Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

EDUCATION Revista Mexicana de Fı́sica E 17 (1) 47–54 JANUARY–JUNE 2020

Power and beauty of the Lagrange equations

L. de la Peña, A. M. Cetto and A. Valdés-Hernández


Instituto de Fı́sica, Universidad Nacional Autónoma de México,
Apartado Postal 20-364, Ciudad de México, Mexico.

Received 14 August 2019; accepted 30 October 2019

The Lagrangian formulation of the equations of motion for point particles is usually presented in classical mechanics as the outcome of a
series of insightful algebraic transformations or, in more advanced treatments, as the result of applying a variational principle. In this paper
we stress two main reasons for considering the Lagrange equations as a fundamental description of the dynamics of classical particles. Firstly,
their structure can be naturally disclosed from the existence of integrals of motion, in a way that, though elementary and easy to prove, seems
to be less popular —or less frequently made explicit— than others in support of the Lagrange formulation. The second reason is that the
Lagrange equations preserve their form in any coordinate system —even in moving ones, if required. Their covariant nature makes them
particularly suited to deal with dynamical problems in curved spaces or involving (holonomic) constraints. We develop the above and related
ideas in clear and simple terms, keeping them throughout at the level of intermediate courses in classical mechanics. This has the advantage
of introducing some tools and concepts that are useful at this stage, while they may also serve as a bridge to more advanced courses.

Keywords: Lagrange equations; classical mechanics; covariance; integrals of motion.

PACS: 45.20.Jj; 45.20.D-; 01.40.-d DOI: https://doi.org/10.31349/RevMexFisE.17.47

1. Introduction gant and powerful demand that a function called action attain
a minimum value along the trajectory followed by the parti-
cle [1,3]. The Hamiltonian formulation, in its turn, represents
Classical dynamics has a mighty range of tools to describe a further elaboration providing a no less elegant set of differ-
the motion of particles; among them we find in elementary ential equations. However, usually these derivations make
treatments the forces, potentials, constants of motion, sys- use of the more advanced calculus of variations, with which
tems of reference and, in a prominent place, the equations we assume the reader-student not to be acquainted.
of motion themselves. The first contact with the latter is un-
doubtedly through Newton’s widely-known second-order dif- In this paper we arrive at the Lagrange equations by fol-
ferential equation that relates accelerations with forces. It is lowing a procedure that allows us to expose in clear and sim-
of interest to recall that the invention of differential calcu- ple terms their connection with the integrals of motion —par-
lus by Newton was driven precisely by the need of a tool to ticularly, the energy— as well as their covariance, i.e., their
describe the motion of bodies. Usually it is only in more ad- form-invariance with respect to a change of coordinate sys-
vanced courses that the student learns about other approaches tem. An additional benefit of this alternative route is that it
serving the same purpose, in particular the Lagrangian for- allows us to disclose in a natural way the general structure
mulation (established about a century after Newton’s), and of the Lagrange equations. Our intention is to develop the
the Hamiltonian formulation (developed about half a century above ideas (and other related ones) in simple terms, work-
later). We thus have at least three different sets of equa- ing all the time at the level of intermediate courses in classical
tions, all of them fundamental, and equivalent in that they mechanics.
can be used alternatively to describe the behavior of the same
classical system. However, what is considered fundamental The paper is structured as follows. In Sec. 2 the Lagrange
depends on the intended descriptive level. From a histori- equations for a conservative system are derived, and the me-
cal perspective, Newton’s Second Law is fundamental, the chanical energy is shown to be the integral of motion asso-
others having been derived from it with the introduction of ciated with them. In Sec. 3 the invariance of the Lagrange
more elaborate principles and demands. In particular, the equations under a transformation of the coordinate system is
Lagrange equations have a most remarkable and important discussed, and the covariant form of Newton’s Second Law is
property, namely that they preserve their form in any system derived. Special attention is paid to the emergence of inertial
of coordinates, even moving ones [1, 2]. This form-invariant forces as a result of the curvature of the coordinates (curvi-
property reflects the fundamental fact that they express a law linear coordinates). A couple of illustrative examples are pre-
of nature, which is naturally independent of our description. sented in Sec. 3.4. Finally, Sec. 4 establishes the connection
Also interesting is the fact —expressing another law of na- between the Lagrangian and the Hamiltonian functions, and
ture— that the Lagrange equations result from the single ele- briefly introduces Hamilton’s equations.
48 L. DE LA PEÑA, A.M. CETTO AND A. VALDÉS-HERNÁNDEZ

