diff --git a/chapter-1.html b/chapter-1.html index 0236701..02f0402 100644 --- a/chapter-1.html +++ b/chapter-1.html @@ -278,179 +278,185 @@

- + 1.2 Hamilton’s variational principle

- + 1.3 First steps with conserved quantities

- + 1.4 Euler-Lagrange equations on smooth manifolds

- + 2 Conservation Laws

- + 2.1 Noether theorem

- + 2.2 The spherical pendulum

- + 2.3 Intermezzo: small oscillations

- + 2.4 Motion in a central potential

- + 2.5 D’Alembert principle and systems with constraints

- + 3 Hamiltonian mechanics

- + 3.1 The Legendre Transform in the euclidean plane

- + 3.2 Intermezzo: cotangent bundle and differential forms

- + 3.3 The Legendre Transform

- + 3.4 Poisson brackets and first integrals of motion

- + 3.5 Variational principles of hamiltonian mechanics

- + 3.6 Canonical transformations

- + 3.7 The symplectic structure on the cotangent bundle

- + 4 Integrable systems

- + 4.1 Lagrangian submanifolds

- + 4.2 Canonical transformations revisited

- + 4.3 Hamilton-Jacobi equations

- + 4.4 The Liouville-Arnold theorem

- + 5 Hamiltonian perturbation theory

- + 5.1 Small oscillations revisited

- + 5.2 Birkhoff normal forms

- + 5.3 A brief look at KAM theory

- + 5.4 Nekhoroshev theorem

- + 6 Conclusion

- + Bibliography

+

+ +Index +

+ + @@ -659,9 +665,10 @@

1.1 Newtonian mecha

-In the elementary formulation we will focus on, our primary interest is in describing the equations of motion for an idealized point particle. This is a point–like object obtained by ignoring the dimensions of the physical object. Note that in -many cases this is a reasonable first approximation. For example, when describing planetary motion around the sun, we can consider the planet and the sun as two point particles without significantly affecting the qualitative properties of the system. -After all the sun is about 150 million kilometers away from the Earth, compared to which their sizes are negligible. The diameter of the sun is just 0.9% of this distance and the diameter of the Earth a mere 0,008%. +In the elementary formulation we will focus on, our primary interest is in describing the equations of motion for an idealized point particle. This is a point–like object obtained by ignoring the dimensions of the +physical object. Note that in many cases this is a reasonable first approximation. For example, when describing planetary motion around the sun, we can consider the planet and the sun as two point particles without significantly affecting the +qualitative properties of the system. After all the sun is about 150 million kilometers away from the Earth, compared to which their sizes are negligible. The diameter of the sun is just 0.9% of this distance and the diameter of the Earth a mere +0,008%. See Figure 1.1.

@@ -684,6 +691,8 @@

1.1 Newtonian mecha

+ +
@@ -693,7 +702,7 @@

1.1 Newtonian mecha

-A point particle is first of all a point, a dimensionless mathematical object in space: its position in space is described by the position vector \(\vb *{x} = (x, y ,z)\). Keep in mind that +A point particle is first of all a point, a dimensionless mathematical object in space: its position in space is described by the position vector \(\vb *{x} = (x, y ,z)\). Keep in mind that

@@ -718,7 +727,7 @@

1.1 Newtonian mecha

-The velocity of the point particle is given by the rate of change of the position vector, i.e., its derivative with respect to time1 +The velocity of the point particle is given by the rate of change of the position vector, i.e., its derivative with respect to time1

@@ -739,7 +748,7 @@

1.1 Newtonian mecha

-We call acceleration, the rate of change of the velocity, i.e., the second derivative +We call acceleration, the rate of change of the velocity, i.e., the second derivative

@@ -762,13 +771,13 @@

1.1 Newtonian mecha

For us, the main distinction between a mathematical point and a point particle, is that the latter usually carries a mass \(m\). This is a scalar quantity that, as we will see, measures its resistance to changes in its state of motion, also -known as inertia. +known as inertia.

-1 Notation: the symbol \(a := b\) means that \(a\) is defined by the expression \(b\), similarly \(b =: a\) is the same statement but read from right to left. +1 Notation: the symbol \(a := b\) means that \(a\) is defined by the expression \(b\). Similarly \(b =: a\) is the same statement but read from right to left.

@@ -781,14 +790,27 @@

1.1 Newtonian mecha > +
+ +

+Newton’s second law of motion +

+
+ +
+ +

-The mechanics of the particle is encoded by Newton’s second law of motion. That is, there exist frames of reference (i.e systems of coordinates) in which the motion of the particle is described by a differential equation involving -the forces \(\vb *{F}\) acting on the point particle, its mass \(m\) and its acceleration as follows2 +There exist frames of reference, that is, systems of coordinates, in which the motion of the particle is described by a differential equation involving the forces \(\vb *{F}\) acting on the point particle, its mass +\(m\) and its acceleration as follows

@@ -808,43 +830,18 @@

1.1 Newtonian mecha

-

-
- - -

-2 This is not completely true, Newton was already talking about momentum, which makes a deep difference but would set us off-course right now. -

- -

-Let’s leave it here for now, we will come back to it later on. +Note that the law as described above is not completely true: Newton was already talking about momentum, which makes a deep difference, but would set us off-course right now.

-In general we will consider systems of \(N\) point particles. These will be described by a set of \(N\) position vectors \(\vb *{x}_k = (x_k, y_k ,z_k)\) with masses \(m_k\), \(k = 1, \ldots , N\). For convenience we will denote3 -\(\vb *{x} = (\vb *{x}_1, \ldots , \vb *{x}_N)\in \mathbb {R}^{3N}\), and \(\vb *{m} = (m_1, \ldots , m_N)\in \mathbb {R}^N\). We call \(\vb *{x}(t)\) the configuration of the system at time \(t\) in the -configuration space \(\mathbb {R}^{3N}\). +Before moving on, let’s anticipate three examples that will come back over and over in the course of this book. Here we will just sketch them, but we will come back to them in more detail later.

-
- - - -

-3 Here we are identifying \(\mathbb {R}^{3\times N}\), the space of \(3\times N\) matrices, and \(\mathbb {R}^{3N}\), the space of \(3N\) vectors. Namely, \(\vb *{x} = (\vb *{x}_1, \ldots , \vb *{x}_N) = (x_1, y_1, -z_1, x_2, y_2, z_2, \ldots , x_N, y_N, z_N)\) is used interchangeably as the vector of positions of the points in three-space or as the vector including all the positions of the \(N\) bodies together. -

- - -
- - -
-
- - - -

-4 Here there is an hidden statement: \(\mathbb {S}^1\times \mathbb {S}^1 \simeq \mathbb {T}^2\). This should be read as: the 2-torus, the surface of a doughnut, is diffemorphic to the product of two circles. Here, -diffeomorphic is a concept from the theory of smooth manifolds, and stands for the existence a smooth bijection between the two constructions. Note that \(\simeq \) usually denotes some kind of equivalence, but which specific kind is -context-dependent and not often well explicited. + +Example 1.1 (Horizontal spring and pendulum). Consider an ideal spring, that is, an idealized system consisting of a point particle of mass \(m\) attached to a spring with stiffness \(k\), sliding on a frictionless surface. +Assume that the motion is one-dimensional, along the axis of the spring, and let \(x\) denote the displacement of the system from its equilibrium position, i.e., the position in which the spring is completely at rest: not compressed nor extended. It is +convenient to assume that the sliding happens on a flat surface. This allows us to exclude gravitational forces from the picture, making this example as simple as it could be.

- - -
- -
-

-(image)(image)

+

-Figure 1.2: Left: a system of two decoupled pendulums in \(\mathbb {R}^3\). Right: corresponding generalized coordinates. +Figure 1.2: Left: horizontal spring. Right: planar pendulum.

- +

-We say that a system of \(N\) particles has \(n\) degrees of freedom if we need \(n\) independent parameters to uniquely specify the system configuration5. We call generalized -coordinates any set of \(n\) parameters \(q = (q^1, \ldots , q^n)\) that uniquely determine the configuration of a system with \(n\) degrees of freedom, and generalized velocities their time derivatives \(\dot q = (\dot q^1, -\ldots , \dot q^n)\). The state of the system is then characterized by the set of (generalized) coordinates and velocities \((q, \dot q) = \left (q^1, \ldots , q^n,\dot q^1, \ldots , \dot q^n\right )\). -

- -

-If you recall differential geometry, you may (correctly) guess that the generalized coordinates will be points on some differentiable manifold \(q\in M\), their evolution will be described by a curve \(q: \mathbb {R} \to M\) parametrized by time and -the state of the system will be a point in the tangent bundle \((q, \dot q)\in TM\), i.e., \(\dot q \in T_q M\) (see also section 1.4). -

- -

-We have now all the elements to translate the Newtonian principle of determinacy in mathematical terms. In 1814, Laplace [Lap51] wrote -

-
- -

-We may regard the present state of the universe as the effect of its past and the cause of its future. An intellect which at a certain moment would know all forces that set nature in motion, and all positions of all items of which nature is composed, if -this intellect were also vast enough to submit these data to analysis, it would embrace in a single formula the movements of the greatest bodies of the universe and those of the tiniest atom; for such an intellect nothing would be uncertain and the -future just like the past would be present before its eyes. -

-
- -

-In other words, this principle states that the initial state \(\left (q(t_0), \dot q(t_0)\right )\) uniquely determines its evolution \(\left (q(t),\dot q(t)\right )\) for \(t > t_0\). The Picard-Lindelöf theorem6 implies that the Newtonian principle of determinacy is locally satisfied by the equations of motion of the mechanical system, i.e., second order differential equations derived from Newton’s law. -

-
- - -

-5 As Example 1.1 shows, the \(n\) degrees of freedom do not have to be the cartesian coordinates of the point particles. -

- -

-6 Also known as “Existence and uniqueness of solutions of initial value problems”, [Kna18, Theorem 3.17]. -

- - -
- -

1.1.1 Motion in one degree of freedom
- - - -

-Before taking a little detour into Lagrangian mechanics, let’s anticipate some of the upcoming concepts in a few simple cases. -

-
- -
-

1.2 Hamilton’s variational principle

- +
1.1.1 From particles to systems of particles
+ - +

+It is a bit limiting to consider only systems with one particle and, indeed, many interesting problems in classical mechanics involve systems on interacting particles. +

-In order to leave more space to discuss Hamiltonian systems and their geometry, we will be brief7 in our account of Lagrangian mechanics and calculus of variations. For a more detailed account refer to [Arn89, Part II]. This, however, should not confuse you: Lagrangian mechanics plays a role as large as Hamiltonian mechanics in the development of classical mechanics. +In general a system of \(N\) point particles will be described by a set of \(N\) position vectors \(\vb *{x}_k = (x_k, y_k ,z_k)\) with masses \(m_k\), \(k = 1, \ldots , N\). For convenience we will denote \(\vb *{x} += (\vb *{x}_1, \ldots , \vb *{x}_N)\in \mathbb {R}^{3N}\), and \(\vb *{m} = (m_1, \ldots , m_N)\in \mathbb {R}^N\). We call \(\vb *{x}(t)\) the configuration of the system at time \(t\) in the +configuration space \(\mathbb {R}^{3N}\).

-In fact, one should not be surprised if there are many sources claiming that the most general formulation of the equations of motion in classical mechanics comes from the principle of least action or Hamilton’s variational principle. +Note that we are identifying \(\mathbb {R}^{3\times N}\), the space of \(3\times N\) matrices, and \(\mathbb {R}^{3N}\), the space of \(3N\) vectors. Namely, \(\vb *{x} = (\vb *{x}_1, \ldots , \vb *{x}_N) = (x_1, y_1, z_1, x_2, +y_2, z_2, \ldots , x_N, y_N, z_N)\) is used interchangeably as the vector of positions of the points in three-space or as the vector including all the positions of the \(N\) bodies together.

+
+ + +
+ +

+As exemplified by the example above, we want to set up a formalism that allows us to describe the configuration of a system in a way that is independent of the number of particles and instead captures the minimal number of parameters that are +really necessary to describe it. +

+ +

+We say that a system of \(N\) particles has \(n\) degrees of freedom if we need \(n\) independent parameters to uniquely specify the system configuration. We call generalized coordinates any set of \(n\) parameters \(q = (q^1, \ldots , q^n)\) that uniquely determine the configuration of a system with \(n\) degrees of freedom, and generalized velocities their time +derivatives \(\dot q = (\dot q^1, \ldots , \dot q^n)\). The state of the system is then characterized by the set of (generalized) coordinates and velocities \((q, \dot q) = \left (q^1, \ldots , +q^n,\dot q^1, \ldots , \dot q^n\right )\). +

+ +

+As Example 1.3 shows, the \(n\) degrees of freedom do not have to be the cartesian coordinates of the point particles. If you recall differential geometry, you may (correctly) guess that the +generalized coordinates will be points on some differentiable manifold \(q\in M\), their evolution will be described by a curve \(q: \mathbb {R} \to M\) parametrized by time and the state of the system will be a point in the tangent bundle \((q, +\dot q)\in TM\), i.e., \(\dot q \in T_q M\). If you don’t have any previous experience with these concepts, don’t dispare! We will take our time to introduce them as they appear throughout the course. See also section 1.4). +

+ +

+We have now all the elements to translate the Newton’s principle of determinacy in mathematical terms. In 1814, Laplace [Lap51] wrote +

+
+ +

+We may regard the present state of the universe as the effect of its past and the cause of its future. An intellect which at a certain moment would know all forces that set nature in motion, and all positions of all items of which nature is composed, if +this intellect were also vast enough to submit these data to analysis, it would embrace in a single formula the movements of the greatest bodies of the universe and those of the tiniest atom; for such an intellect nothing would be uncertain and the +future just like the past would be present before its eyes. +

+
+ +

+In other words, this principle states that the initial state \(\left (q(t_0), \dot q(t_0)\right )\) uniquely determines its evolution \(\left (q(t),\dot q(t)\right )\) for \(t > t_0\). We had already observed this fact in +Example 1.1 of the pendulum, where the initial position and velocity uniquely determined the future trajectory. The Picard-Lindelöf theorem, also known as “Existence and uniqueness of solutions +of initial value problems” [Kna18, Theorem 3.17], implies that Newton’s principle of determinacy is locally satisfied by the equations of motion of the mechanical system, i.e., second order differential equations derived from Newton’s law. +

+ +

1.2 Hamilton’s variational principle

+ + + + + +

+In order to leave more space to discuss Hamiltonian systems and their geometry, we will be brief in our account of Lagrangian mechanics and calculus of variations, although not as brief as this twitter thread. This, however, should not confuse you: Lagrangian mechanics plays a role as +large as Hamiltonian mechanics in the development of classical mechanics! In fact, many sources claim that the most general formulation of the equations of motion in classical mechanics comes from the principle of least action or +Hamilton’s variational principle. According to this principle, stated in 1834, the equations of motion of a mechanical system are characterized by a function +

+ + + +

+ +\begin{equation} +L \equiv L(q, \dot q, t) : \mathbb {R}^n \times \mathbb {R}^n \times \mathbb {R} \to \mathbb {R}, +\end{equation} + +

+ +

+called the lagrangian (function) of the system. +

-Given a curve \(\gamma :[t_1, t_2] \to \mathbb {R}^n\), we define the action functional +Given a curve \(\gamma :[t_1, t_2] \to \mathbb {R}^n\), we define the action functional as

- + @@ -1445,15 +1464,33 @@

1.2 Hamilton’s v > +
+ +

+Principle of least action +

+
+ +
+ + + +

-Assume that the configuration of our system is \(q_1 = (q_1^1, \ldots , q_1^n)\) at an initial time \(t=t_1\) and \(q_2 = (q_2^1, \ldots , q_2^n)\) at a final time \(t = t_2\). The principle of least action, or Hamilton’s -principle, states that the evolution of our system in the time interval \([t_1, t_2]\) corresponds to the curve \(q(t)\) which is the critical point of the action functional \(S[q]\) on the space of curves \(q(t)\) with \(q(t_1) = q_1\) and -\(q(t_2) = q_2\). +Consider a system with \(n\) degrees of freedom. Assume that the system is in the configuration \(q_1 = (q_1^1, \ldots , q_1^n)\) at an initial time \(t=t_1\) and \(q_2 = (q_2^1, \ldots , q_2^n)\) at a final time \(t = t_2\). +

+ +

+The principle of least action, or Hamilton’s principle, states that the evolution of the system in the time interval \([t_1, t_2]\) corresponds to the curve \(q(t)\) which is the critical point of the action functional +\(S[q]\) on the space of curves \(\gamma :[t_1, t_2] \to \mathbb {R}^n\) with fixed endpoints \(\gamma (t_1) = q_1\) and \(\gamma (t_2) = q_2\).

