for example, assuming t P
= 0 for simplicity, we might have ␾ = A共t / tR兲共1 − t / tR兲 or ␾ so that ␦2S ⬎ 0 if
= A sin共␲t / tR兲. However, ␾ need not be so restricted; the mT具␾ ˙ 2 兩U⬙ 兩T3 ,
˙ 2典 ⬎ ␾ 共27兲
only restrictions are that ␾ go to zero at both P and R and be max max

piecewise smooth. Statements about the value of ␦2S for or


each different terminal event R are taken to be true for all
possible ␾ for the particular R that satisfy these conditions. ␾˙ rms T0
T⬍ , 共28兲
For a sufficiently short worldline the action is always a ␾˙ max 2␲
minimum compared with that of adjacent curves, as men-
tioned in Sec. III. The formalism developed in Sec. IV con- ⬙ 兩兲1/2 and ␾˙ rms ⬅ 具␾˙ 2典1/2 is the root-
where T0 / 2␲ = 共m / 兩Umax
firms this result as follows. 共Here we follow and elaborate
Whittaker,36 apart from a qualification given in the follow- mean-square value of ␾ ˙.
ing.兲 We rewrite Eq. 共19兲 using U⬙共x兲: For the harmonic oscillator T0 is exactly equal to the pe-
riod, and for a general oscillator T0 is a time of the order of
the period. Assume for simplicity that ␾共t兲 is smooth and is
␦ 2S = −
␣2
2

P
R
␾2U⬙ dt +
␣2
2

P
R
m␾
˙ 2 dt. 共21兲 nonvanishing over the whole range T with exceptions only at
discrete points, for example, ␾共t兲 = A sin共n␲共t − t P兲 / T兲; a
similar argument can be constructed if this condition is vio-
Because ␾ = 0 at P, we can write lated. The ratio ␾˙ rms / ␾
˙ max is then of order unity; for ex-

441 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 441
ample, for the latter form of ␾共t兲 we have ␾ ˙ rms / ␾
˙ max

= 1 / 2. Thus for times T less than about 共1 / ␲兲T0 / 2 we have
␦2S ⬎ 0. For the various oscillators studied in Secs. VIII and
IX we will see that the half-period is a better estimate of the
time limit for which ␦2S ⬎ 0. For example, for the harmonic
oscillator we show that for all times up to exactly one half-
period, ␦2S ⬎ 0, so that the action is a minimum for times
less than a half-period. We show in Sec. VI that for any
system the precise time limit for which ␦2S ⬎ 0 for all varia-
tions is tQ − t P, the time to reach the kinetic focus.
Because the location of the initial event P is arbitrary, it
follows that the action is a minimum on a short segment
anywhere along a true worldline. It is not difficult to show Fig. 10. Let R be the earliest event along the true worldline x0共t兲, labeled 0,
that a necessary and sufficient condition for a curve to be a such that ␦2S = 0 for worldline PR along x0共t兲. The unique variational func-
true worldline is that all short segments have minimum ac- tion achieving ␦2S = 0 for worldline 0 corresponds to a varied curve labeled
tion. This result is valid irrespective of whether the action for 1. We show that for this location of R, curve 1 is a true worldline and R is
the kinetic focus Q. The arbitrary curve 2 is used to verify that curve 1 is a
the complete worldline is a minimum or a saddle point. true worldline.
As discussed in Sec. IV, if U⬙共x兲 is zero or negative at
every x along the worldline, then ␦2S in Eq. 共19兲 is always
positive, with the result that worldlines of every length have
minimum action for particles in these potentials. For ex- well and Whittaker focus on the Maupertuis action W. We
ample, the gravitational potential energy functions U1 for adapt their work to the Hamilton action S, extend and sim-
vertical motion near Earth’s surface and U2 for radial motion plify it, and show that their argument is incomplete.
above the Earth 共radius rE兲 have the standard forms Consider a true worldline x0共t兲, labeled 0 in Fig. 10. As we
have seen in Sec. V, the action S0 along a sufficiently short
U1共x兲 = mgx 共0 艋 x Ⰶ rE兲, 共29兲 segment PR⬘ of worldline 0 is a minimum, which leads to
␦2S ⬎ 0 for all variations. We imagine terminal event R⬘ lo-
U2共x兲 = − 共GMm兲/共rE + x兲 共0 艋 x兲. 共30兲 cated at later and later positions along the worldline until it
In both cases U⬙共x兲 is zero or negative everywhere, so that reaches R, the event at which, by hypothesis, ␦2S → 0 for the
Eq. 共19兲 tells us that worldlines of any length have minimum first time for some variation; that is, the integral in Eq. 共19兲
action. 共Further discussion of the nature of the stationary defining ␦2S vanishes for some choice of ␾. We shall find
action for trajectories in these gravitational potentials ap- that the earliest event at which ␦2S vanishes is connected to
pears in Appendix B; differences arise when we examine the initial event P by a unique type of variation, namely a
two-dimensional trajectories and when we compare the true worldline that coalesces with the original worldline in
Hamilton action S with the Maupertuis action W.兲 the limit for which their initial velocities at P coincide.
There are an infinite number of potential energy functions Hence the earliest event at which ␦2S vanishes satisfies the
with the property U⬙共x兲 艋 0 everywhere 关another example is definition of the kinetic focus Q.
As we assumed, ␦2S ⬎ 0 for all R⬘ up to R⬘ = R, the first
U共x兲 = −Cx2兴, leading to minimum action along worldlines of
event for which ␦2S = 0 for a particular variation ␣␾. To
any length. Nevertheless, the class of such functions is small prove that R is the kinetic focus Q, we need to consider small
compared with the class of potential energy functions for variations because Q involves a coalescing second worldline.
which U⬙共x兲 ⬎ 0 everywhere 共such as the harmonic oscillator In the typical case the integral in Eq. 共20兲 defining ␦3S does
potential U共x兲 = kx2 / 2兲, or for which U⬙共x兲 is positive for not vanish, but letting ␣ → 0 will ensure ␦3S 共proportional to
some locations and zero or negative for other locations 关such ␣3兲 does not exceed ␦2S 共proportional to ␣2兲 for R⬘ ap-
as the Lennard-Jones potential U共x兲 = C12 / x12 − C6 / x6兴. For proaching R.56 The fact that ␦3S does not exceed ␦2S keeps
this larger class of potentials the particular choice of world- S − S0 ⬎ 0 共not just ␦2S ⬎ 0兲 for R⬘ up to R. 关The untypical
line 共length and location兲 determines whether the action has case, in which the integral in Eq. 共20兲 vanishes, is discussed
a minimum or whether it falls into the class for which the in the following.兴 In the limit R⬘ → R, we let ␣ → 0 so that
action is a saddle point. the varied curve x1共t兲 = x0共t兲 + ␣␾共t兲 coalesces with the true
worldline x0共t兲, and S − S0 = 0 at R⬘ = R.
VI. MINIMUM ACTION WHEN WORLDLINE To satisfy the definition of the kinetic focus, we need to
show that x1共t兲 is a true worldline just short of the limit ␣
TERMINATES BEFORE KINETIC FOCUS
→ 0 共just short of the limit R⬘ → R兲, not just at the limit ␣
The central result of this section is that the event along a = 0. We will prove this result by contradiction: assume curve
worldline nearest to initial event P at which ␦2S goes to zero x1共t兲 共curve 1 in Fig. 10兲 is not a true worldline. Consider the
is the kinetic focus Q. The key idea is that a unique true arbitrary comparison curve 2 in Fig. 10, which differs from 1
worldline connects P to Q, a worldline that coalesces with by the arbitrary variation ␣2␾2, with ␣2 small. Assume
the original worldline and thus satisfies the definition of the 共wrongly兲 that curve 1 is not a true worldline, so that the
kinetic focus in Sec. II. The primary outcome of this proof is first-order variation ␦S in S between curves 1 and 2 is non-
that the action is a minimum if a worldline terminates before zero for arbitrary ␣2␾2, and the sign of ␣2 can be chosen to
reaching its kinetic focus. make S2 ⬍ S1. But because S1 = S0 to second order, we must
Our discussion is inspired by the classic work of have S2 ⬍ S0, which is a contradiction; R is the earliest zero
Culverwell21 共see Ref. 52 for a textbook discussion兲. Culver- of S − S0 for any small variation, so that small variations giv-

442 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 442
ing S − S0 ⬍ 0 are impossible. To avoid the contradiction,
curve 1 must be a true worldline. Thus we have proven that
the unique variation ␣␾ that connects P and R = Q when ␦2S
goes to zero for the first time corresponds to a true worldline.
This argument covers the typical case, where the integral
in Eq. 共20兲 defining ␦3S does not vanish. The only common
untypical case49,57 is the harmonic oscillator, which we will
discuss in Sec. VIII. The harmonic oscillator potential U
= kx2 / 2, for which ⳵3U / ⳵x3 in Eq. 共20兲 vanishes, so that
␦3S = 0 identically. Similarly, the variation ␦4S and higher
variations all vanish because ⳵4U / ⳵x4 and higher potential
derivatives vanish. Thus for the harmonic oscillator ␦2S = S
− S0 and S − S0 remains positive up to R⬘ = R for arbitrary ␣
共not just small ␣兲. The preceding argument with ␣ → 0 is
valid also for the harmonic oscillator, so that the coalescing
true worldline at R again shows that R is the kinetic focus Q.
However, it is not necessary here to take the limit ␣ → 0.
Figure 3 for the harmonic oscillator shows that all true
worldlines beginning at P intersect again where ␦2S first van-
ishes, which is the kinetic focus Q. By varying the amplitude
of the alternative true worldlines for the harmonic oscillator,
we can find one that coalesces with the original worldline
and thus satisfies the definition of the kinetic focus. The ar-
gument for other untypical cases49 is similar to that for the Fig. 11. Schematic illustration of the topological evolution of the minimum
typical case. A → trough B → saddle C of the action S for two “directions” in function
In summary we have shown that as terminal event R takes space as the final time tR increases from tR ⬍ tQ to tR = tQ to tR ⬎ tQ, respec-
tively, where tQ is the kinetic focus time. 共Figure adapted from Ref. 59.兲
up positions along the worldline farther away from the initial
point P, the special varied curve that leads to the earliest
vanishing of ␦2S is typically a unique49,58 true worldline that
can coalesce with the original worldline. This R satisfies the
definition of the kinetic focus Q. Because the varied world-
line for which ␦2S = 0 for the first time is unique, it follows A simple topological picture of the action landscape in
that all other curves PQ adjacent to the original worldline function space is emerging 共see Fig. 11兲. For a short world-
have ␦2S ⬎ 0. line PR, the action S is a minimum: the action increases in
For bound motion in a time-independent potential, world- all directions away from the stationary point in function
lines that can coalesce will typically cross more than once. In space 关panel 共a兲 in Fig. 11兴. For longer PR, we may reach a
the literature all of these sequential limiting crossings are kinetic focus R = Q for which S is trough-shaped, that is, flat
called kinetic foci. The above argument is valid only for the in one special direction and increasing in all other directions
first such crossing, which we refer to as the kinetic focus. away from the stationary point 关panel 共b兲 in Fig. 11兴. 共The
As we have shown, a sufficient condition for the kinetic trough is completely flat for the harmonic oscillator, and flat
focus Q is that it is the earliest terminal event R for which to at least second order49 for all other systems.兲 As we shall
␦2S = 0. In the following section we show the converse nec- see in Sec. VII, as R moves beyond Q, the trough bends
essary condition: Given the definition of the kinetic focus Q
downward, placing the action at a saddle point; that is, S
as the first event at which a second true worldline can coa-
lesce with PQ, the necessary consequence is that ␦2S = 0 for decreases in one direction in function space and increases in
worldline PQ for the variation leading to coalescence. Taken all other directions away from the stationary point 关panel 共c兲
together, the arguments in these two sections prove the fol- in Fig. 11兴. Although not discussed in this paper, the pattern
lowing theorem, which is the fundamental analytical result of may continue as R moves still further beyond the kinetic
our paper: focus Q. If R reaches a second event Q2 at which ␦2S = 0
共called the second kinetic focus in the literature兲, a trough
A necessary and sufficient condition for Q to be a again develops for one special variational function ␾ 共differ-
kinetic focus of worldline PQ is that Q is the ear- ent from the first special ␾兲; at Q2 the action is flat in one
liest event on the worldline for which ␦2S = 0. direction in function space, decreases in one direction, and
increases in all other directions. Beyond Q2 the trough be-
This earliest vanishing of ␦2S occurs for one special type of comes a maximum, and we have a saddle point that is a
variation ␣␾ 共which turns out to correspond to a true world- maximum in two directions and a minimum in all others.
line兲, with ␾ unique 共up to a factor兲 and typically ␣ → 0; for Similar topological changes occur if we reach still later ki-
all other variations ␦2S remains positive at the kinetic focus. netic foci events Q3, Q4, and so forth, a result in agreement
The Culverwell-Whittaker argument 共more complicated than with Morse’s theorem,34 which states that the number of di-
that just given兲 is incomplete in that it addresses only the rections n in function space for which the action is a maxi-
sufficiency part of the theorem 共the necessary part given in mum at a saddle point for worldline PR is equal to the num-
Sec. VII is new兲, and it overlooks the usual case where the ber n of kinetic foci between the end events P and R of the
limit ␣ → 0 is necessary to locate the kinetic focus. worldline.