2. The Lagrange equations and integrals of The above equations establish a correspondence between
motion the constant of motion ξ and the vector G(ξ) , which, accord-
ing to Eq. (6), is either zero or orthogonal to the velocity q̇
2.1. Integrals of motion along the trajectory. Equation (7) exhibits in a natural way the
characteristic structure of the Lagrange equations. Indeed, in
For the description of the motion of a particle it is often the following section we will see that an analysis of the com-
convenient to use a system of (generalized) coordinates {qi } (ξ)
ponents Gi takes us directly to the Lagrange equations, re-
suited for the particular problem, instead of Cartesian coordi- vealing their equivalence with Newton’s Second Law, and to
nates {xi }. In what follows we shall consider such a system the identification of the mechanical energy as the associated
{qi }, which is arbitrary except for the condition that the coor- integral of motion.
dinates can be expressed as regular, continuous and invertible
It is worthwhile to recall at this point that a closed me-
functions of the Cartesian coordinates {xi }, so that
chanical system with N degrees of freedom may have at most
qi = qi ({xk }), xi = xi ({ql }). (1) 2N − 1 nontrivial and functionally independent constants of
motion, that is, functions of q, q̇ and t whose value does not
This is an instance of a passive transformation [1], i.e., one change with time. The complete set of these constants de-
involving a change of coordinate system only, without affect- termines the trajectory of the particle in the 2N -dimensional
ing in any way the physical system. For simplicity we shall phase space. Among the possible constants of motion there
omit the appearance of the time as a parameter in the transfor- are some that have a particular importance, the so-called inte-
mation; this restricts our discussion to point transformations, grals of motion. These are (continuous, single-valued, differ-
that is, to purely geometric transformations [1-6]. entiable) time-independent functions of the generalized co-
Let us consider a differentiable scalar function ϕ(q, q̇), ordinates and their corresponding momenta (or velocities),
with q̇ = dq/dt and q the vector with N components {qi }, defined over the entire accessible phase space and having a
N being the number of degrees of freedom. As a first step constant value along the trajectory. Thus each one serves in
we analyze the time derivative of ϕ, which we obtain apply- principle to eliminate a degree of freedom from the descrip-
ing the chain rule, tion.
· ¸
dϕ X ∂ϕ ∂ϕ
= q̇i + q̈i 2.2. The Lagrange equations of motion
dt i
∂qi ∂ q̇i
X · ∂ϕ d ∂ϕ
µ ¶
d ∂ϕ
¸
Let us consider a particle of mass m under the action of
= q̇i + q̇i − q̇i . (2) a time-independent, conservative force f (x) = −∇V (x),
i
∂qi dt ∂ q̇i dt ∂ q̇i
with V (x) the potential energy. In Cartesian coordinates the
A rearrangement of terms leads to the expression mechanical energy of this particle reads
à !
d X ∂ϕ X µ d ∂ϕ ∂ϕ

E = 12 mẋ2 + V (x). (8)
q̇i −ϕ = q̇i − , (3)
dt i
∂ q̇i i
dt ∂ q̇i ∂qi
According to our previous results, for this E(x, ẋ) we can
which can be written in the more compact form construct a ϕ(ξ) such that
dγ X
= q̇i Gi (4) X ∂ϕ(ξ)
dt E= ẋi − ϕ(ξ) . (9)
i
i
∂ ẋi
with
X The general solution ϕ(ξ) (x, ẋ) of this differential equation
∂ϕ d ∂ϕ ∂ϕ
γ(q, q̇)≡ q̇i − ϕ, Gi (q, q̇)≡ − . (5) reads
∂ q̇i dt ∂ q̇i ∂qi
i X X
ϕ(E) (x, ẋ) = 21 m ẋ2i − V (x) + Qi (x)ẋi , (10)
Assume now that the function ϕ is selected such that the
i i
corresponding γ does not evolve in time, i.e., dγ/dt = 0. We
denote such constant function with ξ, and the corresponding where the Qi are arbitrary functions of x only. From Eqs. (7)
ϕ with ϕ(ξ) . Equations (4) and (5) give then and (10) we obtain for the components Gi
(E)

dξ X µ ∂Qi ¶
= 0 = q̇ · G(ξ) , (6) (E) ∂Qj
dt Gi = mẍi − fi + − ẋj (11)
j
∂xj ∂xi
(ξ)
where G is the vector with components
dQi ∂ X
d ∂ϕ(ξ) ∂ϕ(ξ) = mẍi − fi + − Qj ẋj . (12)
Gi
(ξ)
= − . (7) dt ∂xi j
dt ∂ q̇i ∂qi