@@ -1464,8 +1501,12 @@

1.2 Hamilton’s v

-Let \(X\) and \(Y\) be Banach spaces8 and \(G\subset X\) an open subset of \(X\). A function \(f: G \to Y\) is called Fréchet differentiable at \(x\in G\), if there exists a bounded linear operator \(A: X \to Y\), such -that9 +To state this properly we need the concept of Banach space, that is a complete normed vector space. If \(X\) denotes a Banach space, then we will denote the corresponding norm by \(\|\cdot \|_X\). +

+ +

+Let \(X\) and \(Y\) be Banach spaces and \(G\subset X\) an open subset of \(X\). A function \(f: G \to Y\) is called Fréchet differentiable at \(x\in G\), if there exists a bounded linear operator \(A: X \to +Y\), such that

@@ -1473,7 +1514,7 @@

1.2 Hamilton’s v @@ -1486,31 +1527,16 @@

1.2 Hamilton’s v

-for any \(h\) in a sufficiently small neighborhood of \(0\in X\). Alternatively, you can say asymptotically as \(\|h\|_X\to 0\). -

- -

-As in the finite-dimensional case, \(A\) is uniquely determined and is called the Fréchet derivative of \(f\) at \(x\) and is denoted \(D f(x)\). -

-
- - - -

-8 Banach space: complete normed vector space. +asymptotically as \(\|h\|_X\to 0\), that is, for any \(h\) in a sufficiently small neighborhood of \(0\in X\). Here, we write \(g = o(\|h\|_X)\) if \(g : X \to Y\) satisfies

-

-9 Here, \(g = o(\|h\|_X)\) if -

- - + @@ -1523,14 +1549,8 @@

1.2 Hamilton’s v

- +As in the finite-dimensional case, \(A\) is uniquely determined and is called the Fréchet derivative of \(f\) at \(x\) and is denoted \(D f(x)\).

- - -

- - -

-Let \(f: X \to \mathbb {R}\) be a (Fréchet) differentiable function and \(X_0 \subset X\) be a subspace of \(X\). Then \(\gamma _\star \) is a critical point of \(f\) with respect to \(X_0\) if +Let \(f: X \to \mathbb {R}\) be a (Fréchet) differentiable function and \(X_0 \subset X\) be a subspace of \(X\). Then \(\gamma _\star \) is a critical point of \(f\) with respect to \(X_0\) if

- + @@ -1627,7 +1659,7 @@

1.2 Hamilton’s v

\begin{equation} -X := \big \{ \gamma : [t_1, t_2] \to M \mid \gamma \mbox { is twice continuously differentiable}\big \}, +\mathcal {C}^2 := \big \{ \gamma : [t_1, t_2] \to M \mid \gamma \mbox { is twice continuously differentiable}\big \}, \end{equation}

@@ -1636,50 +1668,62 @@

1.2 Hamilton’s v equipped with the norm

- +

\begin{equation} -\|\gamma \|_X := \|\gamma \|_\infty + \|\dot \gamma \|_\infty + \|\ddot \gamma \|_\infty , +\|\gamma \|_{\mathcal {C}^2} := \|\gamma \|_\infty + \|\dot \gamma \|_\infty + \|\ddot \gamma \|_\infty , \end{equation}

-is a Banach space, and \(X_0 = \big \{h\in X \mid h(t_1) = h(t_2) = 0\big \}\) with the induced norm, is a Banach subspace of \(X\). +is a Banach space, and \(\mathcal {C}^2_0 = \big \{h\in \mathcal {C}^2 \mid h(t_1) = h(t_2) = 0\big \}\) with the induced norm, is a Banach subspace of \(\mathcal {C}^2\).

Therefore, the action defined above is a functional

- +

\begin{equation} -S : X \to \mathbb {R},\qquad \gamma \mapsto S[\gamma ], +S : \mathcal {C}^2 \to \mathbb {R},\qquad \gamma \mapsto S[\gamma ], \end{equation}

-and the evolution of the system is described by the critical points with respect to \(X_0\) of \(S\) on the space of \(\gamma \in X\) with prescribed endpoints. +and the evolution of the system is described by the critical points with respect to \(\mathcal {C}^2_0\) of \(S\) on the space of \(\gamma \in \mathcal {C}^2\) with prescribed endpoints. +

+ + + +

+1 1: Consider rewriting using q and v instead of qdot.

@@ -1689,19 +1733,19 @@

1.2 Hamilton’s v
  • - + Theorem 1.1. Let \(L = L(q, \dot q, t) : \mathbb {R}^{n}\times \mathbb {R}^{n}\times \mathbb {R} \to -\mathbb {R}\) be differentiable. The equations of motion for the mechanical system with lagrangian \(L\) are given by the Euler-Lagrange equations +\mathbb {R}\) be differentiable. The equations of motion for the mechanical system with lagrangian \(L\) are given by the Euler-Lagrange equations

    - + @@ -1724,10 +1768,61 @@

    1.2 Hamilton’s v +

  • +
    + + +
    - + -
    - - - -

    -10 Why is that the case? Assume that (1.35) is non vanishing, then you can choose \(h\) with vanishing endpoints but non vanishing on the -support of (1.35). With such an \(h\) the integral above is strictly positive, contradicting the hypothesis. -

    - - - -
    @@ -1935,18 +2021,18 @@

    1.2 Hamilton’s v
  • - -Remark 1.2. The solution \(q=q(t)\) of the Euler-Lagrange equations is just a critical point of the \(S\) + +Remark 1.3. The solution \(q=q(t)\) of the Euler-Lagrange equations is just a critical point of the \(S\) functional:

    - + @@ -1964,15 +2050,15 @@

    1.2 Hamilton’s v It does not have to be a minimum. However, under some additional conditions it can be proven to be a local minimum. For example, if the matrix of second derivatives

    - + @@ -2004,18 +2090,18 @@

    1.2 Hamilton’s v
  • - + Corollary 1.2. If the lagrangian of the system is non–degenerate, i.e., it satisfies the condition

    - + @@ -2078,17 +2164,17 @@

    1.2 Hamilton’s v

    -Here \(\left (\Lambda ^{ij}\right ) = \left (\Lambda ^{ij}(q, \dot q, t)\right )\) are the coefficients of the inverse to the matrix (1.38), i.e., +Here \(\left (\Lambda ^{ij}\right ) = \left (\Lambda ^{ij}(q, \dot q, t)\right )\) are the coefficients of the inverse to the matrix (1.40), i.e.,

    - + @@ -2103,7 +2189,7 @@

    1.2 Hamilton’s v

    -and it is crucial to remember that we are always using Einstein’s convention11. +and it is crucial to remember that we are always using Einstein’s convention2.

  • @@ -2111,10 +2197,10 @@

    1.2 Hamilton’s v
    - +

    -11 That is, we are summing over repeated indices, for example, \(y^{ji} x_i \equiv \sum _i y^{ji} x_i\). +2 That is, we are summing over repeated indices, for example, \(y^{ji} x_i \equiv \sum _i y^{ji} x_i\).

    @@ -2134,18 +2220,18 @@

    1.2 Hamilton’s v
  • - -Remark 1.3. The lagrangian of a mechanical system is defined + +Remark 1.4. The lagrangian of a mechanical system is defined only up to total derivatives. Or, in other words, the equations of motion remain unchanged if we add a total derivative to the lagrangian function:

    - + @@ -2163,7 +2249,7 @@

    1.2 Hamilton’s v

    The action \(\widetilde S\) of a system with lagrangian \(\widetilde L\) is

    - + @@ -2173,14 +2259,14 @@

    1.2 Hamilton’s v Z t2 S[q] e = e q̇, t) dt - L(q, (1.43) + L(q, (1.45) t1 d Z t2 Z t2 - = L(q, q̇, t) dt + f (q, t) dt (1.44) + = L(q, q̇, t) dt + f (q, t) dt (1.46) t1 t1 dt - = S[q] + f (q2 , t2 ) − f (q1 , t1 ). (1.45) + = S[q] + f (q2 , t2 ) − f (q1 , t1 ). (1.47) @@ -2198,13 +2284,13 @@

    1.2 Hamilton’s v As the additional \(f\)-dependent part is constant,

    - + @@ -2257,12 +2343,12 @@

    1.2 Hamilton’s v moving the two subsystems farther apart from each other, in the limit of infinite distance, the lagrangian of the full system tends to the limit lagrangian

    - + @@ -2280,8 +2366,8 @@

    1.2 Hamilton’s v -

    1.2.1 Dynamics of point particles: from Lagrange back to Newton
    - +
    1.2.1 Dynamics of point particles: from Lagrange back to Newton
    + @@ -2362,7 +2448,7 @@
    1.2.1 Dynamics of Then, the galilean principle of relativity says that the lagrangian of a closed mechanical system is invariant, modulo the sum of total derivatives, with respect to the galilean transformations.

    -
    +

    @@ -2389,7 +2475,7 @@

    1.2.1 Dynamics of

    -We will now see how we can derive Newton’s laws from Hamilton’s principle as a consequence of the galilean principle of relativity. For an alternative discussion on this topic see [[Arn89, Chapters 1.1 and 1.2].

    @@ -2400,17 +2486,17 @@
    1.2.1 Dynamics of
  • - + Theorem 1.3. The lagrangian of an isolated point particle in an inertial system of coordinates has the form

    - + @@ -2436,7 +2522,7 @@
    1.2.1 Dynamics of
  • - +
      @@ -2448,13 +2534,13 @@
      1.2.1 Dynamics of transformations implies that it must be dependent on the square of the velocities:

      - + @@ -2472,13 +2558,13 @@
      1.2.1 Dynamics of The invariance with respect to uniform motion now implies that the lagrangian must actually be proportional to \(\|\dot {\vb *{x}}\|^2\). For convenience, let’s first look at the case of small velocities: applying the galilean transformation

      - + @@ -2496,13 +2582,13 @@
      1.2.1 Dynamics of in the limit \(\epsilon \to 0\), we can apply a Taylor expansion around \(\|\dot {\vb *{x}}\|^2\) to get

      - + @@ -2520,18 +2606,18 @@
      1.2.1 Dynamics of The invariance of the equations of motion then implies that the linear term in \(\epsilon \) should be a total derivative \(\frac {\dd }{\dd t} f(t, \vb *{x}) = 2\,\langle \vb *{v}, \dot {\vb *{x}}\rangle \,L'(\|\dot {\vb *{x}}\|^2)\). This can happen iff \(\dot f(t, \vb *{x}) = 0\) and \(\langle \frac {\partial f}{\partial \vb *{x}}, \dot x\rangle = 2\,\langle \vb *{v}, \dot {\vb *{x}}\rangle L'(\|\dot {\vb *{x}}\|^2)\) is linear also in \(\|\dot {\vb *{x}}\|^2\). In particular this means that \(L'(\|\dot {\vb *{x}}\|^2) = \mathrm {const} =: \frac {m}2\), that is, the lagrangian is of the form (1.48) for some constant value \(m\in \mathbb {R}\). Finally, the lagrangian (1.48) is transformed by a +href="chapter-1.html#eq:singleptlag">1.50) for some constant value \(m\in \mathbb {R}\). Finally, the lagrangian (1.50) is transformed by a general galilean transformation \(\vb *{x} \mapsto \vb *{x} + \vb *{v}t\) into

      - + @@ -2564,7 +2650,7 @@
      1.2.1 Dynamics of
    • - + Corollary 1.4 (Newton’s first law). In an inertial frame of reference, an isolated point particle either does not move or is in uniform motion with constant velocity.

      @@ -2576,7 +2662,7 @@
      1.2.1 Dynamics of
    - +
      @@ -2584,16 +2670,16 @@
      1.2.1 Dynamics of
    • - Proof. It follows immediately by computing the Euler-Lagrange equations for (1.48): + Proof. It follows immediately by computing the Euler-Lagrange equations for (1.50):

      - + @@ -2622,13 +2708,13 @@
      1.2.1 Dynamics of Remembering the additive property of the lagrangians, we can show that for a system of \(N\) particles, which do not interact, the lagrangian is the sum

      - + @@ -2745,7 +2831,7 @@
      1.2.1 Dynamics of

      The vector \(\vb *{F}_k\) is called the force acting on the \(k\)-th point particle, and you should immediately recognize Newton’s second law in (1.56). +href="chapter-1.html#eq:newton2">1.58).

      @@ -2755,18 +2841,18 @@
      1.2.1 Dynamics of
    • - + Exercise 1.1. Prove Newton’s third law, i.e., for each of the \(k\) point particles it holds

      - + @@ -2884,18 +2970,18 @@
      1.2.1 Dynamics of
    • - + Example 1.6. The motion of \(N\) point particles with masses \(m_1, \ldots , m_N\) in their gravitational field, is described by the lagrangian

      - + @@ -2948,7 +3034,7 @@
      1.2.1 Dynamics of

      -whose equation of motion should remind you of (1.10) from Example 1.3. +whose equation of motion should remind you of (1.10) from Example 1.2.