443 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 443
regard x0共t兲 as a varied curve of x1共t兲, because it is closely
adjacent to the other true worldline x1共t兲. Equations similar
to Eqs. 共5兲 and 共6兲 are
x1共t兲 = x0共t兲 + ␣␾共t兲, x0共t兲 = x1共t兲 − ␣␾共t兲, 共31兲
and

ẋ1共t兲 = ẋ0共t兲 + ␣␾
˙ 共t兲, ẋ0共t兲 = ẋ1共t兲 − ␣␾
˙ 共t兲. 共32兲
We have
S 1 = S 0 + ␦ S 0 + ␦ 2S 0 + ␦ 3S 0 + ¯ . 共33兲
Fig. 12. By definition the kinetic focus Q of the initial event P is the first In this case both x0共t兲 and x1共t兲 are true worldlines, so that
event at which two adjacent true worldlines x0共t兲 and x1共t兲 coalesce. We we can also write the inverse expression
show that ␦2S = 0 at the kinetic focus Q for this variational function ␾ in the
limit R1 → Q, and that the action is a saddle point when the terminal event S 0 = S 1 + ␦ S 1 + ␦ 2S 1 + ␦ 3S 1 + ¯ . 共34兲
R2 lies anywhere on the worldline beyond the kinetic focus Q. 共The lower
␣␾ has the opposite sign from the upper ␣␾, and the upper and lower We subtract Eq. 共34兲 from Eq. 共33兲 and use ␦S0 = 0 and ␦S1
functions ␾ are slightly different due to the end-events R1 and R⬘1 being = 0, because both x0共t兲 and x1共t兲 are true worldlines. The
slightly different.兲
result is
2共S1 − S0兲 = 共␦2S0 − ␦2S1兲 + 共␦3S0 − ␦3S1兲 + ¯ . 共35兲
We then find expressions for ␦2S from Eq. 共19兲:
VII. SADDLE POINT IN ACTION WHEN
WORLDLINE TERMINATES BEYOND KINETIC
FOCUS
␦ 2S 0 =
␣2
2
冕P
R
共− ␾2U⬙共x0兲 + m␾
˙ 2兲dt, 共36a兲

In Sec. VI we showed that a sufficient condition for the


earliest event to be the kinetic focus Q is that the earliest
␦ 2S 1 =
␣2
2
冕P
R
共− ␾2U⬙共x1兲 + m␾
˙ 2兲dt. 共36b兲
event at which ␦2S = 0 is connected to the initial event P by a
unique true coalescing worldline. All alternative curves PQ The first parenthesis on the right side of Eq. 共35兲 has the
lead to ␦2S ⬎ 0. In this section we demonstrate the corre- form


sponding necessary condition, namely, given an alternative
true worldline between P and R that coalesces with the origi- ␣2 R
␦ 2S 0 − ␦ 2S 1 = dt关U⬙共x1兲 − U⬙共x0兲兴␾2 , 共37兲
nal worldline as R → Q and therefore defines Q as the kinetic 2 P
focus, we have ␦2S = 0 for this worldline. By using an exten-
sion of this analysis, we also show that when R lies beyond where terms in ␾
˙ have cancelled. We expand U⬙共x1兲 to first
the kinetic focus Q the action of worldline PQR is a saddle order in ␣:
point.
The essence of the proof of the first statement is outlined U⬙共x1兲 ⬇ U⬙共x0兲 + U⵮共x0兲共x1 − x0兲 = U⬙共x0兲 + U⵮共x0兲␣␾ .
in the following heuristic argument by Routh.29 Consider two 共38兲
intersecting true worldlines P → R connecting P to R. As-
sume R is close to the kinetic focus Q so that the two world- If Eq. 共38兲 is substituted into Eq. 共37兲, the resulting integral
lines differ infinitesimally, as required in the definition of Q. contributes a further factor of ␣, yielding a result of O共␣3兲:
Let the action along the two worldlines be S and S + ␦S,
respectively. Because both are true worldlines, the first-order
variation of each is equal to zero, ␦S = 0 and ␦共S + ␦S兲 = 0.
␦ 2S 0 − ␦ 2S 1 ⬇
␣3
2
冕P
R
dtU⵮共x0兲␾3 . 共39兲
The difference of these two relations gives ␦2S = 0 for R near
Q and hence ␦2S = 0 for R = Q. In the following we show that The fact that the right-hand side of Eq. 共39兲 is proportional
this argument is correct in the sense that ␦2S vanishes not to ␣3 means that in Eq. 共35兲 we cannot neglect terms in ␦3S
only at R = Q, but also vanishes to O共␣2兲 for R near Q, dif- that are also proportional to ␣3. 共Later terms are proportional
fering from zero by O共␣3兲 for R near Q. to ␣4 or higher.兲 From Eq. 共20兲 and the signs of ␣ in Eq.
To make Routh’s argument rigorous, consider the two al- 共31兲, we have
ternative true worldlines x0共t兲 and x1共t兲 in Fig. 12 that con-
nect the initial event P to the terminal event R, where R is
close to the kinetic focus Q of x0共t兲. When R reaches Q, the
␦ 3S 0 = −
␣3
6
冕 R

P
␾3U⵮共x0兲dt, 共40a兲

two worldlines coalesce according to the definition of the


kinetic focus Q. For definiteness, take x1共t兲 to be the top
worldline in Fig. 12, which is closely adjacent to the true
␦ 3S 1 = +
␣3
6
冕 P
R
␾3U⵮共x1兲dt, 共40b兲
worldline x0共t兲. Hence, it is a member of the set of adjacent
curves used for the variation in Sec. IV, and therefore we can and hence the second term on the right-hand side of Eq. 共35兲
employ the formalism of that section. Conversely, we can becomes

444 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 444
␦ 3S 0 − ␦ 3S 1 = −
␣3
6
冕R

P
关U⵮共x1兲 + U⵮共x0兲兴␾3 dt
P0R1⬘R2 and P1⬘R1⬘R2 have the segment R1⬘R2 in common,
and because S共P1⬘R1⬘兲 ⬍ S共P0R1⬘兲 as shown previously, we
have S共P1⬘R1⬘R2兲 ⬍ S共P0R1⬘R2兲. Hence we have found a
⬇−

3
3
冕P
R
U⵮共x0兲␾3 dt, 共41兲
variation P1⬘R1⬘R2 with a smaller action than the original
worldline P0R1⬘R2. But we know it is easy to find other varia-
tions giving a larger action 共just add wiggles somewhere兲.
where we have set U⵮共x1兲 ⬇ U⵮共x0兲, which is correct to Thus P0R1⬘R2 has a saddle point61 in the action for all R2
O共␣3兲 in Eq. 共41兲. later than the kinetic focus Q.
The substitution of Eqs. 共39兲 and 共41兲 in Eq. 共35兲 gives These demonstrations are valid for the typical case 共␦3S
⫽ 0 for the special variation ␾兲. The most common atypical
S1 − S0 ⬇
␣3
12
冕P
R
U⵮共x0兲␾3 dt for R near Q to O共␣3兲. case is the harmonic oscillator 共␦kS = 0 for all k for the spe-
cial variation ␾兲; proofs of the previous points are given
explicitly in Sec. VIII for this system. Other atypical cases,49
共42兲 for example, when ␦3S = 0 and ␦4S ⫽ 0 for the special varia-
Equation 共42兲 makes precise the earlier heuristic argument tion ␾, so that S1 − S0 ⬃ O共␣4兲 for R near Q in place of Eq.
due to Routh. We see that S1 − S0 is O共␣3兲 for R near Q and 共42兲, require an extension of the previous argument.
therefore vanishes for R → Q. If we compare Eqs. 共42兲 and We have established, for the typical case and the most
共33兲 共with ␦S0 = 0兲 and note that the coefficient ␣3 / 12 in Eq. common atypical case, the two central results of our paper:
共42兲 differs from the coefficient −␣3 / 6 in Eq. 共40a兲 for ␦3S0, 共1兲 The worldline PR has the minimum action if the terminal
we see that not only is 共␦2S0 − ␦2S1兲 ⬀ ␣3, but ␦2S0 is also event R is earlier than the kinetic focus event Q of the initial
O共␣3兲. In Ref. 60 we show that ␦2S0 is O共␣3兲 more directly event P. 共2兲 The worldline PQR has a saddle point in the
but less elegantly 共using equations of motion rather than action when the terminal event R lies beyond the kinetic
purely variational arguments兲. The result is that S1 − S0 is focus event Q of initial event P.62 These results are correct
O共␣3兲 and hence vanishes for R → Q, which yields the de- wherever on the worldline we freely choose to place the
initial event P with respect to which the later kinetic focus
sired necessary condition: Given the definition of Q, which
event Q of P is established. Note that a true maximum of the
involves two true worldlines coalescing as R → Q, we have
action for worldline PR is never found, in agreement with
␦2S = 0 for worldline PQ for the special variation leading to the result of the intuitive argument at the end of Sec. IV. In
the coalescence that defines the kinetic focus Q. the literature26 these results are often expressed as follows:
Next we extend our results to show that when R lies im-
For Lagrangians of type 共11兲 having ⳵2L / ⳵ẋ2 ⬎ 0, Jacobi’s
mediately beyond the kinetic focus Q, the action of worldline
necessary and sufficient condition for a weak47 local mini-
PQR is a saddle point. For a saddle point to occur the sign of
mum of the stationary action is that a kinetic focus does not
S1 − S0 in Eq. 共42兲 must change from positive to negative as R
occur between the end events P and R.
passes through the kinetic focus Q. To interpret the sign in
Eq. 共42兲 consider the worldline x1共t兲, the uppermost world-
line in Fig. 12, which crosses the original worldline x0共t兲 at
R1 slightly earlier than the kinetic focus Q of worldline x0共t兲. VIII. HARMONIC OSCILLATOR
We have seen that ␦2S0 ⬎ 0 for all variations for short PR
along x0共t兲 and that ␦2S0 does not vanish until R reaches Q. For the harmonic oscillator it is particularly easy to use
Fourier series to compare the action along a worldline with
Thus ␦2S0 and S1 − S0 are positive for R1 slightly earlier than
the action along every 共at least every piecewise-smooth兲
Q. Figure 12 shows and Eq. 共42兲 makes quantitative that as R
curve alternative to the worldline.65 The harmonic oscillator
takes positions from R1 to Q and then positions at Q and
Lagrangian has the form
beyond Q to R1⬘, the variational function ␣␾ vanishes, and
then changes sign to become negative. From Eq. 共42兲 we see L = K − U = 21 mẋ2 − 21 kx2 , 共43兲
that when the variation of x0共t兲 is the adjacent true worldline
so that Eq. 共19兲 becomes
x1⬘共t兲, we have S1⬘ − S0 ⬍ 0 for R1⬘ slightly later than Q, so
that S共P1⬘R1⬘兲 ⬍ S共P0R1⬘兲. However, there are other variations
in x0共t兲 that generate S1⬘ − S0 ⬎ 0, such as displacing or add-
␦ 2S =
␣2
2

P
R
˙ 2兲dt = ␣ m
共− k␾2 + m␾
2

2
冕 P
R
共␾
˙ 2 − ␻2␾2兲dt,
0

ing wiggles to a short segment 共recall the discussion at the


共44兲
end of Sec. IV兲. Thus when R1⬘ is just beyond Q we can
increase or decrease the action compared to S0, depending on where

冉冊
which variation we choose, which is the definition of a
k 1/2
2␲
saddle point: for worldline x0共t兲 or P0R1⬘, the action S0 is a ␻0 = = , 共45兲
saddle point for R1⬘ just beyond Q. m T0
We now demonstrate that worldline x0共t兲 has a saddle and T0 is the natural period. All third and higher partial de-
point in the action not only for R1⬘ just beyond Q, but also for rivatives of L with respect to x and ẋ are zero, so that there
all terminal events on the worldline beyond Q, no matter are only second-order variations in S due to ␾ 关see Eq. 共20兲兴.
how far beyond Q they lie. Imagine a point R2 further along Therefore we have
x0共t兲 from R1⬘ by an arbitrary amount, so that the true world-
S − S 0 = ␦ 2S for the harmonic oscillator. 共46兲
line in Fig. 12 is now P0R1⬘R2. Use the bottom worldline
P1⬘R1⬘ in Fig. 12 to construct the comparison curve P1⬘R1⬘R2, We set P = 共x P , 0兲 and R = 共xR , tR兲 and express the variational
which is not a true worldline due to the kink at R1⬘. Because function ␾共t兲 using a Fourier series:

445 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 445
冉 冊
⬁ lator worldlines starting from P pass through the same ki-
n␻t
␾共t兲 = 兺 an sin , 共47兲 netic focus, Fig. 3兲. For all other choices of the an 共for
n=1 2 example, a2 = A, all other an = 0兲, we have S − S0 = ␦2S ⬎ 0 in
Eq. 共50兲, a result established for the typical case in Sec. VI.
where ␻ = 2␲ / tR. The function ␾共t兲 automatically goes to
Think of each choice of all the coefficients a1 , a2 , a3 , . . . as
zero at the initial and final events P and R, respectively. The a point in function space. Then we see that at the kinetic
constants an can be chosen arbitrarily, corresponding to our focus the action is a minimum for all “directions” in function
free choice of ␾共t兲. Because of the completeness of the Fou- space except for the direction with a1 ⫽ 0, all other an = 0, for
rier series, the an can represent every possible piecewise- which S − S0 = ␦2S = 0. If we think of each an as plotted along
smooth trial curve alternative to any true worldline. We sub- a different direction in function space, we can picture this
stitute Eq. 共47兲 into Eq. 共44兲. The squares of ␾˙ and ␾ lead to exceptional stationary S case at the kinetic focus as an action
double summations: trough in function space; that is, flat in one special direction,

冕冋 冉 冊 冉 冊
⬁ ⬁ increasing in all others. 共In the typical case49 the trough is
␣ 2m tR
nn⬘␻2 n␻t n ⬘␻ t
␦ S=
2
兺 兺 a na n⬘
2 n=1 n =1 4
cos
2
cos
2
flat only to second order.兲 In contrast, for case I where tR
⬍ T0 / 2, ␦2S ⬎ 0 along every direction in function space.
⬘ 0

冉 冊 冉 冊册
Case III, tR ⬎ T0 / 2, the final time is greater than one half-
n␻t n ⬘␻ t period 共␻ / 2 = ␲ / tR ⬍ ␻0 = 2␲ / T0兲. In this case we can choose
− ␻20 sin sin dt. 共48兲
2 2 an to find adjacent curves with action either greater or less
than the action along the original worldline. For a1 = A and
As t goes from zero to tR, the arguments of the harmonic all other an = 0, ␦2S ⬍ 0, so the action for the worldline is
functions go from zero to n␲ or n⬘␲, both of which represent greater than that for the adjacent curve. In contrast, if we
an integer number of half-cycles. Because the different har- choose an = A for any term n = N for which 共N␻ / 2兲2 ⬎ ␻20 and
monics are orthogonal, terms with n⬘ ⫽ n integrate to zero for all other an = 0, then ␦2S ⬎ 0 and the action for the worldline
any number of complete half-cycles. Hence Eq. 共48兲 simpli- is smaller than for the curve. In brief, for a final time greater
fies to than half a period of the harmonic oscillator, the action for

冕 冋冉 冊 冉 冊
⬁ the worldline is neither a true maximum nor a true minimum;
␣ 2m tR
n␻ 2
n␻t
␦ S=
2
兺 a2
2 n=1 n 0 2
cos2
2
it is a saddle point. The corresponding result was shown for
the typical case in Sec. VII. Figure 11 shows schematically

− ␻20 sin2 冉 冊册
n␻t
2
dt. 共49兲
the evolution of S from case I to case II to case III as tR
increases.
All three cases apply to the second variation ␦2S for all
harmonic oscillator true worldlines x0共t兲, for example, those
The integrals from t = 0 to t = tR in Eq. 共49兲 are over n half-
cycles. For any integer number of half-cycles the average that do not start from 共x P , t P兲 = 共0 , 0兲, such as x0共t兲
cosine squared and the average sine squared are both equal to = A0 sin共␻0t + ␪0兲, and include the no-excursion or equilib-
1 / 2. Therefore we have rium worldline x0共t兲 = 0. In all cases, for actual worldlines S