Rev. Mex. Fı́s. E 17 (1) 47–54


POWER AND BEAUTY OF THE LAGRANGE EQUATIONS 49

On the other hand, by separating the terms in Eq. (10) Further (see Eq. (11)),
that encode the dynamical information from the auxiliary Qi ,
(E) X
(E) µ ¶
ϕ(E) can be rewritten in the form d ∂ϕK ∂ϕK ∂Qi ∂Qj
X − = − ẋj = 0, (23)
dt ∂ ẋi ∂xi ∂xj ∂xi
ϕ(E) = L(x, ẋ) + Qi (x)ẋi , (13) j
i
a result that discloses the well-known and very useful prop-
with X erty of invariance of the Lagrange equations with respect to
L(x, ẋ) = 21 m ẋ2i − V (x). (14) the addition of a total time derivative of a scalar function of
i
x (and t, we may add),
We thus obtain using Eq. (7)
dK (E)
(E) d ∂ϕ(E) ∂ϕ(E) L → L0 = L + = ϕK , (24)
Gi = − (15) dt
dt ∂ ẋi ∂xi
µ ¶ with K(x, t) freely selected [1, 2, 4, 6]. L and L0 are equiv-
d ∂L ∂L dQi ∂ X
= − + − Qj ẋj . (16) alent Lagrangians, meaning that the presence of K is totally
dt ∂ ẋi ∂xi dt ∂xi j
inconsequential, since it does not affect the integrals of mo-
Comparison of this with Eq. (12) gives tion in any way. Thus there is an infinity of Lagrangians
associated with each mechanical system. This gives us the
d ∂L ∂L freedom to select one that is appropriate for our purposes, as
− = mẍi − fi . (17)
dt ∂ ẋi ∂xi is sometimes done in the literature. From Eq. (19) we also
From here it follows that Newton’s Second Law (for a con- verify that L can be multiplied by an arbitrary constant. Thus
servative system) any Newtonian system accepts an infinity of equivalent La-
grangians.
∂V As mentioned above, in more advanced textbooks it is
mẍi = − = fi (18)
∂xi common to derive the Lagrange equations from a variational
is equivalent to the set of equations principle, known as Hamilton’s principle, applied to the ac-
tion of the system, defined as
d ∂L ∂L
− = 0, (19)
dt ∂ ẋi ∂xi Zt2
which hold irrespective of the functions Qi , and hence of the S= L[x(t), ẋ(t); t]dt. (25)
specific vector G(E) . The equivalence means that any one of t1
the equations (18) and (19) implies the other one, via (17).
The function L(x, ẋ) is the Lagrangian of the system, The variational principle becomes then the single one that
and Eqs. (19) are the Lagrange equations of motion, which, enters in place of the set of Newton’s equations of motion;
as just seen, are equivalent to Newton’s Second Law. Note it states that the deviation δS of the action for any infinites-
that L is a scalar, and as such it remains invariant under a imal arbitrary deviation from the real trajectory followed by
change of (spatial) coordinates. the particle, for initial and final times t1 and t2 fixed, is null.
Introducing Eq. (19) into (16) one obtains Otherwise stated, the action acquires an extremum (usually
a minimum) value along the real trajectory from t1 to t2 . A
(E) dQi ∂ X relatively immediate application of this demand to Eq. (25)
Gi = − Qj ẋj , (20)
dt ∂xi j leads to the Lagrange equations of motion [1-6]. Since the
variation of L0 given by equation (25) is the same as for L for
and hence, from Eq. (4), t1 and t2 fixed, the variational principle holds for all equiva-
dE X (E)
X dQi X dQj lent Lagrangians.
= ẋi Gi = ẋi − ẋj = 0. (21) The fact that the whole set of equations of motion fol-
dt i i
dt j
dt
lows from a single principle of minimum action is of course
This verifies Eq. (6) (for ξ = E), as expected. most appealing, and it is not surprising that it gives rise to
If the vector Q is selected as irrotational, the terms under extremely valuable tools for the development of theoretical
the summation sign in Eq. (11) vanish and G(E) becomes the physics. However, some caution is needed to understand its
(E) significance. Newton’s Second Law is strictly local, mean-
null vector along the trajectory, Gi = mẍi − fi = 0, which
is the trivial solution of Eq. (6). Since in this case one can ing that the particle responds instantaneously to the force ap-
write Qi = ∂K/∂xi with K(x) an arbitrary scalar function plied just on it, at its position and at precisely that instant.
of x, Eq. (13) becomes The principle of least action, by contrast, refers to what hap-
pens along the entire trajectory from t1 to t2 ; it says that the
(E)
X ∂K(x) dK
ϕK = L + ẋi = L + . (22) particle adjusts its trajectory from the initial to the final time
i
∂xi dt so that the accumulated action will result in a minimum. Of