    • @@ -2958,16 +3044,16 @@
      1.2.1 Dynamics of
    • -To understand better why \(T\) in (1.55) is called kinetic energy, it is useful to look at its relation with Newton’s second law (1.56): +To understand better why \(T\) in (1.57) is called kinetic energy, it is useful to look at its relation with Newton’s second law (1.58):

      - + @@ -2987,12 +3073,12 @@
      1.2.1 Dynamics of There is more, in fact

      - + @@ -3142,13 +3228,13 @@
      1.2.1 Dynamics of Note that the magnetic vector potential is not unique! For any function \(f\), the transformation

      - + @@ -3167,14 +3253,14 @@
      1.2.1 Dynamics of will produce the same field. This is known as gauge transformation. Under this transformation the lagrangian is transformed as

      - + @@ -3201,13 +3287,13 @@
      1.2.1 Dynamics of -

      1.3 First steps with conserved quantities

      - +

      1.3 First steps with conserved quantities

      + -
      1.3.1 Back to one degree of freedom
      - +
      1.3.1 Back to one degree of freedom
      + @@ -3216,12 +3302,12 @@
      1.3.1 Back to one Consider the equation

      - + @@ -3234,16 +3320,16 @@
      1.3.1 Back to one

      -It should come as no surprise, that introducing the auxiliary variable \(y(t) = \dot x(t)\), (1.67) is equivalent to the system of first order equations +It should come as no surprise, that introducing the auxiliary variable \(y(t) = \dot x(t)\), (1.69) is equivalent to the system of first order equations

      - + @@ -3257,16 +3343,16 @@
      1.3.1 Back to one

      -The solutions of (1.68) are parametric curves12 \((x(t),y(t)):\mathbb {R}\to \mathbb {R}^2\) in the \((x,y)\)-space. If \(y\neq 0\), we +The solutions of (1.70) are parametric curves3 \((x(t),y(t)):\mathbb {R}\to \mathbb {R}^2\) in the \((x,y)\)-space. If \(y\neq 0\), we can apply the chain rule, \(\frac {\dd y}{\dd t} = \frac {\dd y}{\dd x} \frac {\dd x}{\dd t}\), to get

      - + @@ -3280,15 +3366,15 @@
      1.3.1 Back to one

      -Reasoning formally13 for a moment, we can rephrase this as +Reasoning formally4 for a moment, we can rephrase this as

      - + @@ -3301,7 +3387,7 @@
      1.3.1 Back to one

      -Getting rid of time and considering equation (1.69) comes with a price, the solution is now an implicit curve \(y(x)\), but also with a huge advantage: this new equation can +Getting rid of time and considering equation (1.71) comes with a price, the solution is now an implicit curve \(y(x)\), but also with a huge advantage: this new equation can be solved exactly!

      @@ -3309,13 +3395,13 @@
      1.3.1 Back to one How is that so? We can separate \(x\) and \(y\) and integrate to get

      - + @@ -3332,12 +3418,12 @@
      1.3.1 Back to one If \(U(x)\) is such that \(F(x) = -\frac {\dd U}{\dd x}\), we can further simplify the equation into

      - + @@ -3353,12 +3439,12 @@
      1.3.1 Back to one where \(C = C_x - C_y \in \mathbb {R}\) is just a number due to the constants of integration. We can locally invert the equation above to get

      - + -
      1.3.2 The conservation of energy
      - +
      1.3.2 The conservation of energy
      + @@ -3569,12 +3655,12 @@
      1.3.2 The conserva derivative of the function \(I\) is zero:

      - + @@ -3595,12 +3681,12 @@
      1.3.2 The conserva In other words, if the function \(I\) remains constant along the paths followed by the system:

      - + @@ -3619,18 +3705,18 @@
      1.3.2 The conserva
    • - + Theorem 1.7. If the lagrangian of the mechanical system does not explicitly depend on time, \(L = L(q, \dot q)\), then the energy of the system

      - + @@ -3656,7 +3742,7 @@
      1.3.2 The conserva
    - +
      @@ -3667,7 +3753,7 @@
      1.3.2 The conserva Proof. Using the Euler-Lagrange equation, we have

      - + @@ -3695,7 +3781,7 @@
      1.3.2 The conserva that is,

      - + @@ -3735,18 +3821,18 @@
      1.3.2 The conserva
    • - + Example 1.9. Let’s consider \(N\) point particles in physical space with -natural lagrangian \(L = T - U\) as in (1.55). Then, +natural lagrangian \(L = T - U\) as in (1.57). Then,

      - + @@ -3765,13 +3851,13 @@
      1.3.2 The conserva and therefore

      - + @@ -3817,13 +3903,13 @@
      1.3.2 The conserva We call \(\vb *{p}_k := m_k \dot {\vb *{x}}_k\) the kinetic momentum, replacing it in the equation above we get that

      - + @@ -3855,13 +3941,13 @@
      1.3.2 The conserva
    • - -Remark 1.7. This is probably a good point to discuss the question: why does nature want to minimize the -action? And why the lagrangian is of the form (1.55)? + +Remark 1.8. This is probably a good point to discuss the question: why does nature want to minimize the +action? And why the lagrangian is of the form (1.57)?

      -Theorem 1.7 tells us that the total energy is conserved, and (1.83) tells us that for closed systems +Theorem 1.7 tells us that the total energy is conserved, and (1.85) tells us that for closed systems this implies that the energy is transferred back and forth between the kinetic and the potential components. We saw towards the end of Section 1.2.1 that while the kinetic energy measures how much the system is moving around, the potential energy measures the capacity of the system to change: its name can be intended as ‘potential’ in the sense of yet unexpressed possibilities.

      @@ -3883,8 +3969,8 @@
      1.3.2 The conserva -
      1.3.3 Fun with the phase portrait
      - +
      1.3.3 Fun with the phase portrait
      + @@ -3897,12 +3983,12 @@
      1.3.3 Fun with the A general natural lagrangians for a system with one degree of freedom in isolation has the form

      - + @@ -3923,13 +4009,13 @@
      1.3.3 Fun with the As we have seen in Section 1.3.1, the conservation of energy,

      - + @@ -3947,15 +4033,15 @@
      1.3.3 Fun with the

      -To begin with, we can use (1.86) to integrate the equations of motion +To begin with, we can use (1.88) to integrate the equations of motion

      - + @@ -3971,7 +4057,7 @@
      1.3.3 Fun with the by quadrature, i.e., solving the equation as

      - + @@ -3996,7 +4082,7 @@
      1.3.3 Fun with the and to reason about further properties of the solutions. To grasp where this comes from, it is enough to spell out

      - + @@ -4031,13 +4117,13 @@
      1.3.3 Fun with the Thanks to time reversibility, the time of travel between \(x_1^*\) and \(x_2^*\) is the same as the time to return from \(x_2^*\) to \(x_1^*\), therefore the period of oscillation is given by

      - + @@ -4089,18 +4175,18 @@
      1.3.3 Fun with the
    • - -Remark 1.8. Integral curves carry a lot more information than what immediately meets the eye. Observe + +Remark 1.9. Integral curves carry a lot more information than what immediately meets the eye. Observe for example that there is a remarkable relation between the period of an oscillation and the area of the integral curve \(H(x,y) = E\): if \(A\) is the area enclosed by the integral curve, then the period \(\mathcal {T}\) can be obtained as

      - + @@ -4123,12 +4209,12 @@
      1.3.3 Fun with the From \(H(x,y) = E\) we get \(y = \pm \sqrt {2(U(x) - E)}\), where I considered unit mass for notational simplicity. Say that the curve intercepts the \(x\)-axis at the two return points \(x_m < x_M\), then the area inside the curve, is given by

      - + @@ -4147,14 +4233,14 @@
      1.3.3 Fun with the As we saw above, the period can be computed as

      - + @@ -4186,18 +4272,18 @@
      1.3.3 Fun with the
    • - + Example 1.10. Looking back at Example 1.2, the lagrangian of a pendulum of length \(l\) and mass \(m\) is given by +href="chapter-1.html#ex:sprPen">1.1, the lagrangian of a pendulum of length \(l\) and mass \(m\) is given by

      - + @@ -4214,16 +4300,16 @@
      1.3.3 Fun with the

      thus \(E = \frac {ml^2 \dot x^2}2 - mgl \cos x\). As one can see in Figure 1.5, the motion is bounded for \(|E| \leq mgl\). For any such motion, the angle of maximal oscillation -corresponds to a solution of \(U(x) = E\), thus is given by \(E = -mgl \cos x^*_E\) for some \(x^*_E\). Finally, (1.90) implies that the period of oscillation is +corresponds to a solution of \(U(x) = E\), thus is given by \(E = -mgl \cos x^*_E\) for some \(x^*_E\). Finally, (1.92) implies that the period of oscillation is

      - + @@ -4303,17 +4389,17 @@
      1.3.3 Fun with the

      -Equation (1.98) implies that for small oscillations, the period of oscillation is also independent of the amplitude (and thus on the energy \(E\)). +Equation (1.100) implies that for small oscillations, the period of oscillation is also independent of the amplitude (and thus on the energy \(E\)). Which is the same result that you get by approximating the potential for small \(x\) by a second order polynomial

      - + @@ -4336,13 +4422,13 @@
      1.3.3 Fun with the Indeed, ignoring the constant \(C\) for conciseness, one finds the harmonic oscillator lagrangian

      - + @@ -4358,16 +4444,16 @@
      1.3.3 Fun with the

      -which we already solved in Example 1.2. In this specific case \(\omega = \sqrt {g/l}\) and the corresponding period of oscillation is +which we already solved in Example 1.1. In this specific case \(\omega = \sqrt {g/l}\) and the corresponding period of oscillation is

      - + @@ -4392,7 +4478,7 @@
      1.3.3 Fun with the
    -
    +

    @@ -4430,7 +4516,7 @@

    1.3.3 Fun with the
  • - + Example 1.11 (Double pendulum). A double pendulum consists of two point particles of masses \(m_1\) and \(m_2\), connected by massless rods of lengths \(l_1\) and \(l_2\). The first pendulum is attached to the ceiling by a fixed pivot while the second one is attached to the point particle of the first pendulum. @@ -4445,13 +4531,13 @@

    1.3.3 Fun with the The lagrangian \(L_1 = T_1 - U_1\) of the first particle is the same as for a simple pendulum:

    - + @@ -4475,13 +4561,13 @@
    1.3.3 Fun with the with \(\eta \) positive in the downward direction:

    - + @@ -4499,13 +4585,13 @@
    1.3.3 Fun with the We can substitute these values into the kinetic energy for the second particle in cartesian coordinates \((\chi ,\eta )\) to get

    - + @@ -4549,7 +4635,7 @@
    1.3.3 Fun with the

    The full lagrangian is then given by

    - + @@ -4557,11 +4643,11 @@
    1.3.3 Fun with the - L = T1 + T2 − U1 − U2 (1.106) + L = T1 + T2 − U1 − U2 (1.108) 1 1 - = (m1 + m2 )l12 ẋ21 + m2 l22 ẋ22 + m2 l1 l2 cos(x1 − x2 )ẋ1 ẋ2 (1.107) + = (m1 + m2 )l12 ẋ21 + m2 l22 ẋ22 + m2 l1 l2 cos(x1 − x2 )ẋ1 ẋ2 (1.109) 2 2 - + (m1 + m2 )gl1 cos x1 + m2 gl2 cos x2 . (1.108) + + (m1 + m2 )gl1 cos x1 + m2 gl2 cos x2 . (1.110) @@ -4602,18 +4688,18 @@
    1.3.3 Fun with the
  • - + Example 1.12 (Lissajous Figures). Consider the system of independent harmonic oscillators

    - + @@ -4640,13 +4726,13 @@
    1.3.3 Fun with the We know from the previous part of this section that the energy of each oscillator is conserved, as well as the total energy:

    - + @@ -4666,7 +4752,7 @@
    1.3.3 Fun with the three sphere” or “Hopf fibration” to have a glimpse of some related fascinating topics.

    -
    +

    @@ -4699,13 +4785,13 @@

    1.3.3 Fun with the On the configuration plane \((x_1, x_2)\), the motion that we observe is

    - + @@ -4720,7 +4806,7 @@
    1.3.3 Fun with the

    -This is contained in a rectangle \([-R_1,R_1]\times [-R_2,R_2]\). When \(\omega \) is irrational, the trajectory fills the rectangle, whereas they are closed curves inside the rectangle when it is rational. See also [[Kna18, Exercise 6.34] and [Arn89, Chapter 2.5, Example 2].

    @@ -4737,7 +4823,7 @@
    1.3.3 Fun with the
  • - + Exercise 1.3 (Huygens problem). Determine the form of even potentials, \(U(x) = U(-x)\), giving rise to isochronal oscillations, i.e., whose period \(\mathcal {T}\) does not depend on \(E\). (Beware, this is hard!)

    @@ -4746,13 +4832,13 @@
    1.3.3 Fun with the Use your result to study the following problem. A point particle of mass \(m\) moves without friction under the effect of a gravitational potential with uniform acceleration \(g\) on the curve

    - + @@ -4785,7 +4871,7 @@
    1.3.3 Fun with the
  • - + Exercise 1.4 (The isoperimetric problem). In the \((x,y)\)-plane in \(P\subset \mathbb {R}^3\) you are given a curve \(\gamma :[t_1,t_2]\to P\) connecting the points \((x_1, y_1, 0)\) and \((x_2, y_2, 0)\). Revolve this curve around the \(x\)-axis. For which curve does the corresponding surface of revolution have minimal area? @@ -4804,7 +4890,7 @@

    1.3.3 Fun with the
  • - + Exercise 1.5 (The brachistochrone problem). Let \(P\), \(Q\) be two given points in the vertical \((x,y)\)-plane, where \(Q\) lies beneath \(P\). A particle, subject to constant gravitational acceleration pointing downwards, moves from \(P\) to \(Q\), starting at rest. Determine the curve along which this motion takes the shortest time. @@ -4818,8 +4904,8 @@

    1.3.3 Fun with the -

    1.4 Euler-Lagrange equations on smooth manifolds

    - +

    1.4 Euler-Lagrange equations on smooth manifolds

    + @@ -4833,12 +4919,12 @@

    1.4 Euler-Lagrange \(\mathbb {R}^{3N}\). Assume that we already know that the system has \(n\) degrees of freedom. If we describe the motion in terms of generalized coordinates

    - + @@ -4854,13 +4940,13 @@

    1.4 Euler-Lagrange a direct application of the chain rule will allow us to rewrite the lagrangian in terms of these new coordinates as

    - + @@ -4908,12 +4994,12 @@

    1.4 Euler-Lagrange in cylindrical coordinates \((r,\phi ,z)\) it would read

    - + @@ -4929,12 +5015,12 @@

    1.4 Euler-Lagrange while in spherical coordinates \((r,\phi ,\theta )\), see also Figure 1.8, it would become

    - + @@ -4950,13 +5036,13 @@

    1.4 Euler-Lagrange The term \(g_{kl} (q)\) should also ring a bell in the context of this example: the arc length \(s\) of a parametrized curve \(q(t) : [t_1,t_2] \to \mathbb {R}^3\), is computed as

    - + @@ -4971,16 +5057,16 @@

    1.4 Euler-Lagrange

    -The strange looking object \(\dd s^2\), called line element or first fundamental form, is just a short-hand notation for the \(g\)-dependent scalar product of generalized velocities that we obtained above14: \(\dd s^2 = g_{kl}(q)\, -\dd q^k \dd q^l\). For example, in cartesian coordinates, we have +The strange looking object \(\dd s^2\), called line element or first fundamental form, is just a short-hand notation for the \(g\)-dependent scalar product of generalized velocities that we obtained above5: \(\dd s^2 = g_{kl}(q)\, \dd +q^k \dd q^l\). For example, in cartesian coordinates, we have

    - + @@ -4996,12 +5082,12 @@

    1.4 Euler-Lagrange while in cylindrical coordinates

    - + @@ -5017,12 +5103,12 @@

    1.4 Euler-Lagrange and in spherical coordinates, see Figure 1.8,

    - + @@ -5041,23 +5127,23 @@

    1.4 Euler-Lagrange

    Let’s consider the more general situation of local coordinates \((q^1, \ldots , q^n)\) on a smooth manifold \(M\). We will only recall the main terminology here, for a review of these concepts refer to [Arn89, Chapter 4.18], [Kna18, Appendix A], [MR99, -Chapter 4], your favourite book of differential geometry15 or my beautiful16 lecture notes [Ser20]. +Chapter 4], your favourite book of differential geometry6 or my beautiful7 lecture notes [Ser20].