冋冉 冊 册
⬁ is a minimum 共or a trough as in case II兲 or a saddle point,
␣2mtR n␻ 2
␦ 2S = 兺 a2
4 n=1 n 2
− ␻20 , 共50兲
never a true maximum, in agreement with the general theory.
For the harmonic oscillator all worldlines starting at initial
event P = 共0 , 0兲 as in Fig. 3, for example, converge next at
where ␻ and ␻0 are defined in Eq. 共45兲 and below Eq. 共47兲. event Q = 共0 , T0 / 2兲, which is therefore the kinetic focus of P.
The harmonic oscillator is atypical in that the period does not We verify this result analytically using the general method of
depend on amplitude, so that all worldlines that start at the Sec. II. For the harmonic oscillator with P = 共0 , 0兲, we ex-
same initial event recross at the same kinetic focus, as shown press the amplitude of displacement in terms of the initial
in Fig. 3. velocity v0:
We now use Eq. 共50兲 to give examples and verify, for the
atypical case of the harmonic oscillator, the results derived v0
for the typical case in earlier sections of this paper. x= sin ␻0t, 共51兲
␻0
Case I, tR ⬍ T0 / 2, the final time is less than one half-period
共␻ / 2 = ␲ / tR ⬎ ␻0 = 2␲ / T0兲. In this case ␦2S ⬎ 0 for all where ␻0 is independent of v0 for the harmonic oscillator.
choices of an and hence for every adjacent curve. For final According to Eq. 共4兲, the time tQ of the kinetic focus is found
times less than T0 / 2 共that is, for worldlines that terminate by taking the partial derivative of Eq. 共51兲 with respect to v0
before they reach the kinetic focus of the initial event兲 the and setting the result equal to zero:
action along the worldline is a minimum with respect to
every adjacent curve. ⳵x 1
= sin ␻0tQ = 0. 共52兲
Case II, tR = T0 / 2, the final time is equal to one half-period ⳵v0 ␻0
共␻ / 2 = ␲ / tR = ␻0 = 2␲ / T0兲. In this case the final event is the
kinetic focus of the initial event, where 关see Eq. 共46兲兴 S Therefore the kinetic focus of the initial event P = 共0 , 0兲 oc-
− S0 = ␦2S goes to zero for the first time for a special type of curs at the time when ␻0tQ = ␲ or tQ = T0 / 2. 共What the
variation. Choose a1 = A and an = 0 for n ⫽ 1, giving ␦2S = 0 in literature27,32 calls the “later kinetic foci” occur for ␻0t
Eq. 共50兲; examples are trial curves 1 and 2 in Fig. 3, which = 2␲ , 3␲ , . . ., but we limit the term kinetic focus to the first
here are also true worldlines. If we take the limit ␣ → 0 共or of these.兲 A similar calculation with P ⫽ 共0 , 0兲 gives the same
the zero displacement limit A → 0兲, the true worldline 1 coa- result, that is, tQ − t P = T0 / 2. The fact that tQ − t P is indepen-
lesces with true worldline 0. 共All half-period harmonic oscil- dent of the initial event P holds only for the harmonic oscil-

446 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 446
lator and does not hold for nonlinear oscillators. In Appendix x = v0t − 21 gt2 共t 艋 T0/2兲, 共54兲
B we discuss the location of the kinetic focus for the trajec-
tories of two-dimensional harmonic oscillators. where the numerical value of g for galactic oscillation de-
In summary, four characteristics of the harmonic oscillator rives from the surface mass density of the galaxy sheet. The
worldlines are exceptional; these characteristics are not true time T0 / 2 of the first half cycle is the time to return to x
for worldlines in most potential energy functions. For an = 0:
arbitrary initial event P, 共1兲 all worldlines from P pass
through the same point 共the kinetic focus兲; 共2兲 the time of the T0 2v0
= . 共55兲
kinetic focus Q of P is half a period T0 / 2 later; 共3兲 the time 2 g
interval is T0 / 2 between all successive kinetic foci; and 共4兲
After crossing into negative values of x, the worldline
when the final event R is not a kinetic focus, only one true
equation has a form similar to Eq. 共54兲:
worldline connects it to P. Underlying these four character-
istics is the exceptional property of the harmonic oscillator:
the frequency is independent of amplitude, which reflects the
linearity of the system.

x = − v0 t −
2v0
g
冊 冉
1
+ g t−
2
2v0
g
冊 2
共T0/2 艋 t 艋 T0兲.

共56兲
IX. NONLINEAR OSCILLATORS If we were considering motion in the region of positive x 共or
We could analyze the action S for the worldlines of an negative x兲 alone, there would be no kinetic focus because
arbitrary oscillator with potential U共x兲 by methods similar to U⬙ is zero in either region, leading to a positive second order
those used for the harmonic oscillator in Sec. VIII. The sec- variation in the action derived from Eq. 共53兲 as discussed in
ond order variation ␦2S from the value S0 for a worldline Sec. V. It is the infinite second derivative U⬙ at the origin of
x0共t兲 is given by Eq. 共19兲: the potential U共x兲 = C 兩 x兩 that creates the kinetic focus for the
piecewise-linear oscillator. The second derivative U⬙共x兲 of
␦ 2S =
␣2
2

P
R
˙ 共t兲2 − U⬙共x0共t兲兲␾共t兲2兴dt,
关m␾ 共53兲
this potential is
U⬙共x兲 = 2C␦共x兲, 共57兲
where ␣␾共t兲 is an arbitrary variation from x0共t兲 that vanishes where ␦共x兲 is the Dirac delta function.
at the end-events P and R. The analysis is complicated66 for Now consider the second order variation ␦2S for a world-
arbitrary U共x兲; it is more instructive to consider instead the line with 共x P , t P兲 = 共0 , 0兲 and the time tR of the terminal point
piecewise-linear oscillator with the potential U共x兲 = C 兩 x兩 and R in the range T0 / 2 艋 tR 艋 T0. As our variational function we
the quartic oscillator with a U-shaped potential U共x兲 = Cx4. choose

冉 冊
Figure 5 illustrates that the piecewise linear oscillator is rep-
resentative of the class whose period increases with increas- n␻t
␾共t兲 = a0 sin , 共58兲
ing amplitude. Figure 4 illustrates that the period of the quar- 2
tic oscillator decreases with increasing amplitude.
where ␻ = 2␲ / tR and a0 is arbitrary. The variational function
A. Piecewise-linear oscillator ␾共t兲 vanishes at the end-points P and R, as it should, and is
a slowly oscillating variation for n = 1 and a rapidly oscillat-
As an example of a piecewise-linear oscillator, consider a ing variation for n large. We substitute Eqs. 共58兲 and 共57兲 in
star that oscillates back and forth through the plane of the Eq. 共53兲. The integration over ␦共x兲 is most easily done by
galaxy and perpendicular to it.67 We approximate the galaxy changing the integration variable from t to x using dt = dx / ẋ.
as a 共freely penetrable兲 sheet of zero thickness and uniform The other integration follows the same pattern as in Sec.
mass density and express the gravitational potential energy VIII. We find

冋冉 冊 冉 冊册
of this configuration as U共x兲 = mg兩x兩, with the value of g
= C / m calculated from the mass density per unit area of the 1 n␻ 2
4 T0 2 2 n␻T0
galaxy surface. On the Earth a piecewise-linear potential of ␦2S = ␣2ma20tR − ␻ sin , 共59兲
4 2 ␲2 tR 0 4
the form68,69 U共x兲 = C 兩 x兩 with C ⬀ g models the horizontal
component of the oscillations of a particle sliding without where ␻0 = 2␲ / T0 and T0 is given by Eq. 共55兲. For suffi-
friction between two equal-angle inclined planes that meet at ciently short tR 共that is, sufficiently large ␻ = 2␲ / tR兲, the
the origin. The same form of the potential roughly models positive 共n␻ / 2兲2 term in Eq. 共59兲 will dominate for any n, so
the interaction between two quarks with x their separation. that ␦2S ⬎ 0. The action S is therefore a minimum for a
The classical, semiclassical, and quantum motion of three worldline with sufficiently short tR. For large tR 共␻ small兲,
quarks on a line interacting with mutual piecewise-linear po- the 共n␻ / 2兲2 term will again dominate for a variation with
tentials have been studied by variational methods 共see Ref. sufficiently large n. In this case we again have ␦2S ⬎ 0. But
73 and references therein兲. for the n = 1 variation the negative term in Eq. 共59兲 dominates
From these possible examples of piecewise-linear oscilla- for sufficiently small ␻ 共tR sufficiently large兲. In this case we
tors, we choose to analyze the star oscillating back and forth have ␦2S ⬍ 0. Thus for sufficiently large tR the action is a
perpendicular to the galaxy. We know that the solution to the saddle point. These results are consistent with the general
star’s oscillation on either side of the galaxy from elementary theorems derived earlier.
analysis of the vertical motion near the surface of the Earth. The dividing line between small and large tR in the pre-
With the initial event chosen as P共x P , t P兲 = 共0 , 0兲 and v0 ⬎ 0 ceding paragraph is the time of the kinetic focus. To find the
as in Fig. 5, the first half-cycle follows the parabolic world- time tQ at the kinetic focus Q of the initial event P 共see Fig.
line 5兲, we use the method developed in Sec. II. Because the

447 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 447
oscillator frequency decreases with increasing amplitude, we
know that the kinetic focus time tQ will exceed T0 / 2. We
apply condition Eq. 共4兲 to Eq. 共56兲, giving
␦ 2S =
␣2
2

0
tR
关m␾
˙ 共t兲2 − 12Cx0共t兲2␾共t兲2兴dt, 共65兲

⳵x
⳵v0

= 0 = − tQ −
2v0
g
+
2v0
g
冊+ g tQ −
2v0
g
冉 冊冉 冊

2
g
,
where we have taken t P = 0 and ␾共t兲 is an arbitrary varia-
tional function. An exact analysis for a general worldline
x0共t兲 is complicated. We therefore analyze an approximate
共60兲 worldline that brings out the salient points.
which yields Consider a periodic worldline x0共t兲 that starts from P
= 共0 , 0兲 with v0 ⬎ 0, as in Fig. 4. For a given energy or am-
tQ =
8v0 2T0 4 T0
3g
=
3
=
3 2
. 冉 冊 共61兲
plitude of motion the worldline can be approximated by12,73
x0共t兲 ⬇ A0 sin共␻0t兲. 共66兲
Thus the kinetic focus is later than the half-period by a factor Unlike the harmonic oscillator, the frequency ␻0 depends on
of 34 , as shown in Fig. 5. the amplitude A0. Action principles can be used in the direct
The spatial location xQ of the kinetic focus Q of a particu- 共Rayleigh-Ritz兲 mode12,73 to estimate ␻0, giving

冉 冊
lar worldline is found from tQ using Eq. 共56兲:
2␲
冉 冊 冉 冊
1/2
3C
2v0 1 2v0 2 ␻0 = ⬇ A0 . 共67兲
xQ = − v0 tQ − + g tQ − . 共62兲 T0 4m
g 2 g
As discussed in Ref. 12, the variational result 共67兲 is accurate
The locus of the various kinetic foci Q of the family of to better than 1%; Eqs. 共66兲 and 共67兲 can both be improved
worldlines in Fig. 5 is the caustic or envelope and can be systematically with the direct variational method if required.
found by relating v0 to tQ. From Eq. 共61兲 we have 关A direct variational method finds true trajectories from a
variational principle 共here an action principle兲 without use of
2v0 3
= tQ . 共63兲 the equations of motion.兴
g 4 We can analyze ␦2S for the quartic oscillator in the same
manner as for the piecewise-linear oscillator. We substitute
If we substitute Eq. 共63兲 into Eq. 共62兲, we obtain the relation
Eqs. 共58兲 and 共66兲 into Eq. 共65兲 and perform the integrations.
for the caustic of the family of piecewise-linear oscillator
The results are similar and we omit the details.
worldlines with P = 共0 , 0兲 and v0 ⬎ 0: As we have discussed, the cut-off time for minimum ac-
xQ = − 1 2
共64兲 tion trajectories is the kinetic focus time tQ; beyond this time
16 gtQ .
the action is a saddle point. We recall 关see Eq. 共4兲 and the
This caustic is a parabola, shown as the heavy gray line in argument there兴 that for a family of worldlines x共t , v0兲 all
Fig. 5. It divides space-time. Each final event 共xR , tR兲 above starting at event P with differing initial velocity v0, the ki-
the caustic can be reached by one or more worldlines of this netic focus of the worldline with initial velocity v0 occurs
family of worldlines; each final event on the caustic can be when ⳵x共t , v0兲 / ⳵v0 = 0. To apply this condition to the world-
reached by just one worldline of the family; and each final line in Eq. 共66兲, we first express A0 and ␻0 in Eq. 共66兲 in
event below the caustic can be reached by no worldline of terms of v0. For brevity we write Eq. 共67兲 as ␻0 = ␤A0, where
the family. For the harmonic oscillator all kinetic foci for a ␤ = 共3C / 4m兲1/2. We also have v0 = ␻0A0 from differentiation
given initial event P fall at the same point 共a focal point37兲, with respect to time of Eq. 共66兲. From these two relations we
the limiting case of a caustic. Caustics for other systems are obtain A0 = ␤−1/2v1/2
0 and ␻0 = ␤ v0 . The condition for the
1/2 1/2
discussed in Secs. IX B and X and Appendix B. kinetic focus is then
Unlike the harmonic oscillator, the time tQ − t P for the
piecewise-linear oscillator to reach a kinetic focus depends ⳵
0 sin共␤ v0 t兲兴 = 0
关␤−1/2v1/2 1/2 1/2
共68兲
on the coordinates 共x P , t P兲 of the initial event P. For ex- ⳵v0
ample, we have shown that tQ − t P = 共 34 兲T0 / 2 for x P = 0. If x P
or
⫽ 0 but still small, we find that tQ − t P ⬍ 共 34 兲T0 / 2 by approxi-
2␤ sin共␤1/2v1/2
0 t兲 + ␤ v0 cos共␤1/2v1/2
1 −1/2 −1/2 −1/2 1/2
mately 共8x P / gT20兲T0 / 2. v0 0 t兲

⫻共 21 兲␤1/2v−1/2
0 t = 0. 共69兲

B. Quartic oscillator We let ␤1/2v1/2


0 = ␻0 and obtain

tan共␻0t兲 = − ␻0t. 共70兲


Pure quartic potentials are rare in nature,74 but a mechani-
cal model is easily constructed.79 A particle is linked by har- Equation 共70兲 is satisfied for t = tQ 共and for times of later
monic springs on both sides along the y axis. The equilib- kinetic foci兲. The smallest positive root ␪Q of tan ␪ = −␪ is
rium position is y = 0 and both springs are assumed to be ␪Q ⬇ 0.646␲. The time of the kinetic focus is then given by
relaxed in this position. Oscillations along the y axis are ␻0tQ ⬇ 0.646␲ or tQ ⬇ 0.646共T0 / 2兲 for worldline 0 in Fig.
harmonic, but for small transverse oscillations in the x direc- 4; the same fraction of the half-period for the other world-
tion the potential has the form U共x兲 = Cx4 + O共x6兲. lines is shown there. Because the worldline Eq. 共66兲 is
Figure 4 shows a family of worldlines for the quartic os- approximate, the location of the kinetic focus is also ap-
cillator. The second order variation of the action for a world- proximate. Note that tQ is earlier than the half-period T0 / 2
line x0共t兲 is given from Eq. 共53兲 as for the quartic oscillator.