Rev. Mex. Fı́s. E 17 (1) 47–54


50 L. DE LA PEÑA, A.M. CETTO AND A. VALDÉS-HERNÁNDEZ

course the principle applies only because it holds for each which confirms that the Lagrange equations look the same in
infinitesimal displacement, so the particle accumulates min- any system of coordinates {qi }. In more formal terms we say
imal actions one after the next along its trajectory. The par- that the Lagrange equations are covariant, or that they are
ticle does not know at time t1 what it will be performing at written in covariant form.
a future time t2 , in order to adjust itself to the demand of But then, what is the importance of covariance? Newton’s
a global minimum action: the global action is minimal be- law is a physical law, and as such it is totally independent of
cause each microscopic action is minimal. For an elementary which system of coordinates is used to express it. The law
derivation of the Lagrange equations from Hamilton’s princi- expresses a fact of nature, whereas our coordinate system is
ple see, e.g., [8]. something external to it, freely and arbitrarily selected. Ex-
pressing the physical law in such a way that it does not de-
pend on the coordinate system, would of course be most ap-
3. Covariance of the equations of motion propriate and satisfactory.
In the foregoing section it became clear that for the single- To guarantee the independence of the law from the se-
particle, conservative problem treated here, the Lagrange lected basis vectors, the matrix that transforms the vector
equations of motion are equivalent to Newton’s equations of components must be the inverse of the matrix that transforms
motion, so one might wonder what is the advantage of us- the basis vectors. We say that such vector components trans-
ing the former. One important difference is that the Lagrange form contravariantly with respect to the basis vectors. This
equations, unlike Newton’s, have the same form in all coor- feature, which is crucial to guarantee that the laws of nature
dinate systems, as we shall now see. This property ultimately are the same in any coordinate system, applied to the sec-
follows (in the present derivation) from the fact that the scalar ond of equations (28), implies that the k-th component of the
product is invariant under geometric transformations. Specif- velocities transforms as
ically, Eqs. (6) and (7) hold for all coordinate systems, so X ∂qk
q̇k = ẋi , (32)
that in terms of another set of coordinates qi0 = qi0 ({ql }) for ∂xi
i
instance, we have ϕ0(ξ) = ϕ(ξ) [q(q 0 )], and
whereas the k-th component of the vector G transforms in-
dE X (E)
X 0(E) versely, in terms of the matrix elements ∂xi /∂qk instead of
=0= q̇i Gi = q̇k0 Gk , (26)
dt i
∂qk /∂xi , as follows from Eq. (30). Below we come back to
k
this point.
with
0(E) d ∂ϕ0(E) ∂ϕ0(E) 3.1. Some geometry
Gk = 0 − . (27)
dt ∂ q̇k ∂qk0
In preparation for the following section we develop here
Since the Cartesian coordinates {xi } and the generalized some mathematical tools that will be useful to distinguish be-
coordinates {qi } are related according to Eq. (1), we have tween the geometric properties and the dynamical properties
X ∂xi X ∂qi of a system, when arbitrary coordinates {qi } are employed.
dxi = dq , dq i = dxk , (28) Consider first the arc element ds. We may express this el-
∂qk k ∂xk
k k ement in both the Cartesian coordinate system and the more
and therefore, general one {qi }; we then have
X X
X (E)
X ³ X ∂xi ´ (E) (ds)2 = ηij dxi dxj = gkl dqk dql , (33)
ẋi Gi (x) = q̇k Gi [x(q)] i,j k,l
i i
∂qk
k
X ³ X ∂x ´ where ηij and gkl are the components of the metric tensor
i (E)
= q̇k Gi [x(q)] . (29) g in terms of the Cartesian and the generalized coordinates,
i
∂qk respectively. Using the first equation in (28) we are led to
k

Thus, the invariance of the scalar product in (26) gives the X ∂xi ∂xj X
(E) (ds)2 = ηij dqk dql = gkl dqk dql ; (34)
law of transformation of Gi (x), ∂qk ∂ql
i,j,k,l k,l
X ∂xi (E) P
(E)
Gk (q) = G (x(q)). (30) thus gkl = i,j ηij (∂xi /∂qk )(∂xj /∂ql ). For a Euclidean
i
∂qk i (flat) space, ηij = δij , and therefore
X ∂xi ∂xi
From Eqs. (13), (16) and (19) we note that the choice Qi = 0 gkl = glk = . (35)
(E) 0(E) ∂qk ∂ql
implies ϕ(E) = L and Gi = 0, hence also Gk = 0 ac- i
cording to Eq. (30). Therefore, from Eq. (27) we arrive at
Thus the components gkl are constant for linear transforma-
d ∂L(q, q̇) ∂L(q, q̇) tions; for curvilinear coordinates they are nontrivial functions
− = 0, (31) of the generalized variables.
dt ∂ q̇i ∂qi