    -Local coordinates, or charts, on \(M\) are local homeomorphisms17 +Local coordinates, or charts, on \(M\) are local homeomorphisms8

    - + @@ -5112,7 +5198,7 @@

    1.4 Euler-Lagrange

    -is the decomposition of the vector \(v\) in the local basis induced by the coordinate chart. This in particular means that \(\frac {\partial }{\partial q^i}\) is just a notation for the \(i\)-th basis versor20 in \(T_q M\). Mapping +is the decomposition of the vector \(v\) in the local basis induced by the coordinate chart. This in particular means that \(\frac {\partial }{\partial q^i}\) is just a notation for the \(i\)-th basis versor11 in \(T_q M\). Mapping \(\frac {\partial }{\partial q^i}\) to the \(i\)-th Euclidean basis versor \(e_i\) of \(\mathbb {R}^n\) allows you to consider tangent vectors in coordinates as genuine euclidean vectors applied at a specific point \(q\).

    @@ -5127,12 +5213,12 @@

    1.4 Euler-Lagrange one can compute that for any tangent vector \(v\in T_qM\),

    - + @@ -5153,13 +5239,13 @@

    1.4 Euler-Lagrange on the fibers:

    - + @@ -5182,14 +5268,14 @@

    1.4 Euler-Lagrange The lagrangian is called non–degenerate if

    - + @@ -5278,13 +5364,13 @@

    1.4 Euler-Lagrange transform in “the opposite way as \(\dot q\)” in the sense that:

    - + @@ -5303,14 +5389,14 @@

    1.4 Euler-Lagrange involves the inverse of the jacobian of \(q\mapsto \widetilde q(q)\). Furthermore, show that the matrix of second derivatives with respect to the coordinates \(\dot q\) of the lagrangian transforms as a \((0,2)\)-tensor, i.e.

    - + @@ -5346,12 +5432,12 @@

    1.4 Euler-Lagrange the space of smooth curves \(q(t): [t_1,t_2] \to M\) with \(q(t_1) = q_1\) and \(q(t_2) = q_2\) as the integral

    - + @@ -5388,7 +5474,7 @@

    1.4 Euler-Lagrange

    -which have the exact same form as (1.24). Under the time-independence and non–degeneracy conditions, the lagrangian defines a dynamical system on the tangent +which have the exact same form as (1.25). Under the time-independence and non–degeneracy conditions, the lagrangian defines a dynamical system on the tangent bundle \(TM\).

    @@ -5402,12 +5488,12 @@

    1.4 Euler-Lagrange \((0,2)\)-tensor on \(TM\) whose matrix representation \(g_{ij}(q)\) is positive definite. Locally, analogously to what we have described at the beginning of the chapter,

    - + @@ -5420,16 +5506,16 @@

    1.4 Euler-Lagrange

    -The length of curves on \(M\) is then defined as in (1.118) and angles between tangent vectors are defined using a generalization of the law of cosines where the Euclidean +The length of curves on \(M\) is then defined as in (1.120) and angles between tangent vectors are defined using a generalization of the law of cosines where the Euclidean scalar product is replaced by the inner product on \(T_qM\):

    - + @@ -5445,12 +5531,12 @@

    1.4 Euler-Lagrange As in the Euclidean space, arc lengths of curves are invariant with respect to monotonic changes of parametrization,

    - + @@ -5466,12 +5552,12 @@

    1.4 Euler-Lagrange Replacing the scalar product, i.e., the inner product in the Euclidean space, we can define the kinetic energy of a point particle with mass \(m\) as

    - + @@ -5484,10 +5570,10 @@

    1.4 Euler-Lagrange

    - +

    -21 Truth be told, we are hiding some extra complications involving infinite dimensional manifolds here, for a full account of the theory involved you can refer to [12 Truth be told, we are hiding some extra complications involving infinite dimensional manifolds here, for a full account of the theory involved you can refer to [MR99, Chapters 7 and 8].

    @@ -5508,13 +5594,13 @@

    1.4 Euler-Lagrange Exercise 1.7. Show that the Euler-Lagrange equations for the functional

    - + @@ -5533,13 +5619,13 @@

    1.4 Euler-Lagrange have the form

    - + @@ -5557,13 +5643,13 @@

    1.4 Euler-Lagrange where \(\Gamma _{ij}^k(q)\) are the Christoffel symbols of the Levi-Civita connection associated with the metric \(\dd s^2\):

    - + @@ -5590,7 +5676,7 @@

    1.4 Euler-Lagrange

    -Solutions of (1.138) are the geodesic curves on the riemannian manifold \(M\). Locally, geodesics are not just minimizing the action \(S\), but also the lengths! +Solutions of (1.140) are the geodesic curves on the riemannian manifold \(M\). Locally, geodesics are not just minimizing the action \(S\), but also the lengths!

    @@ -5600,7 +5686,7 @@

    1.4 Euler-Lagrange
  • - + Exercise 1.8.

      @@ -5608,7 +5694,7 @@

      1.4 Euler-Lagrange
    • -1. Derive formula (1.83) for a natural lagrangian on a riemannian manifold \((M,g)\). +1. Derive formula (1.85) for a natural lagrangian on a riemannian manifold \((M,g)\).

      @@ -5617,7 +5703,7 @@

      1.4 Euler-Lagrange

      -2. Given any lagrangian of the form \(L = L(q, \dot q)\) on \(TM\), show that the definition of energy (1.78) does not depend on +2. Given any lagrangian of the form \(L = L(q, \dot q)\) on \(TM\), show that the definition of energy (1.80) does not depend on the choice of local coordinates on \(M\).
      Hint: look at Exercise 1.6.

      @@ -5630,11 +5716,11 @@

      1.4 Euler-Lagrange

      3. Consider the following free lagrangian on a riemannian manifold \((M, g)\):

      - +

      @@ -5666,23 +5752,23 @@

      1.4 Euler-Lagrange
    • - + Example 1.13. Let us consider a spherical pendulum of mass \(m\) and length \(l\), i.e., a point particle of mass \(m\) moving without friction on the surface of a sphere of radius \(l\). Let’s ignore gravity for the moment.

      The system has two degrees of freedom, its configuration space \(M\) is a 2-sphere \(S^2\) of radius \(l\). Using spherical coordinates \((\theta , \phi )\) – see also Figure 1.8 – and (1.121), we find the free lagrangian +class="textup">(1.123), we find the free lagrangian

      - + @@ -5714,9 +5800,9 @@

      1.4 Euler-Lagrange
    • - + Exercise 1.9. Start from a free lagrangian in cartesian coordinates. As we are imposing that the particle -lives on the surface of the sphere, change variables using the appropriate spherical coordinates to derive (1.141). +lives on the surface of the sphere, change variables using the appropriate spherical coordinates to derive (1.143).

    • @@ -5725,7 +5811,7 @@

      1.4 Euler-Lagrange

  • -
    +

    @@ -5765,15 +5851,15 @@

    1.4 Euler-Lagrange

    -In analogy with our discussion of natural lagrangians (1.55), on riemannian manifolds we call natural the lagrangians of the form +In analogy with our discussion of natural lagrangians (1.57), on riemannian manifolds we call natural the lagrangians of the form

    - + @@ -5796,10 +5882,10 @@

    1.4 Euler-Lagrange
  • - -Remark 1.9. It is often necessary to deal with mechanical systems in which the interactions between + +Remark 1.10. It is often necessary to deal with mechanical systems in which the interactions between different bodies take the form of constraints. These constraints can be very complicated and hard to incorporate when working directly with the differential equations of motion. It is often easier to construct a lagrangian out of the kinetic and -potential energy of the system and derive the correct equations of motion from there. Mechanical lagrangians will often be of the form (1.142). +potential energy of the system and derive the correct equations of motion from there. Mechanical lagrangians will often be of the form (1.144).

  • diff --git a/chapter-2.html b/chapter-2.html index 4b996bd..7240ea7 100644 --- a/chapter-2.html +++ b/chapter-2.html @@ -234,7 +234,7 @@ - + @@ -278,179 +278,185 @@

    - + 1.2 Hamilton’s variational principle

    - + 1.3 First steps with conserved quantities

    - + 1.4 Euler-Lagrange equations on smooth manifolds

    - + 2 Conservation Laws

    - + 2.1 Noether theorem

    - + 2.2 The spherical pendulum

    - + 2.3 Intermezzo: small oscillations

    - + 2.4 Motion in a central potential

    - + 2.5 D’Alembert principle and systems with constraints

    - + 3 Hamiltonian mechanics

    - + 3.1 The Legendre Transform in the euclidean plane

    - + 3.2 Intermezzo: cotangent bundle and differential forms

    - + 3.3 The Legendre Transform

    - + 3.4 Poisson brackets and first integrals of motion

    - + 3.5 Variational principles of hamiltonian mechanics

    - + 3.6 Canonical transformations

    - + 3.7 The symplectic structure on the cotangent bundle

    - + 4 Integrable systems

    - + 4.1 Lagrangian submanifolds

    - + 4.2 Canonical transformations revisited

    - + 4.3 Hamilton-Jacobi equations

    - + 4.4 The Liouville-Arnold theorem

    - + 5 Hamiltonian perturbation theory

    - + 5.1 Small oscillations revisited

    - + 5.2 Birkhoff normal forms

    - + 5.3 A brief look at KAM theory

    - + 5.4 Nekhoroshev theorem

    - + 6 Conclusion

    - + Bibliography

    +

    + +Index +

    + +

  • @@ -643,14 +649,14 @@

    Hamiltonian Mechanics

    -

    2 Conservation Laws

    - +

    2 Conservation Laws

    + -

    2.1 Noether theorem

    - +

    2.1 Noether theorem

    +

    @@ -676,7 +682,7 @@

    2.1 Noether theore
  • - + Example 2.1 (kinetic momentum). Consider a mechanical system with \(n\) degrees of freedom described by the lagrangian \(L=L(q,\dot q)\), \(q=(q^1,\ldots ,q^n)\). If there is \(i\) such that \(L\) does not depend on \(q^i\), i.e.

    @@ -894,7 +900,7 @@

    2.1 Noether theore

    - +

    1 The operation \(\Phi ^*:C^\infty (TM) \to C^\infty (TM)\) defined here is a special case of a more general operation, called pullback, which we will see again in different settings in future chapters. @@ -987,7 +993,7 @@

    2.1 Noether theore
  • - + Remark 2.1. To any one–parameter group of diffeomorphism we can associate a smooth vector field \(X : q\in M \mapsto X(q)\in T_qM\) by

    @@ -1008,7 +1014,7 @@

    2.1 Noether theore \begin{equation} -X(q) := \frac {\dd }{\dd s} \Phi _s(q) \Big |_{s=0}. +X(q) := \dv {s} \Phi _s(q) \Big |_{s=0}. \end{equation} @@ -1095,7 +1101,7 @@

    2.1 Noether theore
  • - + Exercise 2.1. Show that \(\Phi _s\) as defined in (2.11) is a one–parameter group of diffeomorphisms on \(M\) for \(s,t \in (-\epsilon ,\epsilon )\) such that \(|s + t| < \epsilon \).

    @@ -1123,7 +1129,7 @@

    2.1 Noether theore

    - +

    3 For a pictorial glimpse to her life, look for #Noethember on instagram or 2.1 Noether theore \begin{equation} -I(q,\dot q) := p_i X^i, \quad p_i(q, \dot q) := \frac {\partial L}{\partial \dot q^i}, \quad X(q) := \frac {\dd }{\dd s} \Phi _s(q) \Big |_{s=0}. +I(q,\dot q) := p_i X^i, \quad p_i(q, \dot q) := \frac {\partial L}{\partial \dot q^i}, \quad X(q) := \dv {s} \Phi _s(q) \Big |_{s=0}. \end{equation} @@ -1186,7 +1192,7 @@

    2.1 Noether theore

    - +
      @@ -1194,7 +1200,7 @@

      2.1 Noether theore
    • - Proof. By definition, see (1.76), \(I\) is a first integral if it is constant on the + Proof. By definition, see (1.78), \(I\) is a first integral if it is constant on the phase curves. Let’s compute its time derivative,

      @@ -1304,7 +1310,7 @@

      2.1 Noether theore
    • - + Example 2.2 (kinetic momentum - reprise). Let’s revisit Example 2.1. For a cyclic coordinate \(q^i\), the lagrangian is invariant with respect to the one–parameter group of translations along the \(q^i\) direction:

      @@ -1370,7 +1376,7 @@

      2.1 Noether theore
    • - + Remark 2.2. Noether theorem is a more general result related to symmetries in classical field theories than what we presented here. The proof of the more general Noether theorem is obtained by reducing the system to a system with one cyclic variable. This can be done by locally rectifying4 the vector field \(X\) on \(M\) and observing that, by Exercise 1.6, the sum \(p_i X^i\) does not depend on the choice of coordinates on \(M\). @@ -1381,7 +1387,7 @@

      2.1 Noether theore

    - +

    4 See [Ser20, Chapter 3.5]. @@ -1400,7 +1406,7 @@

    2.1 Noether theore
  • - + Exercise 2.2. Assume that the one–parameter group of diffeomorphisms \(\Phi _s\) is not a symmetry for \((M,L)\) in the sense of our definition, but preserves the action \(S\) associated to the lagrangian. In that case, the invariance condition (2.17) holds only up to total derivatives [GF00, Chapter 4.20], i.e., there is a function \(f(q)\) on \(M\) such that @@ -1471,7 +1477,7 @@

    2.1 Noether theore
  • - + Remark 2.3. There are also discrete symmetries in nature which don’t depend on a continuous parameter. The so-called parity, for example, is the invariance under reflection, \(\vb *{x} \mapsto -\vb *{x}\). These types of symmetries give rise to conservation laws in quantum physics but not in classical physics.