448 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 448
The location xQ of the kinetic focus Q共xQ , tQ兲 of a particu- Some typical worldlines with P共x P , 0兲 are shown in Fig. 6.
lar worldline is found from tQ using Eq. 共66兲: A direct worldline of arbitrary length has minimum action
共no kinetic focus兲. For indirect worldlines the kinetic foci Q
xQ = A0 sin共␻0tQ兲 = A0 sin ␪Q , 共71兲
are not the minimum x turning points, but rather the tangent
where ␪Q = 0.646␲ and A0 is the amplitude of the particular points to the straight-line caustic given by
worldline. The locus of the various kinetic foci Q共xQ , tQ兲
of the family of worldlines in Fig. 4 or caustic can be
found by relating A0 to tQ. From Eq. 共67兲 we have A0
xQ = 冉 冊
2C
m
1/2
tQ
xP
. 共77兲

= ␻0 / ␤, where ␤ = 共3C / 4m兲1/2 and ␻0tQ = ␪Q. Thus we find The derivation and discussion of ␦2S for comparison curves
B are similar to those given for the piecewise-linear oscillator
xQ = , 共72兲 in Sec. IX.
tQ The derivation of the kinetic foci 共xQ , tQ兲 and the caustic
where B = ␪Q sin ␪Q / ␤ = 1.82共4m / 3C兲1/2. This caustic in Fig. equation 共77兲 by our standard method is cumbersome for this
4 is a simple hyperbola and it too divides space-time 共see potential, so we use an alternative argument 共cf. Ref. 54兲. We
the discussion of the piecewise-linear oscillator caustic兲. can eliminate tT and xT from Eq. 共75兲 using Eqs. 共74兲 and
We have seen examples for which the kinetic focus time is 共76兲 and obtain a relation involving x, t, and the 共conserved兲
earlier than, equal to, and later than the half-period, corre- energy E. By doing some routine algebra, we can solve for
sponding, respectively, to oscillators whose frequency in- E:
creases with amplitude, is independent of amplitude, and de-
creases with amplitude. 冋
共2t2/m兲E = x2 + x2P ± 2xx P 1 −
2C t2
m x2x2P
册 1/2
. 共78兲

X. REPULSIVE INVERSE SQUARE POTENTIAL The ⫾ signs refer either to an indirect/direct pair of world-
lines or to two indirect worldlines; both situations are pos-
The previous examples were systems with exclusively sible as seen in Fig. 6. A kinetic focus arises here when two
bound motions. We now show the corresponding results for a indirect worldlines coalesce into one; the locus of kinetic
system whose unbound worldlines describe scattering from foci forms the caustic or envelope in Fig. 6. When the tra-
the potential jectories coalesce, their energies coincide. From Eq. 共78兲 we
see that the condition for coinciding energies is the vanishing
C
U共x兲 = , 共73兲 of the term in square brackets. The caustic relation is thus
x2 found to be Eq. 共77兲.
with C ⬎ 0. It might seem surprising that a worldline in a Equation 共77兲 for the caustic is seen to be plausible by the
scattering potential, where motion is unbound, can have a following argument. From Fig. 6, worldline 2, which starts
kinetic focus, because there is no kinetic focus for free par- from rest 共v0 = 0 or zero initial slope兲, has the caustic as its
ticle worldlines or for worldlines in the scattering potentials asymptote. The equation for this asymptote is easily calcu-
U共x兲 = Cx and U共x兲 = −Cx2. The difference is due to the cur- lated from Eq. 共75兲 by setting tT = 0 and E = C / x2P and taking
t large.
vatures of the potentials: the inverse square potential 共73兲 has
As we have seen, the indirect worldlines each have a ki-
U⬙共x兲 ⬎ 0, whereas the other potentials have U⬙共x兲 艋 0. As
netic focus. In contrast to the oscillator systems studied ear-
discussed qualitatively in Sec. III, potentials with U⬙ ⬎ 0 are lier, subsequent kinetic foci do not exist for this system.
stabilizing/focusing, which can lead to a kinetic focus.
For a given initial position x P and final position xR in the
potential 共73兲, a worldline may be direct 共direct motion from
x P to xR兲 or indirect 共backward motion from x P to a turning XI. GENERALIZATIONS
point xT followed by forward motion from xT to xR兲. For
indirect worldlines the turning point 共xT , tT兲 occurs where the Extensions of the results of this paper to two- and three-
kinetic energy is equal to zero, so that the total energy E is dimensional motion and multiparticle systems are straight-
equal to the potential energy 共73兲, yielding forward, primarily because the action and energy are scalars;
adding dimensions or particles sums the corresponding scalar
C
xT2 = . 共74兲 quantities. We let xi denote the coordinates, for example,
E 共x1 , x2兲 = 共x , y兲 for motion of a single particle in two dimen-
The worldlines x共t兲 for the potential 共73兲 are calculated by sions; 共x1 , x2 , x3兲 = 共x , y , z兲 for motion of a single particle in
integrating the energy conservation relation. We assume three dimensions; and 共x1 , . . . , x6兲 for two particles in three
共x P , t P兲 = 共x P , 0兲 and find dimensions, where 共x1 , x2 , x3兲 = 共x , y , z兲 for particle one and
共x4 , x5 , x6兲 = 共x , y , z兲 for particle two. Equation 共5兲 generalizes
2E to
x2 = xT2 + 共t ± tT兲2 , 共75兲
m
xi = x共0兲
i + ␣␾i , 共79兲
where ⫾ apply to the direct/indirect worldline, respectively.
For an indirect worldline with the initial event P共x P , 0兲, Eq. where x共0兲
i and xi are the coordinates of a point on the actual
共75兲 can be solved for the turn-around time tT: worldline and varied curve, respectively. This formalism

冉 冊 1/2 leads to obvious generalizations of the equations of Sec. IV.


m In particular, for one or more particles of mass m, the La-
tT = 共x2P − xT2 兲1/2 . 共76兲
2E grangian is

449 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 449
1 potentials,92–94 but as far as we know not for chaotic world-
L = 兺 mẋ2i − U, 共80兲 lines in particular.
i 2
Chaotic behavior can arise in higher dimensions even
without explicitly time-dependent potentials. As an example,
where U can be a function of all the xi and time. Equation the worldlines for the Henon-Heiles oscillator
共19兲 generalizes to

␦ S=
2 ␣2
2
冕冋
R

P
− 兺 ␾ i␾ j
ij
⳵ 2U
⳵ x i⳵ x j
+ 兺 m␾
i

˙ 2 dt.
i 共81兲
U共x,y兲 = 21 k共x2 + y 2兲 + ␣x2y − 31 ␤y 3 共83兲

are chaotic for certain values of the initial conditions and


The kinetic energy in Eq. 共80兲 is quadratic in the velocities parameters.95 Some studies of ␦2S and kinetic foci have been
and thus positive, which leads to S having a minimum or a done on periodic worldlines for this system,93 but we are not
saddle point 共never a maximum兲 for true worldlines 共see the aware of any studies for chaotic worldlines.
argument at the end of Sec. IV兲. It would be interesting to It would be interesting and challenging to study ␦2S for
investigate the possible extension of this result and the cor- chaotic worldlines.96 We hypothesize that kinetic foci will
responding result for the Maupertuis action W 共see Appendix not exist if the worldline is sufficiently chaotic. In such cases
A兲 to more general Lagrangians, including relativistic worldlines with incremental difference in velocity at initial
Lagrangians and Lagrangians containing terms linear in the event P may recross pseudorandomly in time, but the severe
velocities 共for example, magnetic field terms, gyroscopic instability may prevent the two worldlines from smoothly
terms兲. coalescing, as required for the existence of a kinetic focus Q.
Appendix B describes the 2D motion of a particle in two Worldlines PR lacking kinetic foci have ␦2S ⬎ 0 for arbitrary
types of gravitational potentials and in harmonic oscillator final events R, so that the action is expected to remain a
potentials. The criteria for the minimum action and location minimum in such cases, even for long worldlines in poten-
of kinetic foci80 are similar, but, unlike one dimension, two tials having U⬙ ⬎ 0, such as Eq. 共82兲.
trajectories connecting P to R can both have minimum action
for the attractive 1 / r potential. The analysis generalizes eas-
ily to other dimensions and multi-particle systems, but the
calculations are more complicated for complex worldlines, XII. SUMMARY
for example, when the motion is chaotic. For many-particle
systems it is unlikely that we will want to specify in advance We have investigated the nature of the stationary value of
the complete final as well as initial configuration, because a the Hamilton action S for the worldlines of a single particle
major goal of mechanics is to find the final configuration.81 in one dimension with potential energy function U共x兲. We
In such cases these powerful deterministic tools may be less showed that when no kinetic focus exists, the action is a
useful than modern statistical mechanical methods, although minimum for worldlines of arbitrary length. When a kinetic
Helmholtz, Boltzmann, Planck, and others have attempted to focus exists, and when a worldline terminates before reach-
base the second law of thermodynamics on action principles ing its kinetic focus, the action is still a minimum. In con-
for the molecular motions.73 trast, when a worldline terminates beyond its kinetic focus,
We have been careful to use partial derivatives of the po- its action is a saddle point. The value of the action S is never
tential function with respect to position, because the potential a true maximum for a true worldline. These results were
energy U共x , t兲 can be an explicit function of time. The results illustrated for the harmonic oscillator, two anharmonic oscil-
of this paper can be applied in principle to motion in time- lators, and a scattering system. Extensions to time-dependent
dependent potentials, in which the energy of the particle may 1D potentials and to multidimensional potentials were dis-
not be a constant of the motion. New qualitative features cussed briefly. The appendices supply parallel results for spa-
may arise if U is explicitly time dependent; for example, we tial orbits described by Maupertuis’ action W and give ex-
expect that ␦2S can remain positive for some long worldlines amples for 2D motion for both S and W. Corresponding
in potentials with U⬙ ⬎ 0.84 As an example, consider the results for some newer action principles have not yet been
quartic oscillator with the time-dependent external forcing derived, and open questions about these newer action prin-
F共t兲. The potential becomes ciples are sketched in Appendix C.

U共x,t兲 = Cx4 − xF共t兲 共82兲


ACKNOWLEDGMENTS
and has U⬙ ⬎ 0 for all x except x = 0. A common choice is
F共t兲 = F0 cos ␻t, but other choices are also of interest, for CGG thanks NSERC for financial support. EFT acknowl-
example, quasiperiodic forcing F共t兲 = F1 cos ␻1t + F2 cos ␻2t, edges James Reid-Cunningham for providing opportunities
with ␻2 / ␻1 irrational. Alternatively, we can introduce para- for extended thought, and Alar Toomre for tutoring in math-
metric forcing by modulating C. The unforced oscillator has ematics. Both authors thank Bernie Nickel for many sugges-
only equilibrium and periodic worldlines, which are stable. tions, including pointing out the insufficiency of the
Depending on the initial conditions and potential parameters, Culverwell-Whittaker argument as shown in Sec. VI and
the forced oscillator can also have unstable periodic providing the original analysis elaborated in Sec. VII. We
worldlines,30 quasiperiodic worldlines,87,88 and chaotic 共ape- also thank Adam Pound, Slavomir Tuleja, Don Lemons,
riodic, bounded, exponentially unstable兲 worldlines.90 Simi- Donald Sprung, and John Taylor for a careful reading of the
larly the potential U共x兲 = C 兩 x兩 can be made time dependent.91 manuscript and many useful suggestions. Brenda Law, Igor
Second variations and kinetic foci have been studied for Tolokh, and Will M. Farr are thanked for their assistance in
some worldlines of various oscillators with time dependent the production of this paper.

450 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 450
APPENDIX A: THE MAUPERTUIS ACTION two arbitrary intermediate positions between x P and xR. Con-
PRINCIPLE sider a second trial path x P → A → B → A → B → xR that has
an extra “loop” inserted, with the momentum p reversed at
There are two major versions of the action and two corre- every point along B → A compared to A → B. This compari-
sponding action principles. The Hamilton or time-dependent son path satisfies the constraint of having the same energy as
action S and the corresponding Hamilton action principle the actual path, but has a larger action because pdx is always
were introduced in Sec. I. The Maupertuis or time- positive. Thus W for the actual path cannot be a true maxi-
independent action W is defined along an arbitrary trial tra- mum.
jectory connecting P共x P , t P兲 to R共xR , tR兲 by

W= 冕xP
xR
p dx = 冕
tP
tR
mẋ
dx
dt
dt = 冕
tP
tR
2K dt, 共A1兲
APPENDIX B: TWO-DIMENSIONAL
TRAJECTORIES
1. Gravitational fields
where the first 共time-independent兲 form is the general defi-
nition with p = ⳵L / ⳵ẋ the canonical momentum, and the last We have shown that the Hamilton action S is a minimum
共time-dependent兲 form is valid generally for normal for all radial/vertical trajectories in the 1 / r and linear gravi-
systems73 in which the kinetic energy K is quadratic in the tational potentials discussed in Sec. V. This minimum action
velocity components. For normal systems W is positive for property may not hold for 2D trajectories, as we shall dis-
all trajectories in all potentials 共unlike S兲. The Maupertuis cuss. A possible nonminimum in the action is more evident
action principle states that in conservative systems W is sta- for the Maupertuis action W, Eq. 共A1兲, for which the true
tionary 共␦W = 0兲 for an actual trajectory when comparing trial trajectories are defined by giving the two end positions
共x P , y P兲 and 共xR , y R兲 and the energy E; we therefore discuss
trajectories all with the same fixed energy E and the same
fixed start and end positions x P and xR. Note that in Mauper- 2D orbits for W first. We choose the x and y axes in the plane
tuis’ principle the energy E is fixed and the duration T = 共tR of the orbit.
For U共x , y兲 = mgy, with y the vertical direction and x the
− t P兲 is not, the opposite conditions of those in Hamilton’s
horizontal, it is well known that two actual spatial orbits
principle. The constraint of fixed end positions x P and xR is
共parabolas兲 with the same energy E can connect two given
common to both principles. Hamilton’s principle is valid for
positions, the origin 共x P , y P兲 = 共0 , 0兲 say, and the final posi-
both conservative systems and nonconservative systems with
U = U共x , t兲. The conventional Maupertuis principle is valid tion 共xR , y R兲, provided that 共xR , y R兲 lies within the “parabola
only for conservative systems; the extension to nonconserva- of safety”—the envelope43,101 of the parabolic orbits of en-
tive systems is discussed in Ref. 73. Maupertuis’ principle ergy E originating at 共0,0兲 共see Fig. 7兲. If 共xR , y R兲 lies on the
can be used in its time-independent form to find spatial orbits parabola of safety, which is the locus of the spatial kinetic
关for example, 共x P , y P兲 → 共xR , y R兲兴 and in its time-dependent foci 共xQ , y Q兲 or caustic, there is one actual orbit between the
form to find space-time trajectories or worldlines, for ex- fixed end positions. If 共xR , y R兲 lies outside the parabola of
ample, 共x P , y P , t P兲 → 共xR , y R , tR兲. safety, no orbit of energy E can connect it to the origin.
The Hamilton and Maupertuis action principles can be These conclusions are evident in Fig. 7. As we have seen in
stated73 succinctly in terms of constrained variations as Sec. II, the intersection of two paths implies the existence of
共␦S兲T = 0 and 共␦W兲E = 0, respectively, where the constraints of a 共different兲 kinetic focus for each of the paths. W is a mini-
fixed T and fixed E are denoted explicitly as subscripts, and mum for the path for which 共xR , y R兲 precedes its kinetic fo-
the constraint of fixed end positions x P and xR is implicit. cus, and is a saddle point for the path for which 共xR , y R兲 lies
Along an arbitrary trial trajectory P → R, S and W are beyond its kinetic focus.
related73 by a Legendre transformation, that is, Similarly, as first shown by Jacobi,22 typically102 two
given positions 共x P , y P兲 and 共xR , y R兲 in the gravitational po-
S = W − ĒT, 共A2兲 tential 共1 / r兲 can be connected by two actual orbits 共ellipses兲
共and therefore four paths of less than one revolution兲 of the
where Ē = 兰ttRH dt / T is the mean energy along the arbitrary same energy E. Again, the fact that more than one true path
P
trajectory, T = 共tR − t P兲 is the duration, and H is the Hamil- can connect the two end positions leads to a nonminimum in
tonian. Equation 共A2兲 follows by integrating over time from the action W for actual paths connecting 共x P , y P兲 to 共xR , y R兲
t P to tR along the arbitrary trajectory the corresponding Leg- when 共xR , y R兲 lies beyond the kinetic focus. An example is
endre transform L = pẋ − H and deserves to be better known. shown in Fig. 13. The intersection points of the orbits show
Along an actual trajectory of a conservative system, Eq. 共A2兲 two ellipses connecting point P to other points. The outer
reduces to the well-known relation97 S = W − ET, where E is curve is the envelope/caustic, which is also elliptical with
the constant energy of the actual trajectory. Using Eq. 共A2兲 foci at P and the force center. The spatial kinetic foci for
enables us to relate the Hamilton and Maupertuis action W lie on this outer ellipse. The second spatial kinetic
principles.73 focus occurs at P itself, following one revolution. There is no
Parallel yet distinct discussions have been developed for envelope for the hyperbolic scattering orbits for the attractive
the second variations of S and W because the kinetic foci, 1 / r potential.104 Methods of determining spatial caustics are
which play such an important role in determining the second similar to those for determining space-time caustics and are
variation, can differ for the two actions.30,98,99 discussed in Ref. 105.
An intuitive argument why W can never be a true maxi- To discuss 2D space-time worldlines for the Hamilton ac-
mum for actual paths was given in Ref. 100 for normal sys- tion S, which depend on the two end positions and the time
tems. Consider an actual path x P → A → B → xR that makes interval that now specify a worldline, note that for the 1 / r
the first form of W in Eq. 共A1兲 stationary. Here A and B are potential, typically two actual worldlines can connect two