Rev. Mex. Fı́s. E 17 (1) 47–54


POWER AND BEAUTY OF THE LAGRANGE EQUATIONS 51

The components of the inverse of the metric tensor An example of a contravariant vector is given in
g−1 =ḡ are such thati Eq. (32) (another one is the law of transformation for fi in
X Eq. (48) below), which transforms with the matrix elements
gil ḡln = δin , (36) ∂qk /∂xi . An example of a covariant vector, transforming
l with ∂xi /∂qk , appears in Eq. (30). In an orthogonal system
and they are given by of coordinates these vectors coincide (except for a possible
interchange of coordinates), so there is no need to distinguish
X ∂qk ∂ql
ḡkl = ḡlk = . (37) between them. However, in general both kinds of indices (of
i
∂xi ∂xi transformations) are necessary; this happens in particular in
the general theory of relativity where space-time is curved
A set of functions that will be of relevance, as we shall due to the presence of matter and energy, and is also the case
see below, are the Christoffel symbols (a kind of affine con- in Eq. (47) below. The scalar product corresponds then to
nections) Γljk , defined as [2, 9, 10] a contraction, i.e., the sum over two equal indices, one from
X ∂ql ∂ 2 xi a factor that transforms covariantly (written as a subindex),
Γljk = Γlkj ≡ . (38) and the other from a factor thatPtransforms contravariantly
i
∂xi ∂qj ∂qk (written as a superindex), as in i q i Gi = q · G, resulting
in an invariant (a scalar). Notice that, in line with classical
A comparison with Eq. (35) suggests that the Γljk will be mechanics terminology, we have avoided the distinction be-
useful to express the derivatives of the metric tensor and tween upper and lower indices when writing covariant and
viceversa. Indeed, by taking the derivative of gkl and using contravariant components. For a more detailed discussion
Eq. (38), one arrives at the expression see, e.g., Refs. [1, 2, 4].
∂gkl X¡ ¢
= gil Γiks + gik Γils . (39)
∂qs i 3.2. Covariant form of Newton’s second law
To invert this relation and express the affine connection in In this section we go back to Eq. (18) and apply a geometric
terms of the metric tensor, one combines the above equation transformation, to express it in terms of generalized coordi-
and its two permutations, thereby obtaining nates. The process will lead us to the covariant form of New-
∂gkl ∂gsl ∂gks X ton’s Second Law, clearly revealing that the inertial forces
+ − =2 gil Γiks . (40) (forces proportional to the mass of the particle) have a geo-
∂qs ∂qk ∂ql i metric origin associated with the curvature of the coordinates.
Multiplying the above equation by ḡln and sum over l leads to We first express ẍi in terms of the generalized coordinates
a formula for the Christoffel symbols in terms of the deriva- {qi } and their time derivatives as follows,
tives of the metric tensor
µ ¶ d X ∂xi X ∂xi
n 1X ∂gkl ∂gsl ∂gks ẍi (q) = q̇k = q̈k
Γks = ḡln + − . (41) dt ∂qk ∂qk
2 ∂qs ∂qk ∂ql k k
l
X 2
∂ xi
Since the affine connections depend on the coordinates + q̇j q̇k . (42)
∂qj ∂qk
only and not on the velocities (see Eq. (38)), they have a ge- j,k

ometric meaning and are directly related to the metric of the


space in question. Since we are looking for an equation for q̈k , it is convenient
With this we have the geometric tools required to de- to multiply the above expression by ∂ql /∂xi , sum over i, and
scribe Newtonian dynamics in an arbitrary coordinate sys- use the equality
tem, which is the aim of the following section; but before X ∂xi ∂ql ∂ql
closing this one it is convenient to add a couple of comments = = δkl , (43)
regarding the notation. i
∂qk ∂xi ∂qk
In the affine connections defined by Eq. (38) we intro-
duced lower and upper indices: these represent different laws to obtain
of transformation, so their meaning is important and goes be- X X ∂ql
yond a mere convenience in the writing. The upper indices q̈l + q̇j q̇k Γljk = ẍi (q) . (44)
∂xi
describe the components of contravariant tensors, whereas j,k i
the lower indices refer to the covariant ones. The contravari-
ant components of a vector are obtained by projecting the The dynamics enters by using Eq. (18) in the form
vector onto the coordinate axes; in their turn, the covariant
components are obtained by projecting onto the lines normal 1 1 X ∂V (q) ∂qj
ẍi = fi = − , (45)
to the coordinate hyperplanes [10, 12]. m m j ∂qj ∂xi