    @@ -1488,12 +1494,12 @@

    2.1 Noether theore -

    2.1.1 Homogeneity of space: total momentum
    - +
    2.1.1 Homogeneity of space: total momentum
    +

    -Consider a system of \(N\) particles with lagrangian (1.55), +Consider a system of \(N\) particles with lagrangian (1.57),

    @@ -1628,7 +1634,7 @@
    2.1.1 Homogeneity

    - +

    5 It may seem here that we forgot to push forward \(\dot x\) by \(\Phi _s\), however the pushforward \((\Phi _s)_* = \Id \) by \(\Phi _s\) acts on \(\dot x\) as the identity. @@ -1802,7 +1808,7 @@

    2.1.1 Homogeneity
    - +

    6 Beware, here (and a few other times below) \(M\) denotes a constant, to not be confused with the base manifold. What is what should be inferrable from the constants. @@ -1821,7 +1827,7 @@

    2.1.1 Homogeneity
  • - + Exercise 2.3. Consider, as in Example 1.7 and Exercise 1.2, a closed system of charged point particles in the presence of a uniform constant magnetic field \(\vb *{B}\). Define

    @@ -1860,8 +1866,8 @@
    2.1.1 Homogeneity -
    2.1.2 Isotropy of space: angular momentum
    - +
    2.1.2 Isotropy of space: angular momentum
    + @@ -1960,7 +1966,7 @@
    2.1.2 Isotropy of
  • - + Proposition 2.3. Let \(A(s)\) be a one–parameter family of orthogonal matrices of the form (2.34). Then,

    @@ -1997,7 +2003,7 @@
    2.1.2 Isotropy of
  • - +
      @@ -2103,7 +2109,7 @@
      2.1.2 Isotropy of
    • - + Exercise 2.4 (Euler theorem). Given an antisymmetric matrix

      @@ -2216,7 +2222,7 @@
      2.1.2 Isotropy of
    • - + Theorem 2.4. If a closed system of \(N\) particles is invariant with respect to rotations, then the components of its total angular momentum

      @@ -2253,7 +2259,7 @@
      2.1.2 Isotropy of

      - +
        @@ -2375,9 +2381,9 @@
        2.1.2 Isotropy of
      • - + Example 2.4. Consider a natural lagrangian (1.55) with an axial symmetry, i.e., whose potential is invariant by simultaneous rotations around a fixed axis. For convenience let’s say that such axis corresponds to the \(z\) axis. +href="chapter-1.html#eq:mechlag">1.57) with an axial symmetry, i.e., whose potential is invariant by simultaneous rotations around a fixed axis. For convenience let’s say that such axis corresponds to the \(z\) axis. Looking back at the proof above, we can infer that the conserved quantity will not be the full total momentum \(\vb *{M} = (M_x, M_y, M_z)\) but instead only its \(z\) component:

        @@ -2421,8 +2427,8 @@
        2.1.2 Isotropy of -
        2.1.3 Homogeneity of time: the energy
        - +
        2.1.3 Homogeneity of time: the energy
        +

        @@ -2441,12 +2447,12 @@

        2.1.3 Homogeneity -
        2.1.4 Scale invariance: Kepler’s third law
        - +
        2.1.4 Scale invariance: Kepler’s third law
        +

        -We have mentioned in Remark 1.3 that the two lagrangians \(L\) and \(\alpha L\), \(\alpha \neq 0\), give rise to the same equations of motion. This does not look like a symmetry in +We have mentioned in Remark 1.4 that the two lagrangians \(L\) and \(\alpha L\), \(\alpha \neq 0\), give rise to the same equations of motion. This does not look like a symmetry in our sense, but it is worth investigating its meaning in light of what we have seen so far.

        @@ -2457,9 +2463,9 @@
        2.1.4 Scale invar
      • - + Theorem 2.5. Let \(L\) be a natural lagrangian \(L=T-U\) as in (1.55) whose potential is a homogeneous function of degree \(k\): +href="chapter-1.html#eq:mechlag">1.57) whose potential is a homogeneous function of degree \(k\):

        @@ -2518,7 +2524,7 @@
        2.1.4 Scale invar
      • - +
          @@ -2640,7 +2646,7 @@
          2.1.4 Scale invar
        • - + Remark 2.4. The reasoning above can be repeated for any Lagrangian \(L = T - U\) where \(T\) is homogeneous of degree \(k\) and \(U\) is homogeneous of degree \(l\), obtaining a similar result.

          @@ -2658,7 +2664,7 @@
          2.1.4 Scale invar
        • - + Example 2.5 (Harmonic oscillator). The potential of the harmonic oscillator

          @@ -2701,9 +2707,9 @@
          2.1.4 Scale invar
        • - + Example 2.6 (Kepler’s third law). Recall Kepler’s problem -lagrangian from Examples 1.3 and 1.6: +lagrangian from Examples 1.2 and 1.6:

          @@ -2763,8 +2769,8 @@
          2.1.4 Scale invar -

          2.2 The spherical pendulum

          - +

          2.2 The spherical pendulum

          + @@ -2886,7 +2892,7 @@

          2.2 The spherical
        • - + Remark 2.5. Be careful here: Euler-Lagrange equations are derived under the assumption that the coordinates are independent. If you replace the quantity \(I\) in the lagrangian itself, you will get an incorrect result!

          @@ -2985,7 +2991,7 @@

          2.2 The spherical
        • - + Exercise 2.5. Use the phase portrait to describe the behavior of \(\theta (t)\), and the motion of the pendulum. Make sure to use the correct intervals of definition of \(\phi \) and \(\theta \).

          @@ -3006,8 +3012,8 @@

          2.2 The spherical -

          2.3 Intermezzo: small oscillations

          - +

          2.3 Intermezzo: small oscillations

          + @@ -3023,7 +3029,7 @@

          2.3 Intermezzo: s

          Let’s revisit what we saw in Sections 1.3.1 and 1.3.3 in slightly more generality and in a more organic way. Let’s consider the Newton equation for a -closed system with one degree of freedom in the presence of conservative forces, i.e., an equation like (1.67): +closed system with one degree of freedom in the presence of conservative forces, i.e., an equation like (1.69):

          @@ -3318,7 +3324,7 @@

          2.3 Intermezzo: s
        • - + Theorem 2.6. A system performing small oscillations is the direct product of \(n\) one-dimensional systems performing small oscillations.

          @@ -3381,7 +3387,7 @@

          2.3 Intermezzo: s
        • - + Theorem 2.7. The system (2.78) has \(n\) characteristic oscillations, the directions of which are pairwise orthogonal with respect to the scalar product given by the kinetic energy.

          @@ -3403,7 +3409,7 @@

          2.3 Intermezzo: s
        • - + Corollary 2.8. Every small oscillation is a sum of characteristic oscillations.

          @@ -3424,7 +3430,7 @@

          2.3 Intermezzo: s
        • - + Remark 2.6. Note that, in general, it is not possible to simultaneously diagonalize three symmetric matrices.

          @@ -3436,8 +3442,8 @@

          2.3 Intermezzo: s -

          2.3.1 Decomposition into characteristic oscillations
          - +
          2.3.1 Decomposition into characteristic oscillations
          + @@ -3530,7 +3536,7 @@
          2.3.1 Decompositi
        • - + Remark 2.7. As opposed to general systems of linear differential equations, resonance terms of the form \(t^\alpha e^{i \omega t}\), \(\alpha \in \big \{1,2,\ldots ,n\big \}\), do not arise in lagrangian systems (even in the case of eigenvalues with multiplicity).

          @@ -3547,8 +3553,8 @@
          2.3.1 Decompositi -
          2.3.2 The double pendulum
          - +
          2.3.2 The double pendulum
          +

          @@ -3695,8 +3701,8 @@

          2.3.2 The double -
          2.3.3 The linear triatomic molecule
          - +
          2.3.3 The linear triatomic molecule
          + @@ -3856,8 +3862,8 @@
          2.3.3 The linear -
          2.3.4 Zero modes
          - +
          2.3.4 Zero modes
          +

          @@ -3941,7 +3947,7 @@

          2.3.4 Zero modes<
        • - + Remark 2.8. We have already seen an example of this in Section 2.3.3: the normal mode \(\vb *{\xi }_1=(1,1,1)^T\) of the triatomic molecule is the zero mode corresponding to the invariance of the system with respect to horizontal translations.

          @@ -3959,8 +3965,8 @@
          2.3.4 Zero modes< -

          2.4 Motion in a central potential

          - +

          2.4 Motion in a central potential

          +
            @@ -3969,9 +3975,9 @@

            2.4 Motion in a c
          • - + Example 2.7 (The two-body problem). Let’s go back to Examples 1.3 and 1.6. Consider a closed systems of two point particles with masses \(m_{1,2}\). We now +id="ex:kepler2"> Let’s go back to Examples 1.2 and 1.6. Consider a closed systems of two point particles with masses \(m_{1,2}\). We now know that, in an inertial frame of reference, their natural lagrangian must have the form

            @@ -4185,7 +4191,7 @@

            2.4 Motion in a c
          • - + Theorem 2.9. The trajectories of (2.107) are planar.

            @@ -4197,7 +4203,7 @@

            2.4 Motion in a c

          - +
            @@ -4289,7 +4295,7 @@

            2.4 Motion in a c
          • - + Exercise 2.6. Use the conservation law

            @@ -4332,7 +4338,7 @@

            2.4 Motion in a c
          • - + Theorem 2.10. The equations of motion describing a point particle in a central potential can always be solved by quadrature.

            @@ -4344,7 +4350,7 @@

            2.4 Motion in a c

          - +
            @@ -4566,7 +4572,7 @@

            2.4 Motion in a c region as depicted in Figure 2.1.

            -
            +

            @@ -4583,7 +4589,7 @@

            2.4 Motion in a c

            -Figure 2.1: The trajectories are confined in an annular region [LL76] +Figure 2.1: The trajectories are confined in an annular region [LL76]

            @@ -4600,7 +4606,7 @@

            2.4 Motion in a c
          • - + Exercise 2.7. Consider the integral (2.119), with \(r_0 = r_m\) and \(r = r_M\), one can use it to define the quantity

            @@ -4668,12 +4674,12 @@

            2.4 Motion in a c -

            2.4.1 Kepler’s problem
            - +
            2.4.1 Kepler’s problem
            +

            -Let’s go back to Example 2.7, and replace \(U\) with the Newtonian gravitational potential \(k/\|\vb *{x}\|\) (see Examples 1.3 and 2.7, and replace \(U\) with the Newtonian gravitational potential \(k/\|\vb *{x}\|\) (see Examples 1.2 and 1.6). The effective potential (2.120), in this case is

            @@ -4727,7 +4733,7 @@
            2.4.1 Kepler’s after which the potential grows approaching \(0\) as \(r\to +\infty \).

            -
            +

            @@ -4842,7 +4848,7 @@

            2.4.1 Kepler’s

            -We can now confirm the last claim in Example 1.3. Indeed, with some knowledge about conic sections one can conclude that (1.2. Indeed, with some knowledge about conic sections one can conclude that (2.129) is an ellipse for \(E<0\), when \(0<e<1\). The semiaxes of the ellipse are then given by

            @@ -4868,7 +4874,7 @@
            2.4.1 Kepler’s Furthermore, when \(E=0\), i.e., \(e = 1\), we have a parabola and, finally, for \(E>0\), i.e., \(e >1\) an hyperbola.

            -
            +

            @@ -5145,7 +5151,7 @@

            2.4.1 Kepler’s
          • - + Exercise 2.8. Obtain the following parametrization for the hyperbolic motion for \(E>0\).

            @@ -5202,7 +5208,7 @@
            2.4.1 Kepler’s
          • - + Exercise 2.9. Obtain the following parametrization for the parabolic motion for \(E=0\).

            @@ -5265,7 +5271,7 @@
            2.4.1 Kepler’s
          • - + Exercise 2.10. Let \(0<e<1\) and \(\tau = \theta - e\sin \theta \). Show that \(\sin \theta \) and \(\cos \theta \) can be expanded in Fourier series as

            @@ -5363,7 +5369,7 @@
            2.4.1 Kepler’s
          • -In Example 1.3, we mentioned a third conserved quantity for the Kepler problem. +In Example 1.2, we mentioned a third conserved quantity for the Kepler problem.

            @@ -5373,7 +5379,7 @@
            2.4.1 Kepler’s
          • - + Exercise 2.11. Show that the components of the Laplace-Runge-Lenz vector

            @@ -5415,7 +5421,7 @@
            2.4.1 Kepler’s
          • - + Exercise 2.12. Consider a small perturbation of the gravitational potential \(U(r) = -\frac kr\) of the form

            @@ -5453,8 +5459,8 @@
            2.4.1 Kepler’s -

            2.5 D’Alembert principle and systems with constraints

            - +

            2.5 D’Alembert principle and systems with constraints

            + @@ -5470,7 +5476,7 @@

            2.5 D’Alembert

            -Let’s try again. Consider now a natural lagrangian (1.55) +Let’s try again. Consider now a natural lagrangian (1.57)

            @@ -5528,7 +5534,7 @@

            2.5 D’Alembert
          • - + Remark 2.9. We call the constraints holonomic if the functions that define them do not depend on the velocities (they can depend on time, their treatment is analogous to what we will do in this section). More general constraints, also taking into account the velocities, are called non holonomic and will not be treated in this course.

            @@ -5577,9 +5583,9 @@

            2.5 D’Alembert
          • - + Theorem 2.11. The equations of motion of a mechanical system of \(N\) particles with natural lagrangian (1.55) under the effect of \(l\) independent holonomic constraints (2.152) have the form +class="textup">(1.57) under the effect of \(l\) independent holonomic constraints (2.152) have the form

            @@ -5614,7 +5620,7 @@

            2.5 D’Alembert

          • - +
              @@ -5752,7 +5758,7 @@

              2.5 D’Alembert

            - +

            7 For a brief introduction on their meaning, I recommend this twitter thread or this 2.5 D’Alembert

          • - + Remark 2.10. The theorem above is often stated as follows: in a constrained system, the total work of the constraint forces on any virtual variations, i.e., vectors \(\delta \vb *{x}\) tangent to the submanifold \(Q\), is zero. One can understand this by observing that for such tangent vectors

            @@ -5834,9 +5840,9 @@

            2.5 D’Alembert
          • - + Theorem 2.12. The evolution of of a mechanical system on \(N\) particles with natural lagrangian (1.55) under the effect of \(l\) independent holonomic constraints (2.152) can be described by +class="textup">(1.57) under the effect of \(l\) independent holonomic constraints (2.152) can be described by mechanical system on \(TQ\) with lagrangian

            @@ -5959,7 +5965,7 @@

            2.5 D’Alembert
          • - + Exercise 2.13. Show that the equations of motion of the constrained motion on \(Q\) are given by the Euler-Lagrange equations of the lagrangian (2.162).

            @@ -5977,7 +5983,7 @@

            2.5 D’Alembert
          • - + Example 2.8. Every submanifold \(Q\subset \mathbb {R}^m\) of the euclidean space is regular with respect to the free lagrangian

            @@ -6051,7 +6057,7 @@

            2.5 D’Alembert
          • - + Exercise 2.14. Consider the lagrangian \(L=\frac 12 g_{ij}(q)\dot q^i \dot q^j\) on the tangent bundle of a riemannian manifold \(Q\) with a metric \(\dd s^2 = g_{ij}(q)\dd q^i \dd q^j\). Show that the solutions \(q=q(t)\) of the Euler-Lagrange equations for \(L\) satisfy \(\langle \dot q, \dot q\rangle = \mathrm {const}\).

            @@ -6069,7 +6075,7 @@

            2.5 D’Alembert
          • - + Exercise 2.15 (Clairaut’s relation). Given a function \(f(z): [a,b] \to \mathbb {R}_+\), consider the surface of revolution in \(\mathbb {R}^3\) implicitly defined by

            @@ -6113,7 +6119,7 @@

            2.5 D’Alembert
          • - + Example 2.9 (Double pendulum revisited). Let’s see if we can re-derive the double pendulum lagrangian using D’Alembert principle. We have a system of two point masses with masses \(m_1\) and \(m_2\) connected by massless rigid rods: one of length \(l_2\) connecting them to each other and one of length \(l_1\) connecting the first to a pivot on the ceiling. We assume that frictions are absent and that the double pendulum is constrained to move in a single plane. @@ -6168,7 +6174,7 @@

            2.5 D’Alembert
          • - + Exercise 2.16. Use D’Alembert principle to compute the constraint force acting on the spherical pendulum9 under the effect of the gravitational potential.