451 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 451
Fig. 15. An elliptical orbit 共tilted ellipse兲 in a 2D isotropic harmonic oscil-
lator potential U共r兲 = 共 2 兲kr2 with force center at O. A family of trajectories is
1

launched from P with equal initial speeds 兩v0兩 and various directions ␪0. One
member of the family is shown. The envelope of the family is the outer
ellipse 共heavy line兲, with foci at P and Q 共coordinates xQ = −x P, y Q = y P = 0兲.
Points on the envelope are the kinetic foci for the spatial orbits. Point Q,
occurring at time tQ = T0 / 2, where T0 is the period, is the kinetic focus for
the space-time trajectories 共worldlines兲. 共Figure adapted from French, Ref.
101.兲

action. For this particular potential, only one space-time path


exists for one dimension, which therefore has minimum ac-
Fig. 13. A family of elliptical trajectories starting at P with the same speed tion S 共see Sec. V兲.
兩v0兩 共the directions ␪0 of v0 differ兲, and hence the same energy E, the same For the potential U共x , y兲 = mgy, only one 2D 共parabolic-
major axis 2a, and the same period T0, in a 1 / r gravitational potential. The shaped兲 actual worldline can connect two given positions
value of v0 exceeds that necessary to generate a circular orbit. The center of 共x P , y P兲 and 共xR , y R兲 in the given time 共tR − t P兲. Kinetic foci
force is the Earth 共heavy circle兲. The dashed circle gives the locus of the
second focus of the ellipses 共a circle centered at P兲. The outer ellipse, with
for the space-time trajectories therefore cannot arise, and
foci at P and the Earth, is the envelope of the family of ellipses and the hence S is always a minimum for actual worldlines for this
locus of the spatial kinetic foci relevant for action W. The space-time kinetic potential.112 The fact that for this potential only one true
foci of the family, relevant for action S, all occur at time tQ = T0. 共Figure worldline can connect two points in a given time is in con-
adapted from Butikov, Ref. 101.兲 trast with the Maupertuis action W for which we have seen
that some pairs of positions 共x P , y P兲 and 共xR , y R兲 can be con-
nected by more than one path of a given energy E so that
given positions 共x P , y P兲 and 共xR , y R兲 in the given time interval kinetic foci can exist for the spatial orbits. This contrast be-
共tR − t P兲.107 The fact that two actual worldlines exist here is tween S and W also holds for the vertical 1D paths in the
illustrated in Fig. 14, which shows two elliptical trajectories potential U = mgy. For S only one actual 1D worldline can
that connect the initial and final points in the same time. connect a given y P to a given y R in the given time 共tR − t P兲,
Choose t P = 0 for simplicity. The kinetic focus time tQ for S is which leads to the conclusion that all 1D actual worldlines
the period T0 as is clear intuitively from Fig. 13, which minimize S 共see Sec. V兲. For W typically two actual 1D
shows a family of trajectories leaving point P and converg- paths of given energy E can connect y P to y R, which leads to
ing back on P in the same time T0. A rigorous proof that the conclusion that not all 1D actual paths minimize W
tQ = T0 can also be given.110 The space-time kinetic focus 共some are saddle points兲. For the potential U共x , y兲 = mgy
here is of the focal point type, as in Fig. 1 for the sphere there is always one actual worldline that can connect two
geodesics and in Fig. 3 for the harmonic oscillator world- given spatial points in a given time, for both 2D and 1D
lines. As seen in Fig. 14, if two trajectories connect P to R in worldlines 共S is always a minimum for these worldlines兲. In
time 共tR − t P兲 ⬍ T0, both have the minimum action; the fact contrast, there may be no actual path that connects two given
that both trajectories have minimum action is in contrast to spatial points for a given energy, for both one and two di-
one dimension, where we have seen in general that when two mensions. Figure 7 shows examples 共final points outside the
space-time paths exist, one of them has a saddle point in the caustic兲 of this nonexistence of actual paths.

2. Harmonic oscillators
The potential for a 2D isotropic harmonic oscillator is
U共x , y兲 = 21 k共x2 + y 2兲 ⬅ 21 kr2. The spatial orbits are ellipses
with the force center 共r = 0兲 at the center of the ellipse. A
family of ellipses launched from P in Fig. 15, with equal
values of 兩v0兩 共and hence equal energies E兲 and various di-
rections of v0, has an envelope/caustic that is also
elliptical.101 The envelope is the outer ellipse 共heavy line兲 in
Fig. 15. The spatial kinetic foci for the action W lie on the
caustic. Reference 105 gives methods for deriving the spatial
caustic.
Fig. 14. Two different elliptical trajectories typically can connect P To locate the space-time kinetic focus, which is relevant
= 共r P , t P兲 to R = 共rR , tR兲 in the same time 共tR − t P兲 for the attractive 1 / r poten- for the action S, we apply the relation given in Ref. 80 for 2D
tial. 共Adapted from Bate et al., Ref. 108.兲 trajectories x共t , v0兲. The matrix ⳵xi / ⳵v0j is diagonal, so that

452 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 452
APPENDIX C: OPEN QUESTIONS FOR SOME
NEWER ACTION PRINCIPLES

Culverwell and Whittaker 共see Sec. VI兲 framed their


analysis in terms of the Maupertuis action W. For sufficiently
short trajectories 共the final position occurs before the kinetic
focus兲 W is always a minimum, and for longer trajectories W
is a saddle point. W is never a true maximum. In this paper
we amended the Culverwell-Whittaker analysis and adapted
it to the Hamilton action S. For times less than the kinetic
focus time tQ, the action S is always a minimum. For longer
times S is a saddle point. S is never a true maximum. We
refer to these results for W and S as “no-max” theorems.
It may be possible to extend the theorems to several newer
action principles.73,115 To state the newer principles, we first
Fig. 16. A periodic orbit of a 2D anisotropic harmonic oscillator with com- recall the notation for the Hamilton principle and the Mau-
mensurate frequencies 共here ␻1 / ␻2 = 2兲 共Ref. 113兲. The space-time kinetic pertuis principle given in Appendix A:
focus occurs at time tQ = T1 / 2, where T1 is the period for x-motion, for any
initial event 共x P , y P , t P = 0兲 共see text兲, and the spatial kinetic focus for initial 共␦S兲T = 0 共Hamilton principle兲, 共C1兲
position 共x P , y P兲 = 共0 , 0兲 is located on a parabolic spatial caustic 共see Ref.
105兲.
共␦W兲E = 0 共Maupertuis principle兲, 共C2兲
where T = tR − t P 共the duration兲 and E 共the energy兲 denote the
constraints. The additional constraints of fixed end positions
the determinant condition reduces to 共⳵x / ⳵v0x兲共⳵y / ⳵v0y兲 = 0, x P and xR are implicit in Eqs. 共C1兲 and 共C2兲 and are under-
and we obtain separate 1D conditions for the x and y mo- stood to hold here and in all the action principles discussed
tions, that is, ⳵x / ⳵v0x = 0 or ⳵y / ⳵v0y = 0. Choose t P = 0 for in the following.
simplicity. We showed in Sec. VIII that the kinetic focus In recent years the Maupertuis principle 共C2兲 has been
time tQ for the 1D harmonic oscillator is T0 / 2, where T0 extended to a generalized Maupertuis principle,73,115
= 2␲ / ␻0 and ␻0 = 共k / m兲1/2. Thus we have tQ = T0 / 2 for the 共␦W兲¯E = 0 共generalized Maupertuis principle兲, 共C3兲
isotropic 2D harmonic oscillator.
The space-time kinetic focus for the trajectories of the 2D where Ē = 兰T0 H dt / T is the mean energy along the arbitrary
anisotropic harmonic oscillator with the potential U共x , y兲 trial trajectory, with H the Hamiltonian, and where for sim-
= 21 k1x2 + 21 k2y 2 and k1 ⫽ k2 can be derived similarly. The or- plicity we choose t P = 0 and tR = T. The constraint of fixed E
bits are Lissajous figures, closed 共periodic兲 for rational ratios in Eq. 共C2兲 has been weakened to one of fixed mean energy
␻1 / ␻2 as in Fig. 16, and open 共quasiperiodic兲 for ␻1 / ␻2 Ē in Eq. 共C3兲. Conservation of energy for actual trajectories
irrational as in Fig. 17, where ␻i = 共ki / m兲1/2. The problem is is now a consequence of the principle 共C3兲, rather than an
again separable into x and y motions, and the method of Ref. assumption as in the original principle 共C2兲.
80 yields for the kinetic focus time tQ the value T0 / 2, where Both the generalized Maupertuis principle 共C3兲 and
T0 is the smaller of T1 and T2 and Ti = 2␲ / ␻i. The determi- Hamilton principle 共C1兲 have associated reciprocal
nation of the spatial kinetic foci, relevant for W, is more principles:73,115
complicated.105
共␦Ē兲W = 0 共reciprocal Maupertuis principle兲, 共C4兲

共␦T兲S = 0 共reciprocal Hamilton principle兲. 共C5兲


The newer principles in Eqs. 共C3兲–共C5兲 have several advan-
tageous features, computational and conceptual, as discussed
in Refs. 12, 73, and 115. Additionally, the reciprocal Mau-
pertuis principle 共C4兲 is the direct classical analogue115,116
共the classical ប → 0 limit兲 of the well known Schrödinger
quantum variational principle involving the mean energy.117
It would be of interest to prove the existence or nonexist-
ence of a no-max or no-min theorem for these newer action
principles. A Routh-type argument 共see Appendix A兲 sug-
gests that the generalized Maupertuis principle 共C3兲 obeys a
no-max theorem, but the examples worked out to date12,73,115
provide no compelling evidence one way or the other for the
other principles.
Fig. 17. A quasiperiodic orbit of a 2D anisotropic harmonic oscillator with Other newer action principles are discussed in Refs. 73
incommensurate frequencies. The outer ellipse is the equipotential contour and 115. We can completely relax the constraints of fixed T
U共x , y兲 = E. The rectangle delimits the region of x-y space actually reached
by the particular orbit 共Ref. 114兲. The space-time kinetic focus occurs at in Eq. 共C1兲 and fixed Ē in Eq. 共C3兲 with the help of Lagrange
time tQ = T2 / 2, where T2 is the period for y-motion, for any initial event multipliers and obtain an unconstrained Hamilton principle,
共x P , y P , t P = 0兲. ␦S = −E␦T, and an unconstrained Maupertuis principle, ␦W