Rev. Mex. Fı́s. E 17 (1) 47–54


52 L. DE LA PEÑA, A.M. CETTO AND A. VALDÉS-HERNÁNDEZ

so that Assume we are interested in the study of a point particle


X ∂ql 1 X ∂V (q) ∂qj ∂ql moving on a plane. We may describe its motion using Carte-
ẍi =− sian coordinates x1 = x, x2 = y, polar coordinates q1 = r,
∂xi m i,j ∂qj ∂xi ∂xi
i q2 = θ, or some other set of coordinates; the choice may de-
1 X ∂V (q) pend on the symmetry properties of the forces acting on the
=− ḡjl . (46) particle, or on constraints imposed on the motion.
m j ∂qj
Let us first consider the simple case of a central force, i.e.,
Substitution into Eq. (44) leads to the desired dynamical a force that acts along the line joining the particle and the
equation, center of force (taken as the origin), and whose magnitude
does not depend on the angle. In such case, because of the
X ∂V (q) X circular symmetry, the polar coordinates are the most appro-
mq̈l = − ḡjl − m q̇j q̇k Γljk . (47)
j
∂qj priate ones. The transformation rules between the Cartesian
j,k
and polar coordinates areiv
This is the explicit covariant form of Newton’s Second Law
(for the single-particle, conservative and unconstrained prob- x(r, θ) = r cos θ, y(r, θ) = r sin θ; (50)
lem), written in a general system of coordinates (though with- p
out paying due attention to the position of the indices). The r(x, y) = x2 + y 2 , θ(x, y) = tan−1 (y/x). (51)
geometry of the specific coordinates being encoded in the (in-
Upon comparison of the arc element
verse of the) metric tensor g and the Christoffel symbols (see,
e.g., Christoffel symbols in Mathematica)ii . (ds)2 = (dx)2 + (dy)2 = (dr)2 + r2 (dθ)2 (52)
Between Eq. (18) and Eq. (47) there is a world of dif-
ference. Yet when the coordinates {qi } are the appropriate with Eq. (34) one gets
ones for the visualization of the dynamical problem of inter-
est, Eq. (47) gives us the most transparent description of the g11 = 1, g12 = g21 = 0, g22 = r2 . (53)
behavior of the system, as we shall see below.
The components of the inverse metric tensor are obtained by
3.3. Curvilinear coordinates and geometric forces inverting the matrix
µ ¶
Equation (47) shows that there are in general two sources for 1 0
g= , (54)
the acceleration q̈l , namely the generalized force f˜l generated 0 r2
by the external force f according to
which gives
X ∂ql X ∂V
f˜l = fi =− ḡjl , (48)
∂xi ∂qj ḡ11 = 1, ḡ12 = ḡ21 = 0, ḡ22 = r−2 . (55)
i j

and the additional inertial force F̃l , always bilinear in the ve- The metric tensor and its inverse can also be obtained of
locities q̇i and given by course by resorting to Eqs. (35) and (50), and to Eqs. (37)
X and (51), respectively.
F̃l = −m q̇j q̇k Γljk . (49) Using now Eq. (48) we obtain for the components of the
j,k generalized force
Now, if the transformation xi → qi is linear, all coefficients ∂V ∂V
f˜l = −ḡl1 − ḡl2 , (56)
∂ 2 xi /∂qj ∂qk in Eq. (38) vanish, resulting in a null inertial ∂q1 ∂q2
force. Thus, inertial forces appear only when the transforma-
tion xi → qi is nonlinear. A reference system that is nonlin- whence
early related to an inertial one constitutes thus a noninertial ∂V ∂V
reference system, in which inertial forces appear as a result f˜r ≡ f˜1 = −ḡ11 =− , (57)
∂q1 ∂r
of the curvature of the space coordinatesiii .
∂V 1 ∂V
f˜θ ≡ f˜2 = −ḡ22 =− 2 . (58)
3.4. Two elementary problems as seen from the general ∂q2 r ∂θ
theory The affine connections follow from Eq. (41), the only ele-
The powerful mathematical tools developed above are essen- ments different from zero being
tial to deal with complex problems, say within general rela- 1 ∂g22
tivity, so the present exposition may be taken as groundwork Γ122 = − ḡ11 = −r, (59)
2 ∂q1
for those higher matters. However, it seems appropriate to
1 ∂g22 1
analyze here a couple of elementary problems, to gain famil- Γ212 = Γ221 = ḡ22 = . (60)
iarity with such tools. 2 ∂q1 r

Rev. Mex. Fı́s. E 17 (1) 47–54


POWER AND BEAUTY OF THE LAGRANGE EQUATIONS 53

From Eq. (49) applied to this case, The corresponding Lagrange equation for the uncon-
¡ ¢ strained generalized coordinate reads
F̃l = −m 2Γl12 q̇1 q̇2 + Γl22 q̇22 , (61)
d ∂T ∂T ∂V
− =− , (71)
we obtain thus dt ∂ q̇1 ∂q1 ∂q1