            @@ -6182,7 +6188,7 @@

            2.5 D’Alembert

          @@ -643,8 +649,8 @@

          Hamiltonian Mechanics

          -

          3 Hamiltonian mechanics

          - +

          3 Hamiltonian mechanics

          +

          @@ -717,8 +723,8 @@

          3 Hamiltonian mec -

          3.1 The Legendre Transform in the euclidean plane

          - +

          3.1 The Legendre Transform in the euclidean plane

          +

          @@ -917,7 +923,7 @@

          3.1 The Legendre

          - +

          1 Metaphorically, but I should probably add a sketch here.

          @@ -932,8 +938,8 @@

          3.1 The Legendre -

          3.2 Intermezzo: cotangent bundle and differential forms

          - +

          3.2 Intermezzo: cotangent bundle and differential forms

          +

          @@ -1212,8 +1218,8 @@

          3.2 Intermezzo: c -

          3.3 The Legendre Transform

          - +

          3.3 The Legendre Transform

          +

          @@ -1277,7 +1283,7 @@

          3.3 The Legendre
        • - + Theorem 3.1. The Legendre transform of a non–degenerate lagrangian is a local diffeomorphism.

          @@ -1288,7 +1294,7 @@

          3.3 The Legendre

        • - +
            @@ -1426,7 +1432,7 @@

            3.3 The Legendre >

            -We call hamiltonian of a mechanical system with lagrangian \(L\), the Legendre transform \(H(q,p)\) of its energy \(E(q,\dot q)\) from (1.78). In local +We call hamiltonian of a mechanical system with lagrangian \(L\), the Legendre transform \(H(q,p)\) of its energy \(E(q,\dot q)\) from (1.80). In local coordinates the hamiltonian of the system is given by

            @@ -1458,7 +1464,7 @@

            3.3 The Legendre
          • - + Remark 3.1. This is what is meant when you read that the hamiltonian is the lift of the total energy \(E(q, \dot q)\) of the system to a function \(H(q,p)\) on the cotangent bundle \(T^*M\).

            @@ -1476,7 +1482,7 @@

            3.3 The Legendre
          • - + Remark 3.2. This is where the literature can disagree. Some textbooks define \(H\) to be the Legendre transform of \(L\) instead. In this way the definition can be rewritten3 as

            @@ -1496,7 +1502,7 @@

            3.3 The Legendre

          - +

          3 This is very nicely discussed in this blog post. @@ -1515,7 +1521,7 @@

          3.3 The Legendre
        • - + Remark 3.3. Perhaps not surprisingly, the Legendre transform also comes with some interesting connections to control theory and differential geometry. Here I will focus on the latter. For a more detailed discussion about the relations to control theory, you can refer to [GF00, Appendix II]. We will not go into the details here and instead only point out the bare-bones of the connection. Optimization problems can often be reduced to find a critical point of some functional, like the action, with respect to some extra external constraints, @@ -1574,7 +1580,7 @@

          3.3 The Legendre \begin{align} -\mathbb {F}L(v) w := \frac {\dd }{\dd s} L(q, v + s w) \big |_{s=0}. +\mathbb {F}L(v) w := \dv {s} L(q, v + s w) \big |_{s=0}. \end{align} That is, \(\mathbb {F}L(v) w\) is the derivative of \(L\) at \(v\) in the direction \(w\) along the fiber \(T_qM\). The last observation is crucial: the map is fiber-preserving in the sense that it maps the fiber \(T_qM\) to the fiber \(T^*_qM\). If we denote local coordinates on \(M\) by \(q=(q^1, \ldots , q^n)\), vectors on \(T_qM\) take the form \(\dot q = \dot q^i \frac {\partial }{\partial q^i}\big |_q \simeq (\dot q^1, \ldots , \dot q^n)\) and the fiber derivative can be @@ -1653,7 +1659,7 @@

          3.3 The Legendre
        • - + Theorem 3.2 (Hamilton’s equations). Euler-Lagrange equations for a lagrangian \(L(q,\dot q)\) are equivalent to Hamilton’s equations

          @@ -1699,7 +1705,7 @@

          3.3 The Legendre

        • - +
            @@ -1808,7 +1814,7 @@

            3.3 The Legendre
          • - + Example 3.1 (A particle in a potential). Let’s start with a simple prototypical example: a particle in \(\mathbb {R}^3\) moving under the influence of a potential \(U(\vb *{x})\). We are, by now, very accustomed to the corresponding natural lagrangian

            @@ -1938,7 +1944,7 @@

            3.3 The Legendre
          • - + Example 3.2 (The hamiltonian for a natural lagrangian on a riemannian manifold). Let \(L\) be a free lagrangian

            @@ -2090,7 +2096,7 @@

            3.3 The Legendre
          • - + Exercise 3.1. Show that a simple substitution allows to recover Euler-Lagrange equations of Exercise 1.7, which gives the geodesic flow on the tangent bundle \(TM\).

            @@ -2119,7 +2125,7 @@

            3.3 The Legendre
          • - + Remark 3.4. The previous example provides a good opportunity to discuss the nature of the momentum from yet another perspective. Let \((M, g)\) be a riemannian manifold. The riemannian metric, or riemannian metric tensor, is a bilinear form \(\langle \cdot , \cdot \rangle _g\) on the tangent spaces of \(M\). In local coordinates \(q\in M\), we can associate to the metric a matrix \(g_{ij}(q)\). With an abuse of notation, we say that \(g_{ij}(q)\) is a \((0,2)\)-tensor. @@ -2213,7 +2219,7 @@

            3.3 The Legendre

          - +

          4 The isomorphisms between the tangent and cotangent spaces defined here are commonly known as musical isomorphisms: the flat and sharp maps respectively correspond to the mapping \(\beta @@ -2234,7 +2240,7 @@

          3.3 The Legendre
        • - + Example 3.3. We saw in Example 3.1 that the hamiltonian of a free particle of mass \(m\) is, in cartesian coordinates,

          @@ -2366,7 +2372,7 @@

          3.3 The Legendre
        • - + Exercise 3.2 (Inverse Legendre transform). Show that the hamiltonian \(H\) defined by (3.26) satisfies the non–degeneracy condition

          @@ -2435,7 +2441,7 @@

          3.3 The Legendre
        • - + Exercise 3.3. Show that the Legendre transform of a non–degenerate time-dependent lagrangian \(L(q,\dot q, t)\) is well defined, the corresponding hamiltonian is

          @@ -2502,7 +2508,7 @@

          3.3 The Legendre
        • - + Example 3.4 (A particle in an electromagnetic field). We saw the Lagrangian for charged particles in an electromagnetic field in Example 1.7 and in Exercise 1.2. In the simpler case of a single particle, the lagrangian takes the form @@ -2674,7 +2680,7 @@

          3.3 The Legendre
        • - + Example 3.5 (A particle in a constant magnetic field). An important special case of Example 3.4 is the case of a uniform magnetic field pointing in the \(z\)-direction: \(\vb * B = (0,0,B)\). One convenient choice of vector potential for \(\vb * B\) is

          @@ -2867,7 +2873,7 @@

          3.3 The Legendre
        • - + Example 3.6 (The relativistic hamiltonian). If \(c>0\) denotes the speed of light, the hamiltonian \(H:\mathbb {R}^3\times \mathbb {R}^3\to \mathbb {R}\) of a free relativistic particle of mass \(m>0\) is

          @@ -2988,8 +2994,8 @@

          3.3 The Legendre -

          3.4 Poisson brackets and first integrals of motion

          - +

          3.4 Poisson brackets and first integrals of motion

          + @@ -3074,7 +3080,7 @@

          3.4 Poisson brack
        • - + Exercise 3.4. Show that the definition of Poisson brackets on \(T^*M\) does not depend on the choice of the local coordinates on \(M\).

          @@ -3096,7 +3102,7 @@

          3.4 Poisson brack
        • - + Theorem 3.3. Poisson bracket defines a structure of infinite-dimensional Lie algebra5 on the space \(\mathcal {C}^\infty (T^*M)\) of smooth functions on the phase space.

          @@ -3187,7 +3193,7 @@

          3.4 Poisson brack

        - +

        5 The corresponding Lie group is the group of symplectomorphisms, a concept that will be introduced later. @@ -3246,7 +3252,7 @@

        3.4 Poisson brack
      • - + Theorem 3.4. Let \(H=H(q,p)\) be the hamiltonian of an hamiltonian system (3.31) and \(F=F(q,p)\in \mathcal {C}^\infty (T^*M)\). Then, along the flow \((q(t), p(t))\) of \(H\), we have

        @@ -3319,7 +3325,7 @@

        3.4 Poisson brack
      • - + Corollary 3.5. Given an hamiltonian system (3.31) with hamiltonian \(H\) and a smooth function \(F\) on its phase space, then \(F\) is a first integral if and only if \(H\) and \(F\) are in involution.

        @@ -3341,7 +3347,7 @@

        3.4 Poisson brack
      • - + Corollary 3.6. The first integrals of a hamiltonian system form a subalgebra of the Lie algebra \(\mathcal {C}^\infty (T^*M)\). Otherwise said, given two first integrals \(F\) and \(G\) of a hamiltonian system, then also the function \(\big \{F,G\big \}\) is a first integral of \(H\).

        @@ -3353,7 +3359,7 @@

        3.4 Poisson brack

      • - +
          @@ -3381,7 +3387,7 @@

          3.4 Poisson brack
        • - + Corollary 3.7. The hamiltonian function \(H\) is a first integral of the hamiltonian system (3.31).

          @@ -3403,7 +3409,7 @@

          3.4 Poisson brack
        • - + Corollary 3.8. The integral curves \(p = p(t)\) and \(q=q(t)\) of the hamiltonian system (3.31) belong to the levelset of the hamiltonian function, i.e.

          @@ -3441,8 +3447,8 @@

          3.4 Poisson brack -

          3.4.1 The symplectic matrix
          - +
          3.4.1 The symplectic matrix
          + @@ -3606,8 +3612,8 @@
          3.4.1 The symplec -
          3.4.2 A brief detour on time-dependent hamiltonians
          - +
          3.4.2 A brief detour on time-dependent hamiltonians
          + @@ -3775,8 +3781,8 @@
          3.4.2 A brief det -

          3.5 Variational principles of hamiltonian mechanics

          - +

          3.5 Variational principles of hamiltonian mechanics

          +

          @@ -3795,7 +3801,7 @@

          3.5 Variational p
        • - + Theorem 3.9. The solutions of a hamiltonian system (3.31) with hamiltonian \(H=H(q,p)\) are the critical points of the functional

          @@ -3837,7 +3843,7 @@

          3.5 Variational p

          - +
            @@ -3931,7 +3937,7 @@

            3.5 Variational p
          • - + Theorem 3.10. Given a function \(H=H(q,p)\) and a real number \(E\), then the (isoenergetic) trajectories of the hamiltonian system with hamiltonian \(H\) are the critical points of the truncated action

            @@ -3998,7 +4004,7 @@

            3.5 Variational p
          • - + Remark 3.5. Maupertuis principle is a variational principle in configuration space: it determines only the trajectories of the motion, not the dynamics on the trajectories.

            @@ -4014,7 +4020,7 @@

            3.5 Variational p

          - +
            @@ -4144,7 +4150,7 @@

            3.5 Variational p
          • - + Example 3.7 (Geodesics and the free motion on riemannian manifolds). Consider a riemannian manifold \((M, g)\), remember that the metric can be written as

            @@ -4305,7 +4311,7 @@

            3.5 Variational p
          • - + Theorem 3.11. Given a riemannian manifold \((M,g)\), the trajectories of a free point particle on the manifold are the geodesics of the metric.

            @@ -4337,7 +4343,7 @@

            3.5 Variational p
          • - + Example 3.8 (Jacobi metric). Let’s consider a mechanical system \((M,L)\), where \(L\) is a natural lagrangian for a particle of mass \(m\),

            @@ -4395,7 +4401,7 @@

            3.5 Variational p
          • - + Theorem 3.12. The trajectories of the mechanical system \((M,L)\) in the domain (3.117) are the geodesic of the metric

            @@ -4432,7 +4438,7 @@

            3.5 Variational p

          - +
            @@ -4544,7 +4550,7 @@

            3.5 Variational p
          • - + Example 3.9 (Fermat principle and geometric optics). Propagation of light in an isotropic medium is described by the hamiltonian

            @@ -4660,7 +4666,7 @@

            3.5 Variational p
          • - + Theorem 3.13. Light in an isotropic medium propagates in a way that minimizes the travel time. The trajectories followed by the light are the geodesics of the metric

            @@ -4734,8 +4740,8 @@

            3.5 Variational p -

            3.6 Canonical transformations

            - +

            3.6 Canonical transformations

            +

            @@ -4857,7 +4863,7 @@

            3.6 Canonical tra
          • - + Example 3.10. The transformation \(\Phi : (q,p) \mapsto (p, -q)\) is a canonical transformation. To show it, it is enough to verify how the Poisson bracket acts on the coordinate functions. Let, as usual,

            @@ -4935,7 +4941,7 @@

            3.6 Canonical tra
          • - + Example 3.11. The cotangent lift \(\Phi \) of a diffeomorphism \(\phi \) defined on the base manifold \(M\), as defined below, is a canonical transformation.

            @@ -5011,7 +5017,7 @@

            3.6 Canonical tra
          • - + Exercise 3.5. Consider the linear transformations on \(\mathbb {R}^{2n}\),

            @@ -5080,7 +5086,7 @@

            3.6 Canonical tra
          • - + Theorem 3.14. Let \(\Phi : T^*M \to T^* M\) be a canonical transformation. Then, every solution \(x = x(t)\) of the hamiltonian system with hamiltonian \(H\) is mapped by \(\Phi \) into a solution \(\Phi (x(t))\) of the hamiltonian system with hamiltonian \((\Phi ^{-1})^* H\).

            @@ -5092,7 +5098,7 @@

            3.6 Canonical tra

          - +
            @@ -5233,8 +5239,8 @@

            3.6 Canonical tra -

            3.6.1 Hamiltonian vector fields
            - +
            3.6.1 Hamiltonian vector fields
            +

            @@ -5293,7 +5299,7 @@

            3.6.1 Hamiltonian
          • - + Exercise 3.6. Let \(\Phi _t:T^*M \to T^*M\) denote a one–parameter group of canonical transformations, i.e \(\big \{\Phi _t^* f, \Phi _t^* g\big \} = \Phi _t^*\big \{f,g\big \}\) for all \(t\in \mathbb {R}\). Show that the vector field \(X\) which generates the group,

            @@ -5472,7 +5478,7 @@
            3.6.1 Hamiltonian

            - +

            6 Don’t let the \(T\)s scare you, here \(TT^*M\) simply means the tangent bundle to the manifold \(T^*M\). @@ -5530,7 +5536,7 @@

            3.6.1 Hamiltonian
            - +
              @@ -5614,7 +5620,7 @@
              3.6.1 Hamiltonian
            • - + Theorem 3.16. The hamiltonian vector field \(X_H\) associated to a smooth function \(H\) on \(T^*M\) is an infinitesimal symmetry of Poisson bracket. Vice versa, for every infinitesimal symmetry \(X\) there exists a function \(H\) on \(T^*M\) such that \(X = X_H\).