453 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 453
and C. Pellegrini, Ref. 14.
= T␦Ē, where 共for conservative systems兲 the Lagrange mul- 17
F. L. Pedrotti, L. S. Pedrotti, and L. M. Pedrotti, Introduction to Optics
tipliers E and T are the energy and duration of the actual 共Pearson Prentice Hall, Upper Saddle River, NJ, 2007兲, p. 22; J. R. Tay-
trajectory, respectively. The unconstrained Hamilton prin- lor, Classical Mechanics 共University Science Books, Sausalito, 2005兲,
ciple and unconstrained Maupertuis principle can also be Problem 6.5; P. J. Nahin, When Least is Best 共Princeton U. P., Princeton,
written in the more suggestive forms ␦共S + ␭T兲 = 0 and ␦共W NJ, 2004兲, p. 133; D. S. Lemons, Perfect Form 共Princeton U. P., Prince-
ton, NJ, 1997兲, p. 8; D. Park, Ref. 14, p. 13; F. A. Jenkins and H. E.
+ ␭Ē兲 = 0, respectively, where ␭ = E or ␭ = −T is the corre- White, Fundamentals of Optics, 4th ed. 共McGraw Hill, New York, 1976兲,
sponding constant Lagrange multiplier. These unconstrained p. 15; R. Guenther, Modern Optics 共Wiley, New York, 1990兲, p. 135; R.
principles still have the constraint of fixed end positions x P W. Ditchburn, Light, 3rd ed. 共Academic, London, 1976兲, p. 209; R. S.
and xR; these constraints can be relaxed by introducing addi- Longhurst, Geometrical and Physical Optics, 2nd ed. 共Longmans, Lon-
don, 1967兲, p. 7; J. L. Synge, Geometrical Optics 共Cambridge U. P.,
tional Lagrange multipliers.73 It would also be of interest to London, 1937兲, p. 3; G. P. Sastry, “Problem on Fermat’s principle,” Am.
prove the existence or nonexistence of no-max or no-min J. Phys. 49, 345 共1981兲; M. V. Berry, review of The Optics of Rays,
theorems for these various principles. Wavefronts and Caustics by O. N. Stavroudis 共Academic, New York,
1972兲, Sci. Prog. 61, 595–597 共1974兲; V. Lakshminarayanan, A. K.
a兲
Ghatak, and K. Thyagarajan, Lagrangian Optics 共Kluwer, Boston, 2002兲,
Electronic mail: cgg@physics.uoguelph.ca p. 16; V. Perlick, Ray Optics, Fermat’s Principle, and Applications to
b兲
Electronic mail: eftaylor@mit.edu General Relativity 共Springer, Berlin, 2000兲, pp. 149 and 152.
1
R. P. Feynman, R. B. Leighton, and M. Sands, The Feynman Lectures on 18
The seminal work on second variations of general functionals by Leg-
Physics 共Addison-Wesley, Reading, MA, 1963兲, Vol. II, Chap. 19. endre, Jacobi, and Weierstrass and many others is described in the his-
2
A. Bork and A. Zellweger, “Least action via computer,” Am. J. Phys. 37, torical accounts of Refs. 19 and 20. Mayer’s work 共Ref. 19兲 was devoted
386–390 共1969兲. specifically to the second variation of the Hamilton action. In our paper,
3
J. P. Provost, “Least action principle on an air table,” Am. J. Phys. 43, we adapt Culverwell’s work 共Ref. 21兲 for the Maupertuis action W to the
774–781 共1975兲. Hamilton action. Culverwell’s work was preceded by that of Jacobi 共Ref.
4
E. F. Taylor, “A call to action,” Am. J. Phys. 71, 423–425 共2003兲. 22兲 and Kelvin and Tait 共Ref. 23兲.
5
J. Hanc, S. Tuleja, and M. Hancova, “Simple derivation of Newtonian 19
H. H. Goldstine, A History of the Calculus of Variations From the 17th
mechanics from the principle of least action,” Am. J. Phys. 71, 386–391 Through the 19th Century 共Springer, New York, 1980兲.
共2003兲. 20
I. Todhunter, A History of the Progress of the Calculus of Variations
6
J. Hanc, S. Tuleja, and M. Hancova, “Symmetries and conservation laws: During the Nineteenth Century 共Cambridge U. P., Cambridge, 1861兲 and
Consequences of Noether’s theorem,” Am. J. Phys. 72, 428–435 共2004兲. 共Dover, New York, 2005兲.
7 21
J. Hanc, E. F. Taylor, and S. Tuleja, “Deriving Lagrange’s equations E. P. Culverwell, “The discrimination of maxima and minima values of
using elementary calculus,” Am. J. Phys. 72, 510–513 共2004兲. single integrals with any number of dependent variables and any continu-
8
J. Hanc and E. F. Taylor, “From conservation of energy to the principle of ous restrictions of the variables, the limiting values of the variables being
least action: A story line,” Am. J. Phys. 72, 514–521 共2004兲. supposed given,” Proc. London Math. Soc. 23, 241–265 共1892兲.
9 22
T. A. Moore, “Getting the most action out of the least action: A proposal,” C. G. J. Jacobi, “Zür Theorie der Variationensrechnung und der Differ-
Am. J. Phys. 72, 522–527 共2004兲. ential Gleichungen.” J. f.Math. XVII, 68–82 共1837兲. An English trans-
10
J. Ogborn and E. F. Taylor, “Quantum physics explains Newton’s laws of lation is given in Ref. 20, p. 243, and a commentary is given in Ref. 19,
motion,” Phys. Educ. 40, 26–34 共2005兲. p. 156.
11 23
E. F. Taylor, J. Hanc, and S. Tuleja, “Variational mechanics in one and W. Thomson 共Lord Kelvin兲 and P. G. Tait, Treatise on Natural Philoso-
two dimensions,” Am. J. Phys. 73, 603–610 共2005兲. phy 共Cambridge U. P., Cambridge, 1879, 1912兲, Part I; reprinted as Prin-
12
C. G. Gray, G. Karl, and V. A. Novikov, “Direct use of variational prin- ciples of Mechanics and Dynamics 共Dover, New York, 1962兲, Part I, p.
ciples as an approximation technique in classical mechanics,” Am. J. 422.
Phys. 64, 1177–1184 共1996兲 and references therein. 24
Gelfand and Fomin 共Ref. 25兲 and other more recent books on calculus of
13
J. L. Lagrange, Analytical Mechanics (Mécanique Analytique) 共Gauthier- variations are rigorous but rather sophisticated. A previous study 共Ref.
Villars, Paris, 1888–89兲, 2nd ed. 共1811兲 共Kluwer, Dordrecht, 1997兲, pp. 26兲 of the nature of the stationarity of worldline action was based on the
183 and 219. The same erroneous statement occurs in work published in Jacobi-Morse eigenfunction method 共Ref. 27兲, rather than on the more
1760–61, ibid, p. xxxiii. geometrical Jacobi-Culverwell-Whittaker approach.
14 25
We easily found about two dozen texts using the erroneous term “maxi- I. M. Gelfand and S. V. Fomin, Calculus of Variations, translated by R. A.
mum.” See, for example, E. Mach, The Science of Mechanics, 9th ed. Silverman 共Prentice Hall, Englewood Cliffs, NJ, 1963兲, Russian edition
共Open Court, La Salle, IL, 1960兲, p. 463; A. Sommerfeld, Mechanics 1961, reprinted 共Dover, New York, 2000兲.
共Academic, New York, 1952兲, p. 208; P. M. Morse and H. Feshbach, 26
M. S. Hussein, J. G. Pereira, V. Stojanoff, and H. Takai, “The sufficient
Methods of Theoretical Physics 共McGraw Hill, New York, 1953兲, Part 1, condition for an extremum in the classical action integral as an eigen-
p. 281; D. Park, Classical Dynamics and its Quantum Analogues, 2nd ed. value problem,” Am. J. Phys. 48, 767–770 共1980兲. Hussein et al. make
共Springer, Berlin, 1990兲, p. 57; L. H. Hand and J. D. Finch, Analytical the common error of assuming S can be a true maximum. See also J. G.
Mechanics 共Cambridge U. P., Cambridge, 1998兲, p. 53; G. R. Fowles and Papastavridis, “An eigenvalue criterion for the study of the Hamiltonian
G. L. Cassiday, Analytical Mechanics, 6th ed. 共Saunders, Forth Worth, action’s extremality,” Mech. Res. Commun. 10, 171–179 共1983兲.
27
1999兲, p. 393; R. K. Cooper and C. Pellegrini, Modern Analytical Me- M. Morse, The Calculus of Variations in the Large 共American Math-
chanics 共Kluwer, New York, 1999兲, p. 34. A similar error occurs in A. J. ematical Society, Providence, RI, 1934兲.
Hanson, Visualizing Quaternions 共Elsevier, Amsterdam, 2006兲, p. 368, 28
As we shall see, the nature of the stationary value of Hamilton’s action S
which states that geodesics on a sphere can have maximum length. 共and also Maupertuis’ action W兲 depends on the sign of second variations
15
E. F. Taylor and J. A. Wheeler, Exploring Black Holes: Introduction to ␦2S and ␦2W 共defined formally in Sec. IV兲, which in turn depends on the
General Relativity 共Addison-Wesley Longman, San Francisco, 2000兲, pp. existence or absence of kinetic foci 共see Secs. II and VII兲. The same
1–7 and 3–4. quantities 共signs of the second variations and kinetic foci兲 are also im-
16
The number of authors of books and papers using “extremum” and “ex- portant in classical mechanics for the question of dynamical stability of
tremal” is endless. Some examples include R. Baierlein, Newtonian Dy- trajectories 共Refs. 23 and 29–31兲, and in semiclassical mechanics where
namics 共McGraw-Hill, New York, 1983兲, p. 125; L. N. Hand and J. D. they determine the phase loss term in the total phase of the semiclassical
Finch, Ref. 14, p. 51; J. D. Logan, Invariant Variational Principles 共Aca- propagator due to a particular classical path 共Refs. 32 and 33兲. The phase
demic, New York, 1977兲 mentions both “extremal” and “stationary,” p. 8; loss depends on the Morse 共or Morse-Maslov兲 index, which equals the
L. D. Landau and E. M. Lifschitz, Mechanics, 3rd ed. 共Butterworth Hei- number of kinetic foci between the end-points of the trajectory 共see Ref.
neman, Oxford, 2003兲, pp. 2 and 3; see also their Classical Theory of 34兲. Further, in devising computational algorithms to find the stationary
Fields, 4th ed. 共Butterworth Heineman, Oxford, 1999兲, p. 25; S. T. Thorn- points of the action 共either S or W兲, it is useful to know whether we are
ton and J. B. Marion, Classical Dynamics of Particles and Systems, 5th seeking a minimum or a saddle point, because different algorithms 共Ref.
ed. 共Thomson, Brooks/Cole, Belmont, CA, 2004兲, p. 231; R. K. Cooper 35兲 are often used for the two cases. As we discuss in this paper, it is the

454 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 454
sign of ␦2S 共or ␦2W兲 that determines which case we are considering. for ␣ → 0, is termed a weak variation. See, for example, C. Fox, An
Practical applications of the mechanical focal points are mentioned at the Introduction to the Calculus of Variations 共Oxford U. P., Oxford, 1950兲,
end of Ref. 37. reprinted 共Dover, New York, 1987兲, p. 3.
29
E. J. Routh, A Treatise on the Stability of a Given State of Motion 共Mac- 48
H. Goldstein, C. Poole, and J. Safko, Classical Mechanics, 3rd ed.
millan, London, 1877兲, p. 103; reissued as Stability of Motion, edited by 共Addison-Wesley, San Francisco, 2002兲, p. 44.
A. T. Fuller 共Taylor and Francis, London, 1975兲. 49
As discussed in Sec. VI, for a true worldline x0共t兲 or PQ, where Q is the
30
J. G. Papastavridis, “Toward an extremum characterization of kinetic sta- 共first兲 kinetic focus, we have ␦2S = 0 for one special variation 共and ␦2S
bility,” J. Sound Vib. 87, 573–587 共1983兲. ⬎ 0 for all other variations兲 as well as ␦S = 0 for all variations. Further, we
31
J. G. Papastavridis, “The principle of least action as a Lagrange varia- show that ␦3S, etc., all vanish for the special infinitesimal variation for
tional problem: Stationarity and extremality conditions,” Int. J. Eng. Sci. which ␦2S vanishes, and that S − S0 = 0 to second-order for larger such
24, 1437–1443 共1986兲; “On a Lagrangean action based kinetic instability variations. In the latter case, typically ␦3S is nonvanishing due to a single
theorem of Kelvin and Tait,” Int. J. Eng. Sci. 24, 1–17 共1986兲. coalescing alternative worldline. In atypical 共for one dimension 共Ref. 58兲兲
32
L. S. Schulman, Techniques and Applications of Path Integration 共Wiley, cases, more than one coalescing alternative worldline occurs, and the first
33
New York, 1981兲, p. 143. nonvanishing term is ␦4S or higher order—see Ref. 32, pp. 122–127. The
M. C. Gutzwiller, Chaos in Classical and Quantum Mechanics 共Springer, harmonic oscillator is a limiting case where ␦kS = 0 for all k for the special
New York, 1990兲, p. 184. variation around worldline PQ, which reflects the infinite number of true
34
In general, saddle points can be classified 共or given an index 共Ref. 27兲兲 worldlines which connect P to Q, and which can all coalesce by varying
according to the number of independent directions leading to maximum- the amplitude 共see Fig. 3兲. In Morse theory 共Refs. 27 and 34兲 the world-
type behavior. Thus the point of zero-gradient on an ordinary horse line PQ is referred to as a degenerate critical 共stationary兲 point.
saddle has a Morse index of unity. The Morse index for an action saddle 50
D. Morin, Introductory Classical Mechanics, 具www.courses.fas
point is equal to the number of kinetic foci between the end-points of the .harvard.edu/~phys16/Textbook/典 Chap. 5, p. V-8.
trajectory 共Ref. 32, p. 90兲. Readable introductions to Morse theory are 51
This statement and the corresponding one in Ref. 52 must be qualified. It
given by R. Forman, “How many equilibria are there? An introduction to is in general not simply a matter of the time interval 共tR − t P兲 being short.
Morse theory,” in Six Themes on Variation, edited by R. Hardt 共American The spatial path of the worldline must be sufficiently short. When, as
Mathematical Society, Providence, RI, 2004兲, pp. 13–36, and B. Van usually happens, more than one actual worldline can connect a given
Brunt, The Calculus of Variations 共Springer, New York, 2004兲, p. 254.
35 position x P to a given position xR in the given time interval 共tR − t P兲, for
F. Jensen, Introduction to Computational Chemistry 共Wiley, Chichester,
short time intervals only the spatially shortest worldline will have the
1999兲, Chap. 14.
36 minimum action. For example, the repulsive power-law potentials U共x兲
E. T. Whittaker, A Treatise on the Analytical Dynamics of Particles and
= C / xn 共including the limiting case of a hard-wall potential at the origin
Rigid Bodies, 4th ed. 共originally published in 1904兲 共Cambridge U. P.,
for C → 0兲 and the repulsive exponential potential U共x兲 = U0 exp共−x / a兲
Cambridge, 1999兲, p. 253. The same example was treated earlier by C. G.
have been studied 共Ref. 53兲. No matter how short the time interval 共tR
J. Jacobi, Vorlesungen Über Dynamik 共Braunschweig, Vieweg, 1884兲,
− t P兲, two different worldlines can connect given position x P to given
reprinted 共Chelsea, New York, 1969兲, p. 46.
37
A closer mechanics-optics analogy is between a kinetic focus 共mechanics兲 position xR. The fact that two different true worldlines can connect the
and a caustic point 共optics兲 共Ref. 38兲. The locus of limiting intersection two points in the given time interval leads to a kinetic focus time tQ
points of pairs of mechanical spatial orbits is termed an envelope or occurring later than tR for the shorter of the worldlines and a 共different兲
caustic 共see Fig. 7 for an example兲, just as the locus of limiting intersec- kinetic focus time tQ⬘ occurring earlier than tR for the other worldline. For
tion points of pairs of optical rays is termed a caustic. In optics, the the first worldline S is a minimum and for the other worldline S is a
intersection point of a bundle of many rays is termed a focal point; a saddle point. Another example is the quartic oscillator discussed in Sec.
mechanical analogue occurs naturally in a few systems, for example, the IX, where an infinite number of actual worldlines can connect given
sphere geodesics of Fig. 1 and the harmonic oscillator trajectories of Fig. terminal events 共x P , t P兲 and 共xR , tR兲, no matter how short the time interval
3, where a bundle of trajectories recrosses at a mechanical focal point. In 共tR − t P兲. Only for the shortest of these worldlines is S a minimum. The
electron microscopes 共Refs. 39 and 40兲, mass spectrometers 共Ref. 41兲, situation is different in 2D 共see Appendix B兲.
52
and particle accelerators 共Ref. 42兲, electric and magnetic field configura- E. T. Whittaker, Ref. 36, pp. 250–253. Whittaker deals with the Mauper-
tions are designed to create mechanical focal points. tuis action W discussed in Appendix A, whereas we adapt his analysis to
38
M. Born and E. Wolf, Principles of Optics, 4th ed. 共Pergamon, Oxford, the Hamilton action S. In more detail, our Eq. 共1兲 corresponds to the last
1970兲, pp. 130 and 734; J. A. Luck and J. H. Andrews, “Optical caustics equation on p. 251 of Ref. 36, with q1 → t, q2 → x and q⬘2 → ẋ.
53
in natural phenomena,” Am. J. Phys. 60, 397–407 共1992兲. L. I. Lolle, C. G. Gray, J. D. Poll, and A. G. Basile, “Improved short-time
39
M. Born and E. Wolf, Ref. 38, p. 738; L. A. Artsimovich and S. Yu. propagator for repulsive inverse power-law potentials,” Chem. Phys. Lett.
Lukyanov, Motion of Charged Particles in Electric and Magnetic Fields 177, 64–72 共1991兲. In Sec. X and Ref. 54 further analytical and numeri-
共MIR, Moscow, 1980兲. cal results are given for the inverse-square potential, U共x兲 = C / x2. For
40
P. Grivet, Electron Optics, 2nd ed. 共Pergamon, Oxford, 1972兲; A. L. given end positions x P and xR, there are two actual worldlines 共x P , t P兲
Hughes, “The magnetic electron lens,” Am. J. Phys. 9, 204–207 共1941兲; → 共xR , tR兲 for given short times 共tR − t P兲. There is one actual worldline for
J. H. Moore, C. C. Davis, and M. A. Coplan, Building Scientific Appara- tR = tQ when the two worldlines have coalesced into one, and there is no
tus, 3rd ed. 共Perseus, Cambridge, MA, 2003兲, Chap. 5. actual worldline for longer times 共remember x P and xR are fixed兲.
41 54
P. Grivet, Ref. 40, p. 822; J. H. Moore et al., Ref. 40. A. G. Basile and C. G. Gray, “A relaxation algorithm for classical paths
42
M. S. Livingston, The Development of High-Energy Accelerators 共Dover, as a function of end points: Application to the semiclassical propagator
New York, 1966兲; M. L. Bullock, “Electrostatic strong-focusing lens,” for far-from-caustic and near-caustic conditions,” J. Comput. Phys. 101,
Am. J. Phys. 23, 264–268 共1955兲; L. W. Alvarez, R. Smits, and G. 80–93 共1992兲.
55
Senecal, “Mechanical analogue of the synchrotron, illustrating phase sta- Better estimates can be found using the Sturm and Sturm-Liouville theo-
bility and two-dimensional focusing,” Am. J. Phys. 43, 292–296 共1975兲; ries; see Papastavridis, Refs. 26 and 66.
A. Chao et al., “Experimental investigation of nonlinear dynamics in the 56
This argument can be refined. In Sec. 8 we show that ␦2S becomes O共␣3兲
Fermilab Tevatron,” Phys. Rev. Lett. 61, 2752–2755 共1988兲. For recent for R → Q but is still larger in magnitude than the ␦3S term, which is also
texts, see A. W. Chao, “Resource Letter PBA-1: Particle beams and ac- O共␣3兲. See Ref. 60 for ␦2S and Eq. 共40a兲 for ␦3S.
celerators,” Am. J. Phys. 74, 855–862 共2006兲. 57
There are other systems for which ⳵3U / ⳵x3, etc., vanish, for example,
43
V. G. Boltyanskii, Envelopes 共MacMillan, New York, 1964兲. U共x兲 = C, U共x兲 = Cx, U共x兲 = −Cx2. For these systems the worldlines cannot
44
P. T. Saunders, An Introduction to Catastrophe Theory 共Cambridge U. P., have ␦2S = 0 关see Eq. 共19兲兴, so that kinetic foci do not exist.
Cambridge, 1980兲, p. 62. 58
In higher dimensions more than one independent variation, occurring in
45
Systems with subsequent kinetic foci are discussed in Secs. VIII and IX. different directions in function space, can occur due to symmetry. In
For examples with only a single kinetic focus, see Figs. 6 and 7. Morse theory the number of these independent variations satisfying ␦2S
46
M. C. Gutzwiller, “The origins of the trace formula,” in Classical, Semi- = 0 is called the multiplicity of the kinetic focus 共Van Brunt, Ref. 34, p.
classical and Quantum Dynamics in Atoms, edited by H. Friedrich and B. 254; Ref. 32, p. 90兲. If the mulitplicity is different from unity, Morse’s
Eckhardt 共Springer, New York, 1997兲, pp. 8–28. theorem is modified from the statement in Ref. 34 to read as follows: The
47
This type of variation, ␦x = ␣␾, ␦ẋ = ␣␾
˙ , where ␦x and ␦ẋ vanish together Morse index of the saddle point in action of worldline PR is equal to the