F̃r = −mΓ122 q̇22 = mrθ̇2 , (62) where we wrote L = T − V , with T the kinetic energy (see
Eqs. (74) and (77) below). Equation (71) can be solved with-
F̃θ = −2mΓ212 q̇1 q̇2 = −2mr−1 ṙθ̇. (63) out knowing the forces of constraint. This is a great advan-
tage, for the forces of constraint depend on the motion of the
Substitution of these results in Eq. (47) gives the equations
particle, and therefore cannot be determined in general un-
of motion in polar coordinates:
til the motion is known. In some instances it is important to
∂V know such forces; these can then be calculated from the La-
mr̈ = − + mrθ̇2 , (64) grange equations for the constrained coordinate (in our case,
∂r
the coordinate q2 ):
1 ∂V 2m
mθ̈ = − 2 − ṙθ̇. (65)
r ∂θ r d ∂T ∂T ∂V
− =− . (72)
Notice that these equations contain each an inertial force dt ∂ q̇2 ∂q2 ∂q2
term. Let us recast them both in a more familiar form. Firstly,
Indeed, by substituting here the solution for q1 (t) obtained
multiplication of (65) by r2 gives
from solving Eq. (71), we find the constraining force
d dL ∂V (−∂V /∂q2 ). This shows that the Lagrange formalism is par-
mr2 θ̈ + 2mrṙθ̇ = (mr2 θ̇) = =− , (66) ticularly suited to deal with systems subject to holonomic
dt dt ∂θ
constraintsv .
where mr2 θ̇ has been identified as the angular momentum
L. This is the dynamical equation for the angular momen-
tum, driven just by the external torque. In particular, it shows 4. The relation between momenta and veloci-
that for radial forces (V = V (r)) the angular momentum is ties. The Hamiltonian
conserved (it is an integral of motion),
In concluding, it seems convenient to briefly recall the con-
L = mr2 θ̇ = constant. (67) nection between the two fundamental functions that serve to
define, each one in their own capacity, the dynamics of a
On the other hand, in terms of L (whether conserved or not), mechanical system, namely the Lagrangian and the Hamil-
Eq. (64) takes the form tonian. For this purpose we first introduce the (generalized)
momentum pi , defined via the Lagrangian as
∂V L2
mr̈ = − + , (68)
∂r mr3 ∂L(q, q̇)
pi = . (73)
2 2
which contains the centrifugal force L /mr = mrθ̇ , an 2 ∂ q̇i
inertial force that propels the particle radially away from the The variables qi , pi are thus said to be canonically conjugate.
origin. In order to calculate pi , we transform L(x, ẋ) to L(q, q̇)
Let us now take as a second example the case of a parti- by resorting to Eqs. (14), (28) and (35),
cle that, in addition to being subject to a conservative force
f (x, y) = −∇V (x, y), is constrained to move along a pre- 1 X ∂xi ∂xi
scribed trajectory on the plane; we may think of a frictionless L(q, q̇) = m q̇k q̇j − V (x(q))
2 ∂qk ∂qj
i,j,k
rail that enforces this dynamics, by exerting at all times a
force on the particle perpendicular to its trajectory. The ge- 1 X
= m gkj q̇k q̇j − V (x(q)). (74)
ometrical shape of the rail can be expressed as a functional 2
j,k
relation that must hold between the two position coordinates
(with a constant) Equation (73) gives therefore a linear relation between the
g(x, y) = a, (69) momenta and the velocities
which reduces the number of degrees of freedom by one. X
pi = m gik q̇k , (75)
We may conveniently choose the generalized coordinates as
k
q1 = q1 (x, y), and q2 = a (q2 will be fixed to a constant
value through the constriction (69)), which can be inverted which can be inverted to obtain
using (69) to solve for the Cartesian coordinates,
1 X
q̇i = ḡik pk . (76)
x = x(q1 , q2 ), y = y(q1 , q2 ). (70) m
k

Rev. Mex. Fı́s. E 17 (1) 47–54


54 L. DE LA PEÑA, A.M. CETTO AND A. VALDÉS-HERNÁNDEZ

Only for rectilinear coordinates (when all the gik are con- It is a simple, illustrative exercise to use this result for the
stant) this relation is constant along the trajectory. Further, derivation of the set of equations of motion [2-4]
Eq. (75) allows us to write the kinetic energy as
∂H ∂H
1 X 1 X = q̇i , = −ṗi . (81)
T = ḡij pi pj = m glk q̇l q̇k . (77) ∂pi ∂qi
2m i,j 2
k,l