              @@ -5626,7 +5632,7 @@
              3.6.1 Hamiltonian
            - +
              @@ -5916,7 +5922,7 @@
              3.6.1 Hamiltonian
            • - + Exercise 3.7. Let \(\Phi _t\) the one–parameter group of canonical transformations generated by the quadratic hamiltonian

              @@ -5984,7 +5990,7 @@
              3.6.1 Hamiltonian
            • - + Exercise 3.8. Let \(\phi _t:M\to M\) denote a one–parameter group of diffeomorphisms of the base \(M\) of the cotangent bundle \(T^*M\). Consider, as in Example 3.11, its cotangent lift

              @@ -6051,8 +6057,8 @@
              3.6.1 Hamiltonian -
              3.6.2 The hamiltonian Noether theorem
              - +
              3.6.2 The hamiltonian Noether theorem
              +

              @@ -6078,7 +6084,7 @@

              3.6.2 The hamilto
            - +

            7 This should be compared with a symmetry of a lagrangian \(L:TM \to \mathbb {R}\), i.e. a diffeomorphism \(\phi : M \to M\) such that \(\phi ^* L = L\). Note that even though the formula looks very similar, the symmetry of the hamiltonian is defined on the whole phase space and not induced by a symmetry on the base as in the Lagrangian case. In the hamiltonian formalism, position and momentum are really on equal footings. @@ -6104,7 +6110,7 @@

            3.6.2 The hamilto
          • - + Theorem 3.17 (Hamiltonian Noether theorem). Let \(H\) be a hamiltonian on \(T^*M\) and \(F\) be a first integral of the hamiltonian system, then \(F\) generates a one–parameter group of symmetries of the hamiltonian system \(H\).

            @@ -6120,7 +6126,7 @@
            3.6.2 The hamilto
          • - +
              @@ -6139,7 +6145,7 @@
              3.6.2 The hamilto
            • - + Corollary 3.18 (of Theorem 3.15). Let \(H, F \in \mathcal {C}^\infty (T^*M)\) be two hamiltonian functions in involution, i.e., \(\{H,F\}=0\). Then, their flows, respectively denoted here by \(\Phi _t^H\) and \(\Phi _s^F\), commute: @@ -6195,7 +6201,7 @@

              3.6.2 The hamilto
            • - + Example 3.12. Consider \((q,p) = (x,y,z,p_x,p_y,p_z)\in T^* \mathbb {R}^3\) and the infinitesimal generator \(M_z = x p_y - y p_x\) (which you may recognize as the \(z\)-component of the angular momentum).

              @@ -6386,7 +6392,7 @@
              3.6.2 The hamilto
            • - + Exercise 3.9. Let \(X(q), Y(q)\) denote two vector fields on \(M\) and define two hamiltonians linear in the momenta as in (3.177):

              @@ -6454,7 +6460,7 @@
              3.6.2 The hamilto
            • - + Exercise 3.10. Given a mechanical system which is invariant with respect to space translations, show that the components of the total momentum

              @@ -6522,8 +6528,8 @@
              3.6.2 The hamilto -

              3.7 The symplectic structure on the cotangent bundle

              - +

              3.7 The symplectic structure on the cotangent bundle

              +

              @@ -6677,7 +6683,7 @@

              3.7 The symplecti

            - +

            8 I.e. \(\dd \omega = 0\)

            @@ -6698,7 +6704,7 @@

            3.7 The symplecti
          • - + Remark 3.6. The fact that \(\omega \) is closed and non-degenerate, in coordinates, can be translated into

            @@ -6740,7 +6746,7 @@

            3.7 The symplecti
          • - + Remark 3.7. Note that for an \(n\times n\) antisymmetric matrix \(A = - A^T\),

            @@ -6806,7 +6812,7 @@

            3.7 The symplecti
          • - + Lemma 3.19. The matrix of the two–form \(\omega \) defined by (3.201) in the coordinates \(q,p\) is given by

            @@ -6846,7 +6852,7 @@

            3.7 The symplecti

          • - +
              @@ -6950,7 +6956,7 @@

              3.7 The symplecti
            • - + Theorem 3.20. The two–form \(\omega \) defined by (3.201) does not depend from the local choice of coordinates \(q^1, \ldots , q^n\) on the base \(M\) of the cotangent bundle \(T^* M\). This two–form is closed and non–degenerate, therefore defining a symplectic structure on the manifold \(T^*M\). @@ -6973,7 +6979,7 @@

              3.7 The symplecti
            • - + Lemma 3.21. Let \(q^1, \ldots , q^n\) be a choice of local coordinates on a smooth manifold \(M\). Then the tautological one–form

              @@ -7035,7 +7041,7 @@

              3.7 The symplecti

            - +
              @@ -7104,7 +7110,7 @@

              3.7 The symplecti

            - +
              @@ -7133,7 +7139,7 @@

              3.7 The symplecti
            • - + Remark 3.8 (A coordinate free explanation of the tautological one–form). The name tautological comes from the physical intuition that velocity and momenta are necessarily proportional to one-another, but the connection between the two concepts runs deep as previous remarks have shown. The tautological one-form enter into this picture as a foundational concept. @@ -7352,7 +7358,7 @@

              3.7 The symplecti
            • - + Exercise 3.11. Show that \(\big \{f,g\big \} = -\mathcal {L}_{X_f} g = \mathcal {L}_{X_g} f\), where \(\mathcal {L}\) is the Lie derivative defined in (3.19). Remember that the inner product of a vector field with a function is zero by definition.

              @@ -7370,7 +7376,7 @@

              3.7 The symplecti
            • - + Exercise 3.12. Let \(X_H\) be an hamiltonian vector field on \(T^*M\). Show that the following holds

              @@ -7442,8 +7448,8 @@

              3.7 The symplecti -

              3.7.1 Symplectomorphisms and generating functions
              - +
              3.7.1 Symplectomorphisms and generating functions
              +

              @@ -7494,7 +7500,7 @@

              3.7.1 Symplectomo
            • - + Theorem 3.22. Let \(\Phi : T^*M \to T^*M\) be a diffeomorphism, then \(\Phi \) is a canonical transformation if and only if it is a symplectomorphism.

              @@ -7506,7 +7512,7 @@
              3.7.1 Symplectomo
            - +
              @@ -7697,7 +7703,7 @@
              3.7.1 Symplectomo
            • - + Theorem 3.23. Let \(\Phi : T^*M \to T^*M\) be a symplectomorphism, then locally there exists a function \(S:T^*M\to \mathbb {R}\) such that

              @@ -7734,7 +7740,7 @@
              3.7.1 Symplectomo
            - +
              @@ -7794,7 +7800,7 @@
              3.7.1 Symplectomo
            • - + Theorem 3.24. Let \(H\) be a function on the manifold \(T^*M\), then the hamiltonian flow \(\Phi _t\) generated by \(H\) is a one–parameter group of symplectomorphisms of \((T^*M, \omega )\).

              @@ -7908,7 +7914,7 @@
              3.7.1 Symplectomo
            • - + Theorem 3.25. Let \(X\) be an infinitesimal symplectomorphism on the symplectic manifold \((P,\omega )\), then locally there exists a function \(H\) such that \(X = X_H\).

              @@ -7926,7 +7932,7 @@
              3.7.1 Symplectomo
            • - + Example 3.13 (Time dependent hamiltonians). It is possible to use the symplectic formulation to describe hamiltonian systems with an explicit time dependence. To this end, one needs to consider the extended phase space \(T^*M\times \mathbb {R}^2\), as we have already seen in Section 3.4.2. @@ -7993,8 +7999,8 @@

              3.7.1 Symplectomo -
              3.7.2 Darboux Theorem
              - +
              3.7.2 Darboux Theorem
              +

              @@ -8013,7 +8019,7 @@

              3.7.2 Darboux The
            • - + Theorem 3.26 (Darboux’ Theorem). Let \((P, \omega )\) be a symplectic manifold and \(\dim P = 2n\). Then, there exist local coordinates

              @@ -8049,7 +8055,7 @@
              3.7.2 Darboux The
            - +
              @@ -8154,7 +8160,7 @@
              3.7.2 Darboux The
            • - + Lemma 3.27. The symplectic structure \(\omega \) in \(\mathbb {R}^{2n}\) induces a symplectic structure in a neighborhood \(V_{x_0}\subset S\).

              @@ -8166,7 +8172,7 @@
              3.7.2 Darboux The
            - +
              @@ -8372,8 +8378,8 @@
              3.7.2 Darboux The -
              3.7.3 Liouville theorem
              - +
              3.7.3 Liouville theorem
              +

              @@ -8388,7 +8394,7 @@

              3.7.3 Liouville t
            • - + Lemma 3.28. On a manifold \(M\) of dimension \(n\), let \(\omega \) be the symplectic form (3.201) on \(T^* M\). Then the exterior product of \(n\) copies of \(\omega \) is proportional to the volume form:

              @@ -8426,7 +8432,7 @@
              3.7.3 Liouville t
            - +
              @@ -8503,7 +8509,7 @@
              3.7.3 Liouville t
            • - + Exercise 3.13. Show that the definition of volume \(\Vol (D)\) presented above does not depend on the choice of local coordinates \(q^1, \ldots , q^n\) on \(M\).

              @@ -8521,7 +8527,7 @@
              3.7.3 Liouville t
            • - + Theorem 3.29. Let \(\Phi :T^*M \to T^*M\) be a canonical transformation and \(D\subset T^*M\) measurable, then

              @@ -8558,7 +8564,7 @@
              3.7.3 Liouville t
            - +
            @@ -643,8 +649,8 @@

            Hamiltonian Mechanics

            -

            4 Integrable systems

            - +

            4 Integrable systems

            +

            @@ -654,8 +660,8 @@

            4 Integrable syst -

            4.1 Lagrangian submanifolds

            - +

            4.1 Lagrangian submanifolds

            + @@ -715,7 +721,7 @@

            4.1 Lagrangian su
          • - + Theorem 4.1. The submanifold \(\Lambda \) defined by (4.1) is lagrangian if and only if the functions \(f_i\) are in involution on \(\Lambda \), i.e.

            @@ -755,7 +761,7 @@

            4.1 Lagrangian su

          - +
            @@ -891,7 +897,7 @@

            4.1 Lagrangian su
          • - + Example 4.1. One of the simplest example of lagrangian submanifolds are

            @@ -962,7 +968,7 @@

            4.1 Lagrangian su
          • - + Theorem 4.2. The submanifold \(\Lambda \) is lagrangian if and only if the following equations are satisfied:

            @@ -1002,7 +1008,7 @@

            4.1 Lagrangian su

          - +
            @@ -1073,7 +1079,7 @@

            4.1 Lagrangian su
          • - + Theorem 4.3. Equations (4.12) define a lagrangian submanifold if and only if the following one–form on \(M\) is closed:

            @@ -1110,7 +1116,7 @@

            4.1 Lagrangian su

          - +
            @@ -1210,7 +1216,7 @@

            4.1 Lagrangian su
          • - + Remark 4.1. For a hamiltonian system with hamiltonian \(H\), we can compute the generating function of a lagrangian submanifold from \(n\) first integrals \(f_1, \ldots , f_n\) in involution such that

            @@ -1326,8 +1332,8 @@

            4.1 Lagrangian su -

            4.2 Canonical transformations revisited

            - +

            4.2 Canonical transformations revisited

            +

            @@ -1443,7 +1449,7 @@

            4.2 Canonical tra
          • - + Remark 4.2. If \(q^i = Q^i\), \(i=1,\ldots ,n\), then the function \(S=S(q)\). In fact, in this case, (4.24) reads

            @@ -1486,8 +1492,8 @@

            4.2 Canonical tra -

            4.2.1 The generating function \(S_1\)
            - +
            4.2.1 The generating function \(S_1\)
            +

            @@ -1567,7 +1573,7 @@

            4.2.1 The generat
          • - + Theorem 4.4. Let \(S_1(q,Q)\) be a function defined in a neighborhood of the point \((q_0, Q_0)\) of two euclidean coordinate spaces. If

            @@ -1605,7 +1611,7 @@
            4.2.1 The generat
          - +
            @@ -1728,7 +1734,7 @@
            4.2.1 The generat
          • - + Remark 4.3. Not all canonical transformations are free: the identical transformation is not free, as \(q=Q\) and thus they are not independent.

            @@ -1746,7 +1752,7 @@
            4.2.1 The generat
          • - + Example 4.2. Let \(\omega > 0\) be some fixed constant. Consider the transformation \(\Phi \) given by

            @@ -1949,8 +1955,8 @@
            4.2.1 The generat -
            4.2.2 The generating functions \(S_2\) and \(S_3\)
            - +
            4.2.2 The generating functions \(S_2\) and \(S_3\)
            +

            @@ -2047,7 +2053,7 @@

            4.2.2 The generat
          • - + Example 4.3. Let \(Q=Q(q)\) be a change of coordinates on the base \(M\) of the cotangent bundle \(T^*M\). This induces a transformation of the conjugate momenta \(p_i\mapsto P_i\) given by \(P_i = p_l \frac {\partial q^l}{\partial Q^i}\), and we already saw that \((q,p)\mapsto (Q,P)\) defines a canonical transformation. The generating function of such transformation is given by @@ -2119,7 +2125,7 @@

            4.2.2 The generat
          • - + Theorem 4.5. Let \(\Phi :T^*M \to T^* M\) be the canonical transformation defined by the functions \(P(q,p)\) and \(Q(q,p)\). Then, in a neighborhood of a point \((q_0, p_0)\) it is possible to choose one of the \(2^n\) collections of functions \((q, Q^{\vb * j}, P_{\vb * i})\) from (4.49) as independent coordinates, i.e., such that @@ -2241,7 +2247,7 @@

            4.2.2 The generat
          - +
            @@ -2292,7 +2298,7 @@
            4.2.2 The generat
          • - + Remark 4.4. Sometimes, in the literature you can find a different classification in the following terms:

              @@ -2344,8 +2350,8 @@
              4.2.2 The generat -
              4.2.3 Infinitesimal canonical transformations
              - +
              4.2.3 Infinitesimal canonical transformations
              +

              @@ -2384,7 +2390,7 @@

              4.2.3 Infinitesim
            • - + Theorem 4.6. Every canonical transformation which is close to the identity admits a generating function of the form

              @@ -2470,7 +2476,7 @@
              4.2.3 Infinitesim
            • - + Theorem 4.7. The change of coordinates (4.55) is an infinitesimal canonical transformation if and only if there exists a function \(K=K(q,p)\) such that

              @@ -2508,7 +2514,7 @@
              4.2.3 Infinitesim
          - +
            @@ -2661,8 +2667,8 @@
            4.2.3 Infinitesim -
            4.2.4 Time-dependent hamiltonian systems
            - +
            4.2.4 Time-dependent hamiltonian systems
            +

            @@ -2726,7 +2732,7 @@

            4.2.4 Time-depend
          • - + Theorem 4.8. Let \((q,p,t,E) \to (Q,P,T,\widetilde E)\) be a canonical transformation in \(T^*M\times \mathbb {R}^2\). If

            @@ -2762,7 +2768,7 @@
            4.2.4 Time-depend
          - +
            @@ -2913,7 +2919,7 @@
            4.2.4 Time-depend
          • - + Example 4.4. Let \(H(q,p) = \frac {p^2}{2} + \alpha q\). Consider a time-dependent canonical transformation \(\Phi _t\) generated by \(H\): let \(q(0) = Q\) and \(p(0) = P\) and

            @@ -3093,7 +3099,7 @@
            4.2.4 Time-depend
          • - + Theorem 4.9. The hamiltonian flow is a canonical transformation.