455 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 455
number of kinetic foci between P and R, with each kinetic focus counted on ‘A constant force generator for the demonstration of Newton’s second
with its multiplicity. law’,” Am. J. Phys. 52, 268 共1984兲.
59 73
J. M. T. Thompson and G. W. Hunt, Elastic Instability Phenomena C. G. Gray, G. Karl, and V. A. Novikov, “Progress in classical and
共Wiley, Chichester, 1984兲, p. 20. quantum variational principles,” Rep. Prog. Phys. 67, 159–208 共2004兲.
60
The fact that ␦2S0 → 0 for ␣ → 0 is not surprising because ␦2S0 is propor- 74
Nearly pure quartic potentials have been found in molecular physics for
tional to ␣2. The surprising fact is that here ␦2S0 vanishes as ␣3 as ␣ ring-puckering vibrational modes 共Ref. 75兲 and for the caged motion of
→ 0 because the integral involved in the definition, Eq. 共36a兲, of ␦2S0 is the potassium ion K+ in the endohedral fullerene complex K+ @ C60 共Ref.
itself O共␣兲. We can see directly that ␦2S0 becomes O共␣3兲 for R near Q for 76兲, where the quadratic terms in the potential are small. Ferroelectric
this special variation ␣␾ by integrating the ␾ ˙ term by parts in Eq. 共36a兲 soft modes in solids are also sometimes approximately represented by
and using ␾ = 0 at the end-points. The result is ␦2S0 = −共␣2 / 2兲兰RP␾关m␾ ¨ quartic potentials 共Refs. 77 and 78兲.
75
+ U⬙共x0兲␾兴 dt. Because x0 and x1 = x0 + ␣␾ are both true worldlines, we R. P. Bell, “The occurrence and properties of molecular vibrations with
can apply the equation of motion mẍ + U⬘共x兲 = 0 to both. We then subtract V共x兲 = ax4,” Proc. R. Soc. London, Ser. A 183, 328–337 共1945兲; J. Laane,
these two equations of motion and expand U⬘共x0 + ␣␾兲 as U⬘共x0兲 “Origin of the ring-puckering potential energy function for four-
+ U⬙共x0兲␣␾ + 共 2 兲U⵮共x0兲共␣␾兲2 + O共␣3兲, m␾¨ + U⬙共x0兲␾ membered rings and spiro compounds. A possibility of pseudorotation,”
1
giving
J. Phys. Chem. 95, 9246–9249 共1991兲.
= −共 2 兲U⵮共x0兲␣␾2 + O共␣2兲. 共If the ␾2, etc., nonlinear terms on the right-
1
76
C. G. Joslin, J. Yang, C. G. Gray, S. Goldman, and J. D. Poll, “Infrared
hand side are neglected in the last equation, it becomes the Jacobi-
rotation and vibration-rotation bands of endohedral Fullerene complexes.
Poincaré linear variation equation used in stability studies.兲 If we use this
result in the previous expression for ␦2S0, we find to lowest nonvanishing K+ @ C60.” Chem. Phys. Lett. 211, 587–594 共1993兲.
77
order ␦2S0 = 共␣3 / 4兲兰RPdtU⵮共x0兲␾3, which is O共␣3兲. This result and Eq. A. S. Barker, “Far infrared dispersion and the Raman spectra of ferro-
共40a兲 for ␦3S0 give the desired result 共42兲 for S1 − S0. electric crystals,” in Far-Infrared Properties of Solids, edited by S. S.
61
If we use arguments similar to those of this section and Sec. VI, we can Mitra and S. Nudelman 共Plenum, New York, 1970兲, pp. 247–296.
78
show that ␦2S vanishes again at the second kinetic focus Q2, and that for J. Thomchick and J. P. McKelvey, “Anharmonic vibrations of an ‘ideal’
R beyond Q2 the wordline PR has a second, independent variation lead- Hooke’s law oscillator,” Am. J. Phys. 46, 40–45 共1978兲.
79
ing to ␦2S ⬍ 0, in agreement with Morse’s general theory 共Ref. 34兲. R. Baierlein, Ref. 16, p. 73.
80
62
If we use L共x , ẋ兲 = pẋ − H共x , p兲, we can rewrite the Hamilton action as a In two dimensions with x = 共x1 , x2兲 and v0 = 共v01 , v02兲, our analytic condi-
phase-space integral, that is, S = 兰RP关pẋ − H共x , p兲兴 dt. We set ␦S = 0 and tion 共4兲 for the kinetic focus of worldline x共t , v0兲 becomes det共⳵xi / ⳵v0j兲
vary x共t兲 and p共t兲 independently and find 共Ref. 63兲 the Hamilton equa- = 0, where det共Aij兲 denotes the determinant of matrix Aij. The generaliza-
tions of motion ẋ = ⳵H / ⳵ p, ṗ = −⳵H / ⳵x. We can then show 共Ref. 64兲 that tion to other dimensions is obvious. This condition 共in slightly different
in phase space, the trajectories x共t兲, p共t兲 that satisfy the Hamilton equa- form兲 is due to Mayer, Ref. 19, p. 269. For a clear discussion, see J. G.
tions are always saddle points of S, that is, never a true maximum or a Papastavridis, Analytical Mechanics 共Oxford U. P., Oxford, 2002兲, p.
true minimum. In the proof it is assumed that H has the normal form 1061. For multidimensions a caustic becomes in general a surface in
H共x , p兲 = p2 / 2m + U共x兲. space-time. The analogous theory for multidimensional spatial caustics,
63
Reference 48, p. 353. relevant for the action W, is discussed in Ref. 105. A simple example of
64
M. R. Hestenes, “Elements of the calculus of variations,” in Modern a 2D surface spatial caustic is obtained by revolving the pattern of Fig. 7
Mathematics for the Engineer, edited by E. F. Beckenbach 共McGraw Hill, about the vertical axis, thereby generating a paraboloid of revolution
New York, 1956兲, pp. 59–91. surface caustic/envelope. Due to axial symmetry, the caustic has a second
65
O. Bottema, “Beisipiele zum Hamiltonschen Prinzip,” Monatsh. Math. 共linear兲 branch, that is, the symmetry axis from y = 0 to y = Y. An analo-
66, 97–104 共1962兲. gous optical example is discussed by M. V. Berry, “Singularities in waves
66
J. G. Papastavridis, “On the extremal properties of Hamilton’s action and rays,” in Physics of Defects, Les Houches Lectures XXXIV, edited by
integral,” J. Appl. Mech. 47, 955–956 共1980兲. R. D. Balian, M. Kleman, and J.-P. Poirier 共North Holland, Amsterdam,
67
C. W. Misner, K. S. Thorne, and J. A. Wheeler, Gravitation 共Freeman, 1981兲, pp. 453–543.
81
San Francisco, 1973兲, p. 318. These authors use the Rayleigh-Ritz direct A dynamics problem can be formulated as an initial value problem. For
variational method 共see Ref. 12 for a detailed discussion of this method兲 example, find x共t兲 from Newton’s equation of motion with initial condi-
with a two-term trial trajectory x共t兲 = a1 sin 共␻t / 2兲 + a2 sin 共␻t兲, where ␻ tions 共x P , ẋ P兲. It can also be formulated as a boundary value problem; for
= 2␲ / tR and a1 and a2 are variational parameters, to study the half-cycle example, find x共t兲 from Hamilton’s principle with boundary conditions
共tR = T0 / 2兲 and one-cycle 共tR = T0兲 trajectories. Because the kinetic focus 共x P , t P兲 and 共xR , tR兲. Solving a boundary value problem with initial value
time tQ ⬎ T0 / 2 for this oscillator, they find, in agreement with our results, problem methods 共for example, the shooting method兲 is standard 共Ref.
that S is a minimum for the half-cycle trajectory 共with a1 ⫽ 0, a2 = 0兲 and 82兲. Solving an initial value problem with boundary value problem meth-
a saddle point for the one-cycle trajectory 共with a1 = 0, a2 ⫽ 0兲. However, ods is much less common 共Ref. 83兲. For an example of a boundary value
in the figure accompanying their calculation, which shows the stationary problem with mixed conditions 共prescribed initial velocities and final
points in 共a1 , a2兲 space, they label the origin 共a1 , a2兲 = 共0 , 0兲 a maximum. positions兲 for about 107 particles, see A. Nusser and E. Branchini, “On
The point 共a1 , a2兲 = 共0 , 0兲 represents the equilibrium trajectory x共t兲 = 0. As the least action principle in cosmology,” Mon. Not. R. Astron. Soc. 313,
we have seen, a true maximum in S cannot occur, so that other “direc- 587–595 共2000兲.
82
tions” in function space not considered by the authors must give See, for example, W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B.
minimum-type behavior of S, leading to an overall saddle point. P. Flannery, Numerical Recipes in Fortran, 2nd ed. 共Cambridge U. P.,
68
The two-incline oscillator potential has the form U共x兲 = C 兩 x兩, with C Cambridge, 1992兲, p. 749.
83
= mg sin ␣ cos ␣ and ␣ the angle of inclination. Here x is a horizontal H. R. Lewis and P. J. Kostelec, “The use of Hamilton’s principle to derive
direction. A detailed discussion of this oscillator is given by B. A. Sher- time-advance algorithms for ordinary differential equations,” Computer
wood, Notes on Classical Mechanics 共Stipes, Champaign, Il, 1982兲, p. Phys. Commun. 96, 129–151 共1996兲; D. Greenspan, “Approximate solu-
157. tion of initial value problems for ordinary differential equations by
69
Other constant force or linear potential systems include the 1D Coulomb boundary value techniques,” J. Math. Phys. Sci. 15, 261–274 共1967兲.
84
model 共Ref. 70兲 U共x兲 = q1q2 兩 x兩, the bouncing ball 共Ref. 71兲, and for x The converse effect cannot occur: a time-dependent potential U共x , t兲 with
艌 0 the constant force spring 共Ref. 72兲. U⬙ ⬍ 0 at all times always has ␦2S ⬎ 0 as seen from Eq. 共19兲. If U共x , t兲 is
70
I. R. Lapidus, “One- and two-dimensional Hydrogen atoms,” Am. J. such that U⬙ alternates in sign with time, kinetic foci 共and hence trajec-
Phys. 49, 807 共1981兲. K. Andrew and J. Supplee, “A Hydrogen atom in tory stability兲 may occur. An example is a pendulum with a rapidly ver-
d-dimensions,” Am. J. Phys. 58, 1177–1183 共1990兲. tically oscillating support point. In effect the gravitational field is oscil-
71
I. R. Gatland, “Theory of a nonharmonic oscillator,” Am. J. Phys. 59, lating. The pendulum can oscillate stably about the 共normally unstable兲
155–158 共1991兲 and references therein; W. M. Hartmann, “The dynami- upward vertical direction 共Ref. 85兲. Two- and three-dimensional ex-
cally shifted oscillator,” Am. J. Phys. 54, 28–32 共1986兲. amples of this type are Paul traps 共Ref. 86兲 and quadrupole mass filters
72
A. Capecelatro and L. Salzarulo, Quantitative Physics for Scientists and 共Ref. 85兲, which use oscillating quadrupole electric fields to trap ions.
Engineers: Mechanics 共Aurie Associates, Newark, NJ, 1977兲, p. 162; The equilibrium trajectory x共t兲 = 0 at the center of the trap is unstable for
C.-Y. Wang and L. T. Watson, “Theory of the constant force spring,” purely electrostatic fields but is stabilized by using time-dependent elec-
Trans. ASME, J. Appl. Mech. 47, 956–958 共1980兲; H. Helm, “Comment tric fields. Focusing by alternating-gradients 共also known as strong focus-