Note that the kinetic energy is always a bilinear function of These are the notorious Hamilton equations of motion that
either the velocities or the momenta. completely describe the evolution of any system defined by
Let us now combine Eqs. (9) and (14), to get the Hamiltonian H(q, p, t). We see that there are 2N differ-
ential Hamilton equations of first order for a system with N
X ∂L X
E(q, q̇) = q̇i −L= q̇i pi − LH(q, p). (78) degrees of freedom, while according to Eq. (19) the dynam-
i
∂ q̇i i ics of the same system is described by N Lagrange differen-
tial equations of second order. For this reason, in some cases
The right-hand side of this equation corresponds to the def-
the description of the dynamics of a system in terms of the
inition of the Hamiltonian H(q, p), which, as is clear from
Hamilton equations of motion is simpler than in Lagrangian
the first equality, for a closed, conservative system coincides
terms.
with the mechanical energyvi .
By using (76) we recover a most important relationship
between the Lagrangian and the Hamiltonian,
Acknowledgments
1 X
H= ḡik pk pi − L, (79)
m The authors gratefully acknowledge financial support from
i,k

which gives DGAPA-UNAM through project PAPIIT IA101918.

1 X
H(q, p) = ḡij pi pj + V (q). (80)
2m i,j

i. The transition from the x-system to the q-system of coordi- 1. H. Goldstein, Ch. Poole and J. Saiko, Classical Mechanics 3rd
nates can be viewed as a mere change of variables, where the ed. (Addison-Wesley, MA, 1980).
coefficients Jkl = (∂xi /∂qk ) are the elements of the Jaco-
2. H. C. Corben and P. Stehle, Classical Mechanics (John Wiley,
bian of the transformation (for the transition from xk to qk )
NY, 1950), Chap. V.
and, similarly, the elements of the inverse matrix J −1 = J¯ are
J¯lk = (∂qk /∂xl ) (for the inverse transition from qk to xk ). 3. N. Rasband, Dynamics (John Wiley, NY, 1983), Chap. III.
ii. Though this form is not common in classical mechanics, it can 4. J. V. José and E. J. Saletan, Classical Dynamics. A contempo-
be found from time to time [2, 11]. rary approach (Cambridge University Press, Cambridge, 1998).
iii. The observation that forces can be associated with curvatures of
5. J. R. Taylor, Classical Mechanics (University Science Books,
the coordinate system —so fundamental for the general theory
CA, 2005).
of relativity—and that a covariant description is needed, is older
than one would think; it was mentioned already by the brilliant 6. K. R. Symon, Mechanics 2nd. ed. (Addison-Wesley, MA,
German mathematician Bernhard Riemann (1826-1866), fol- 1960).
lowed by another brilliant mathematician, the British William 7. R. P. Feynman, R. B. Leighton and M. Sands, The Feynman
K. Clifford (1845-1879), as discussed in Ref. [13], Chapter 5. Lectures on Physics, Vol. II (Addison-Wesley, Reading, MA,
iv. In Eq. (51) it should be understood that tan−1 (y/x) is the two- 1964).
argument inverse tangent, which takes into account the signs of
both x and y. 8. J. Hanc, E. F. Taylor and S. Tuleja, Am. J. Phys. 72 (2004) 510.
v. Here we have dealt with holonomic constraints, which depend 9. See, e.g., R. Price, Am. J. Phys. 50 (1982) 300.
on the position coordinates. In the more general case, con- 10. S. Weinberg, Gravitation and Cosmology. Principles and Ap-
straints may include restrictions on the velocity. If these con- plications of the General Theory of Relativity (John Wiley, NY,
straints can be integrated so as to lead to an expression of the 1972).
form (69), they are still holonomic. There are cases, however,
in which the equations of constraint cannot be integrated (take 11. T. M. Kalotas and B. G. Wybourne, J. Phys. A 15 (1982) 2077.
for instance, a circular wire rolling on a plane without slipping); 12. See, e.g., R. Courant, Differential and Integral Calculus,
we then speak of nonholonomic constraints. For a more detailed Vol.II,(John Wiley, NY, 1962), Chapt. III.
discussion and additional examples see Ref. [6], Chapter 9.
13. M. Jammer, Concepts of Space: The History of Theories of
vi. The function E(q, q̇), whose value is just that of the Hamil-
Space in Physics (Dover, N.Y., 1954).
tonian H(q, p) but expressed in terms of velocities instead of
momenta, is called the energy function; see Ref. [1], Sec. 2.7.
Rev. Mex. Fı́s. E 17 (1) 47–54

You might also like