            @@ -3104,7 +3110,7 @@
            4.2.4 Time-depend
          - +
            @@ -3238,7 +3244,7 @@
            4.2.4 Time-depend
          • - + Remark 4.5. Choosing \((q,Q,t)\) as independent variables in the proof above one gets

            @@ -3277,8 +3283,8 @@
            4.2.4 Time-depend -

            4.3 Hamilton-Jacobi equations

            - +

            4.3 Hamilton-Jacobi equations

            +

            @@ -3561,7 +3567,7 @@

            4.3 Hamilton-Jaco
          • - + Theorem 4.10. Let \(H(q,p,t)\) be an hamiltonian function and \(S(q,b,t)\) be a complete integral of the Hamilton-Jacobi equation (4.95) depending on \(n\) constants \(b=(b_1,\ldots ,b_n)\). If \(S\) satisfies the non–degeneracy condition

            @@ -3615,7 +3621,7 @@

            4.3 Hamilton-Jaco
          • - + Theorem 4.11 (Liouville theorem). Let \(H(q,p)\) denote the hamiltonian of a hamiltonian system with \(n\) degrees of freedom. Assume that there exist \(n\) first integral \(f_i\) of \(H\) which are in involution, i.e.

            @@ -3707,8 +3713,8 @@

            4.3 Hamilton-Jaco -

            4.3.1 Separable hamiltonians
            - +
            4.3.1 Separable hamiltonians
            + @@ -3788,7 +3794,7 @@
            4.3.1 Separable h
          • - + Example 4.5. Cyclic coordinates are an immediate example of separable coordinates. Assume for example that \(q^1\) is cyclic, then the conjugate momentum \(p_1 = b_1\) is constant. The generating function has the form

            @@ -3901,7 +3907,7 @@
            4.3.1 Separable h
          • - + Example 4.6. A simple example of separable system is given by a sum of hamiltonians depending each on just one pair of conjugate coordinates:

            @@ -3947,7 +3953,7 @@
            4.3.1 Separable h
          • - + Example 4.7 (Planar harmonic oscillator). Let’s reconsider an old friend of ours, the planar harmonic oscillator defined by a point particle of mass \(m\) attracted by an elastic force centered at \(0\in \mathbb {R}^2\) with stiffness \(k\).

            @@ -4172,7 +4178,7 @@
            4.3.1 Separable h
          • - + Example 4.8. Let’s consider a more involved example of a point of mass \(m\) moving in \(\mathbb {R}^3\) under the influence of an arbitrary potential \(U\).

            @@ -4448,8 +4454,8 @@
            4.3.1 Separable h -
            4.3.2 A few exercises
            - +
            4.3.2 A few exercises
            +
              @@ -4458,7 +4464,7 @@
              4.3.2 A few exerc
            • - + Exercise 4.1. Show that the infinitesimal transformation

              @@ -4543,7 +4549,7 @@
              4.3.2 A few exerc
            • - + Exercise 4.2. Let \(Q = q^2 + \frac 12 \cos (q)\). Find \(P(q,p)\) such that the corresponding coordinate transformation is canonical. Compute the corresponding generating function.

              @@ -4561,7 +4567,7 @@
              4.3.2 A few exerc
            • - + Exercise 4.3 (Parabolic coordinates). Let \((r,\phi ,z)\) denote the cylindrical coordinates in \(\mathbb {R}^3\). Find the complete integral of the Hamilton-Jacobi equations for a point particle of mass \(m\) under the influence of the potential

              @@ -4632,7 +4638,7 @@
              4.3.2 A few exerc
            • - + Exercise 4.4 (Prolate spheroidal coordinates). Consider a point particle of mass \(m\) under the gravitational attraction of two bodies which are fixed at the points \((d,0,0)\) and \((-d,0,0)\), \(d>0\). Show that the Hamilton-Jacobi equation is separable introducing the prolate spheroidal coordinates obtained by @@ -4675,7 +4681,7 @@

              4.3.2 A few exerc
            • - + Exercise 4.5. Consider the motion of a free particle of mass \(m\) on a surface \(S\subset \mathbb {R}^3\) parametrized by

              @@ -4712,8 +4718,8 @@
              4.3.2 A few exerc -

              4.4 The Liouville-Arnold theorem

              - +

              4.4 The Liouville-Arnold theorem

              + @@ -4822,7 +4828,7 @@

              4.4 The Liouville
            • - + Remark 4.6. In the integral (4.148) you can choose the initial point arbitrarily as long as it varies smoothly with respect to \(E\).

              @@ -4995,7 +5001,7 @@

              4.4 The Liouville
            • - + Theorem 4.12 (Liouville-Arnold). Consider a completely integrable hamiltonian system on the phase space \(\mathbb {R}^{2n}\). Assume that there exists \(E^0 = (E^0_1, \ldots , E^0_n)\) such that the functions

              @@ -5113,7 +5119,7 @@

              4.4 The Liouville
            • - + Remark 4.7. A separable hamiltonian system is completely integrable. Indeed, for a separable system there exists a canonical transformation \((q,p) \mapsto (Q,P)\) such that the hamiltonian \(H\) in the new coordinates does not depend on \(Q\), \(H=H(P)\). The functions \(P_1, \ldots , P_n\), which are in involution by construction, commute with \(H\). @@ -5126,7 +5132,7 @@

              4.4 The Liouville

            - +
              @@ -5375,7 +5381,7 @@

              4.4 The Liouville
            • - + Lemma 4.13. Let \(M\) be a \(n\)-dimensional compact connected manifold. Assume that at all points \(x\in M\) there exist \(n\) linearly independent vector fields \(X_1, \ldots , X_n\) which are pairwise commuting, i.e.,

              @@ -5411,7 +5417,7 @@

              4.4 The Liouville

            - +
              @@ -5508,7 +5514,7 @@

              4.4 The Liouville
            • - + Exercise 4.6. Show that the action \(\Phi _{\vb * t}\) is transitive, that is, for every two points \(x, y\in M\), there exists \(\vb * t\in \mathbb {R}^n\) such that \(y = \Phi _{\vb * t} (x)\).

              @@ -5578,7 +5584,7 @@

              4.4 The Liouville
            • - + Exercise 4.7. Show that the subgroup \(\Gamma \) does not depend on \(x_0\). Hint: \(\Gamma \) is transitive. Furthermore, \(\Gamma \) is a discrete subgroup of \(\mathbb {R}^n\), that is, there is an open set \(U \ni \vb * 0\) such that \(U\cap \Gamma = \{\vb * 0\}\).

              @@ -5600,7 +5606,7 @@

              4.4 The Liouville
            • - + Lemma 4.14. Let \(\Gamma \) be a discrete subgroup of \(\mathbb {R}^n\). Then, there exist \(m\) linearly independent vectors \(e_1, \ldots , e_m \in \mathbb {R}^n\), \(0\leq m \leq n\), such that

              @@ -5812,7 +5818,7 @@

              4.4 The Liouville
            • - + Exercise 4.8. Extend the proof to completely integrable system on an arbitrary symplectic manifold. (Solution: find the differences between our proof and [Kna18, Theorem 13.3]).

              @@ -5825,8 +5831,8 @@

              4.4 The Liouville -

              4.4.1 Action-angle variables
              - +
              4.4.1 Action-angle variables
              +

              @@ -5871,7 +5877,7 @@

              4.4.1 Action-angl
            • - + Theorem 4.15. Under the hypotheses of the Liouville-Arnold theorem (Theorem 4.12), for every compact connected component \(M_{E^0}\) there are a neighborhood \(M_{E^0}\subseteq U = U(M_{E^0}) \subseteq \mathbb {R}^{2n}\) and a symplectomorphism

              @@ -5933,7 +5939,7 @@
              4.4.1 Action-angl
            - +
            @@ -643,8 +649,8 @@

            Hamiltonian Mechanics

            -

            5 Hamiltonian perturbation theory

            - +

            5 Hamiltonian perturbation theory

            +

            @@ -659,8 +665,8 @@

            5 Hamiltonian per -

            5.1 Small oscillations revisited

            - +

            5.1 Small oscillations revisited

            +

            @@ -856,7 +862,7 @@

            5.1 Small oscilla
          • - + Lemma 5.1. Let \(A\) be a symmetric matrix and \(J\) be the symplectic matrix. The characteristic polynomial of the matrix \(B=JA\) has the form

            @@ -892,7 +898,7 @@

            5.1 Small oscilla

          - +
          @@ -643,8 +649,8 @@

          Hamiltonian Mechanics

          -

          6 Conclusion

          - +

          6 Conclusion

          +

          @@ -660,8 +666,8 @@

          6 Conclusion

          -

          Bibliography

          - +

          Bibliography

          +
            @@ -792,6 +798,16 @@

            Bibliography

          • +

            +[Eur24] European Mathematical Society. “Elliptic Integral”. In: Encyclopedia of Mathematics. EMS Press, 2024. url: http://encyclopediaofmath.org/index.php?title=Elliptic_integral&oldid=46813. +

            + + +
          • +
          • + +

            [GF00] I.M. Gelfand and S.V. Fomin. Calculus of Variations. Dover Books on Mathematics. Dover Publications, New York, 2000. isbn: 9780486414485. @@ -972,6 +988,224 @@

            Bibliography

          + +

          Index

          + + + +

          +A +

          + +

          +acceleration, section 1.1 +

          + +

          +action, section 1.2 +

          + +

          +action +

          + +

          +critical point, section 1.2 +

          + +

          +B +

          + +

          +Banach space, section 1.2 +

          + +

          +C +

          + +

          +configuration, section 1.1.1 +

          + +

          +space, section 1.1.1 +

          + +

          +D +

          + +

          +degrees of freedom, section 1.1.1 +

          + +

          +diffemorphic, example 1.3 +

          + +

          +E +

          + +

          +elliptic integrals, example 1.1 +

          + +

          +equations of motion, section 1.1.1 +

          + +

          +Euler-Lagrange equations, theorem 1.1 +

          + +

          +F +

          + +

          +frames of reference, section 1.1 +

          + +

          +Frechet +

          + +

          +derivative, section 1.2 +

          + +

          +differentiable, section 1.2 +

          + +

          +G +

          + +

          +generalized +

          + +

          +coordinates, section 1.1.1 +

          + +

          +velocities, section 1.1.1 +

          + +

          +gravitational constant, example 1.2 +

          + +

          +H +

          + +

          +Hamilton’s principle, section 1.2 +

          + +

          +I +

          + +

          +inertia, section 1.1 +

          + +

          +L +

          + +

          +lagrangian, section 1.2 +

          + +

          +N +

          + +

          +Newton +

          + +

          +principle of determinacy, section 1.1.1 +

          + +

          +second law, section 1.1 +

          + +

          +universal gravitation, example 1.2 +

          + +

          +P +

          + +

          +pendulum, example 1.1 +

          + +

          +ideal, example 1.1 +

          + +

          +point particle, section 1.1 +

          + +

          +system, section 1.1.1 +

          + +

          +position, section 1.1 +

          + +

          +principle of least action, section 1.2 +

          + +

          +S +

          + +

          +spring, example 1.1 +

          + +

          +constant, example 1.1 +

          + +

          +Hooke’s law, example 1.1 +

          + +

          +state space, section 1.1.1 +

          + +

          +V +

          + +

          +variation, item 5 +

          + +

          +velocity, section 1.1 +

          diff --git a/index.html b/index.html index 1f3343b..bf70c63 100644 --- a/index.html +++ b/index.html @@ -472,290 +472,295 @@

          Contents

          - -1.1.1 Motion in one degree of freedom + +1.1.1 From particles to systems of particles

          - + 1.2 Hamilton’s variational principle

          - + 1.2.1 Dynamics of point particles: from Lagrange back to Newton

          - + 1.3 First steps with conserved quantities

          - + 1.3.1 Back to one degree of freedom

          - + 1.3.2 The conservation of energy

          - + 1.3.3 Fun with the phase portrait

          - + 1.4 Euler-Lagrange equations on smooth manifolds

          - + 2 Conservation Laws

          - + 2.1 Noether theorem

          - + 2.1.1 Homogeneity of space: total momentum

          - + 2.1.2 Isotropy of space: angular momentum

          - + 2.1.3 Homogeneity of time: the energy

          - + 2.1.4 Scale invariance: Kepler’s third law

          - + 2.2 The spherical pendulum

          - + 2.3 Intermezzo: small oscillations

          - + 2.3.1 Decomposition into characteristic oscillations

          - + 2.3.2 The double pendulum

          - + 2.3.3 The linear triatomic molecule

          - + 2.3.4 Zero modes

          - + 2.4 Motion in a central potential

          - + 2.4.1 Kepler’s problem

          - + 2.5 D’Alembert principle and systems with constraints

          - + 3 Hamiltonian mechanics

          - + 3.1 The Legendre Transform in the euclidean plane

          - + 3.2 Intermezzo: cotangent bundle and differential forms

          - + 3.3 The Legendre Transform

          - + 3.4 Poisson brackets and first integrals of motion

          - + 3.4.1 The symplectic matrix

          - + 3.4.2 A brief detour on time-dependent hamiltonians

          - + 3.5 Variational principles of hamiltonian mechanics

          - + 3.6 Canonical transformations

          - + 3.6.1 Hamiltonian vector fields

          - + 3.6.2 The hamiltonian Noether theorem

          - + 3.7 The symplectic structure on the cotangent bundle

          - + 3.7.1 Symplectomorphisms and generating functions

          - + 3.7.2 Darboux Theorem

          - + 3.7.3 Liouville theorem

          - + 4 Integrable systems

          - + 4.1 Lagrangian submanifolds

          - + 4.2 Canonical transformations revisited

          - + 4.2.1 The generating function \(S_1\)

          - + 4.2.2 The generating functions \(S_2\) and \(S_3\)

          - + 4.2.3 Infinitesimal canonical transformations

          - + 4.2.4 Time-dependent hamiltonian systems

          - + 4.3 Hamilton-Jacobi equations

          - + 4.3.1 Separable hamiltonians

          - + 4.3.2 A few exercises

          - + 4.4 The Liouville-Arnold theorem

          - + 4.4.1 Action-angle variables

          - + 5 Hamiltonian perturbation theory

          - + 5.1 Small oscillations revisited

          - + 5.2 Birkhoff normal forms

          - + 5.3 A brief look at KAM theory

          - + 5.4 Nekhoroshev theorem

          - + 6 Conclusion

          - + Bibliography

          +

          + +Index +

          +