456 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 456
ing兲 in particle accelerators and storage rings is based on the same idea the two forms for W differ from each other 共Refs. 30 and 98兲 and from
共Ref. 42兲. those for S.
100
85
M. H. Friedman, J. E. Campana, L. Kelner, E. H. Seeliger, and A. L. E. J. Routh, A Treatise on Dynamics of a Particle 共Cambridge U. P.,
Yergey, “The inverted pendulum: A mechanical analog of the quadrupole Cambridge, 1898兲, reprinted 共Dover, New York, 1960兲, p. 400.
101
mass filter,” Am. J. Phys. 50, 924–931 共1982兲. A. P. French, “The envelopes of some families of fixed-energy trajecto-
86
P. K. Gosh, Ion Traps 共Oxford U. P., Oxford, 1995兲, p. 7. ries,” Am. J. Phys. 61, 805–811 共1993兲; E. I. Butikov, “Families of
87
J. J. Stoker, Nonlinear Vibrations in Mechanical and Electrical Systems Keplerian orbits,” Eur. J. Phys. 24, 175–183 共2003兲.
102
共Wiley, New York, 1950兲, p. 112. Stoker’s statements on series conver- We assume that we are dealing with bound orbits. Similar comments
gence need amendment in light of the Kolmogorov-Arnold-Moser apply to scattering orbits 共hyperbolas兲. Just as for the orbits in the linear
共KAM兲 theory 共Ref. 89兲. See J. Moser, “Combination tones for Duffing’s gravitational potential discussed in the preceding paragraph, here too
equation,” Commun. Pure Appl. Math. 18, 167–181 共1965兲; T. Kapita- there are restrictions and special cases 共Ref. 19, p. 164; Ref. 103, p. 122兲.
niak, J. Awrejcewicz and W.-H. Steeb, “Chaotic behaviour in an anhar- If the second point 共xR , y R兲 lies within the “ellipse of safety” 共the enve-
monic oscillator with almost periodic excitation,” J. Phys. A 20, L355– lope 共French, Ref. 101兲兲 of the elliptical trajectories of energy E originat-
L358 共1987兲; A. H. Nayfeh, Introduction to Perturbation Techniques ing at 共x P , y P兲, then two ellipses with energy E can connect 共x P , y P兲 to
共Wiley, New York, 1981兲, p. 216; A. H. Nayfeh and B. Balachandran, 共xR , y R兲. If 共xR , y R兲 lies on the ellipse of safety, then one ellipse of energy
Applied Nonlinear Dynamics 共Wiley, New York, 1995兲, p. 234; S. Wig- E can connect 共x P , y P兲 to 共xR , y R兲, and if 共xR , y R兲 lies outside the ellipse of
gins, “Chaos in the quasiperiodically forced Duffing oscillator,” Phys. safety, then no ellipse of energy E can connect the two points. Usually the
Lett. A 124, 138–142 共1987兲. initial and final points 共x P , y P兲 and 共xR , y R兲 together with the center of
88
G. Seifert, “On almost periodic solutions for undamped systems with force at 共0,0兲 共one focus of the elliptical path兲 define the plane of the
almost periodic forcing,” Proc. Am. Math. Soc. 31, 104–108 共1972兲; J. orbit. If 共x P , y P兲, 共xR , y R兲, and 共0,0兲 lie on a straight line, the plane of the
Moser, “Perturbation theory of quasiperiodic solutions and differential orbit is not uniquely defined, and there is almost always an infinite num-
equations,” in Bifurcation Theory and Nonlinear Eigenvalue Problems, ber of paths of energy E in three dimensions that can connect
edited by J. B. Keller and S. Antman 共Benjamin, New York, 1969兲, pp. 共x P , y P , z P = 0兲 to 共xR , y R , zR = 0兲. A particular case of the latter is a periodic
283–308; J. Moser, “Perturbation theory for almost periodic solutions for orbit where 共xR , y R兲 = 共x P , y P兲. Because the orbit can now be brought into
undamped nonlinear differential equations,” in International Symposium coalescence with an alternative true orbit by a rotation around the line
on Nonlinear Differential Equations and Nonlinear Mechanics, edited by joining 共x P , y P兲 to 共0,0兲, a third kinetic focus arises for elliptical periodic
J. P. Lasalle and S. Lefschetz 共Academic, New York, 1963兲, pp. 71–79; orbits in three dimensions 共see Ref. 33, p. 29兲.
103
M. S. Berger, “Two new approaches to large amplitude quasi-periodic N. G. Chetaev, Theoretical Mechanics 共Springer, Berlin, 1989兲.
104
motions of certain nonlinear Hamiltonian systems,” Contemp. Math. For the repulsive 1 / r potential, the hyperbolic spatial orbits have a 共para-
108, 11–18 共1990兲. bolic shaped兲 caustic/envelope 共French, Ref. 101兲.
89
G. M. Zaslavsky, R. Z. Sagdeev, D. A. Usikov, and A. A. Chernikov,
105
If the orbit equation has the explicit form y = y共x , ␪0兲, or the implicit form
Weak Chaos and QuasiRegular Patterns 共Cambridge U. P., Cambridge, f共x , y , ␪0兲 = 0, the spatial kinetic focus is found from ⳵y / ⳵␪0 = 0 or
1991兲, p. 30. ⳵ f / ⳵␪0 = 0, respectively. Here ␪0 is the launch angle at 共x P , y P兲 共see Fig.
90
See, for example, M. Tabor, Chaos and Integrability in Nonlinear Dy- 15 for an example兲. The derivation of these spatial kinetic focus condi-
namics 共Wiley, New York, 1989兲, p. 35; J. M. T. Thompson and H. B. tions is similar to the derivation of the space-time kinetic focus condition
Stewart, Nonlinear Dynamics and Chaos, 2nd ed. 共Wiley, Chichester, of Eq. 共4兲 共see Ref. 106, p. 59兲. In contrast, if the orbit equation is defined
2002兲, pp. 310. parametrically by the trajectory equations x = x共t , ␪0兲 and y = y共t , ␪0兲, the
91
For example, the equilibrium position can be modulated. A somewhat spatial kinetic focus condition is ⳵共x , y兲 / ⳵共t , ␪0兲 = 0. This Jacobian deter-
similar system is a ball bouncing on a vertically oscillating table. The minant condition is similar to that of Ref. 80 for the space-time kinetic
motion can be chaotic. See, for example, J. Guckenheimer and P. Holmes, focus 共see Ref. 106, p. 73 for a derivation兲. As an example, consider a
Nonlinear Oscillations, Dynamical Systems and Bifurcations of Vector family of figure-eight-like harmonic oscillator orbits of Fig. 16, launched
Fields 共Springer, New York, 1983兲, p. 102; N. B. Tufillaro, T. Abbott, and from the origin 共x P , y P兲 = 共0 , 0兲 at time t P = 0, all with speed v0 共and there-
fore the same energy E兲, at various angles ␪0. The trajectory equations are
J. Reilly, An Experimental Approach to Nonlinear Dynamics and Chaos
x = 共v0 / ␻1兲cos ␪0 sin ␻1t and y = 共v0 / ␻2兲sin ␪0 sin ␻2t, where ␻1 = 2␻2.
共Addison-Wesley, Redwood City, CA, 1992兲, p. 23; A. B. Pippard, The
The determinant condition for the spatial kinetic focus reduces to
Physics of Vibration 共Cambridge U. P., Cambridge, 1978兲, Vol. 1, pp.
cos ␪0 tan ␻2t = 1, which locates the kinetic focus 共in time兲 for the orbit
253, 271.
with launch angle ␪0. Elimination of ␪0 and t from these three equations
The forced Duffing oscillator with U共x , t兲 = 共 2 兲kx2 + Cx4 − xF0 cos ␻t is
92 1
leads to the locus of the 共first兲 spatial kinetic foci, the spatial caustic/
studied in Ref. 30. For k = 0 the Duffing oscillator reduces to the quartic
envelope equation y 2 = 共v0 / ␻2兲2 − 2共v0 / ␻2兲 兩 x兩, which is a parabolic
oscillator. shaped curve with two cusps on the y-axis.
93
R. H. G. Helleman, “Variational solutions of non-integrable systems,” in 106
R. H. Fowler, The Elementary Differential Geometry of Plane Curves
Topics in Nonlinear Dynamics, edited by S. Jorna 共AIP, New York, 1978兲, 共Cambridge U. P., Cambridge, 1920兲.
pp. 264–285. This author studies the forced Duffing oscillator with 107
The finding of the two elliptical 共or hyperbolic or parabolic兲 shaped tra-
U共x , t兲 = 共 2 兲kx2 − Cx4 − xF0 cos ␻t 共note the sign change in C compared to
1
jectories from observations giving the two end-positions and the time
Ref. 92兲, and the Henon-Heiles oscillator with the potential in Eq. 共83兲. interval is a famous problem of astronomy and celestial mechanics,
In Ref. 54 the harmonic potential U共x , t兲 = 共 2 兲k关x − xc共t兲兴2 with an oscillat-
94 1
solved by Lambert 共1761兲, Gauss 共1801–1809兲, and others 共Ref. 108兲.
ing equilibrium position xc共t兲 is studied. The worldlines for this system 108
R. R. Bate, D. D. Mueller, and J. E. White, Fundamentals of Astrody-
are all nonchaotic. namics 共Dover, New York, 1971兲, p. 227; H. Pollard, Celestial Mechanics
95
M. Henon and C. Heiles, “The applicability of the third integral of the 共Mathematical Association of America, Washington, 1976兲, p. 28; P. R.
motion: Some numerical experiments,” Astron. J. 69, 73–79 共1964兲. Escobal, Methods of Orbit Determination 共Wiley, New York, 1965兲, p.
96
There have been a few formal studies of action for chaotic systems, but 187. For the elliptical orbits, more than two trajectories typically become
few concrete examples seem to be available. See, for example, S. Bolotin, possible at sufficiently large time intervals; these additional trajectories
“Variational criteria for nonintegrability and chaos in Hamiltonian sys- correspond to more than one complete revolution along the orbit 共Ref.
tems,” in Hamiltonian Mechanics, edited by J. Seimenis 共Plenum, New 109兲.
109
York, 1994兲, pp. 173–179. R. H. Gooding, “A procedure for the solution of Lambert’s orbital
97
Reference 48, p. 434. boundary-value problem,” Celest. Mech. Dyn. Astron. 48, 145–165
98
H. Poincaré, Les Méthodes Nouvelles de la Mécanique Céleste 共Gauthier- 共1990兲.
Villars, Paris, 1899兲, Vol. 3; New Methods of Celestial Mechanics 共AIP, 110
It is clear from Fig. 13 that a kinetic focus occurs after time T0. To show
New York, 1993兲, Part 3, p. 958. rigorously that this focus is the first kinetic focus 共unlike for W where it
99
The situation is complicated because, as Eq. 共A1兲 shows, there are two is the second兲, we can use a result of Gordon 共Ref. 111兲 that the action S
forms for W, that is, the time-independent 共first兲 form and the time- is a minimum for time t = T0. If one revolution corresponds to the second
dependent 共last兲 form. Spatial kinetic foci 共discussed in Appendix B兲 kinetic focus, the trajectory P → P would correspond to a saddle point.
occur for the time-independent form of W. Space-time kinetic foci occur The result tQ = T0 can also be obtained algebraically by applying the gen-
for the time-dependent form of W, as for S. Typically the kinetic foci for eral relation 共4兲 to the relation r = r共t , L兲 for the radial distance, where we

457 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 457
114
use angular momentum L as the parameter labeling the various members J. C. Slater and N. H. Frank, Introduction to Theoretical Physics
of the family in Fig. 13. We obtain tQ from 共⳵r / ⳵L兲t = 0. The latter equa- 共McGraw Hill, New York, 1933兲, p. 85.
tion implies that 共⳵t / ⳵L兲r = 0, because 共⳵r / ⳵L兲t = −共⳵r / ⳵t兲L共⳵t / ⳵L兲r. At 115
C. G. Gray, G. Karl, and V. A. Novikov, “The four variational principles
fixed energy E 共or fixed major axis 2a兲, the period T0 is independent of L of mechanics,” Ann. Phys. 共N.Y.兲 251, 1–25 共1996兲.
for the attractive 1 / r potential, so that the solution of 共⳵t / ⳵L兲r = 0 occurs 116
C. G. Gray, G. Karl, and V. A. Novikov, “From Maupertuis to
for t = T0, which is therefore the kinetic focus time tQ. Schrödinger. Quantization of classical variational principles,” Am. J.
111
W. B. Gordon, “A minimizing property of Keplerian orbits,” Am. J.
Phys. 67, 959–961 共1999兲.
Math. 99, 961–971 共1977兲. 117
112 E. Schrödinger, “Quantisierung als eigenwert problem I,” Ann. Phys. 79,
Note that for the actual 2D trajectories in the potential U共x , y兲 = mgy,
kinetic foci exist for the spatial paths of the Maupertuis action W, but do 361–376 共1926兲, translated in E. Schrödinger, Collected Papers on Wave
not exist for the space-time trajectories of the Hamilton action S. This Mechanics 共Blackie, London, 1928兲, Chelsea reprint 1982. For modern
result illustrates the general result stated in Appendix A that the kinetic discussions and applications, see, for example, E. Merzbacher, Quantum
foci for W and S differ in general. Mechanics, 3rd ed. 共Wiley, New York, 1998兲, p. 135; S. T. Epstein, The
113
A. P. French, Vibrations and Waves 共Norton, New York, 1966兲, p. 36. Variation Method in Quantum Chemistry 共Academic, New York, 1974兲.

458 Am. J. Phys., Vol. 75, No. 5, May 2007 C. G. Gray and E. F. Taylor 458

You might also like