律证论坛
Would you like to react to this message? Create an account in a few clicks or log in to continue.

Quantum Field Theory II

4页/共4 上一页  1, 2, 3, 4

向下

Quantum Field Theory II - 页 4 Empty 回复: Quantum Field Theory II

帖子 由 一星 2014-08-30, 01:12

Fourier analysis[edit]


Quantum Field Theory II - 页 4 220px-Sawtooth_Fourier_Analysys.svg

Superposition of sinusoidal wave basis functions (bottom) to form a sawtooth wave (top)


Quantum Field Theory II - 页 4 220px-Harmoniki

Spherical harmonics, an orthonormal basis for the Hilbert space of square-integrable functions on the sphere, shown graphed along the radial direction


One of the basic goals of Fourier analysis is to decompose a function into a (possibly infinite) linear combination of given basis functions: the associated Fourier series. The classical Fourier series associated to a function f defined on the interval [0, 1] is a series of the form
Quantum Field Theory II - 页 4 0f7eb7ee13ebd39ba8fccffe3ed435ad
where
Quantum Field Theory II - 页 4 659ea7061f3269a36cea1cfda44989b4
The example of adding up the first few terms in a Fourier series for a sawtooth function is shown in the figure. The basis functions are sine waves with wavelengths λ/n (n=integer) shorter than the wavelength λ of the sawtooth itself (except for n=1, the fundamental wave). All basis functions have nodes at the nodes of the sawtooth, but all but the fundamental have additional nodes. The oscillation of the summed terms about the sawtooth is called the Gibbs phenomenon.

A significant problem in classical Fourier series asks in what sense the Fourier series converges, if at all, to the function f. Hilbert space methods provide one possible answer to this question.[33] The functions en(θ) = e2πinθ form an orthogonal basis of the Hilbert spaceL2([0,1]). Consequently, any square-integrable function can be expressed as a series
Quantum Field Theory II - 页 4 96fc184f136b95d3d2eb02f13a25911b
and, moreover, this series converges in the Hilbert space sense (that is, in the L2 mean).

The problem can also be studied from the abstract point of view: every Hilbert space has an orthonormal basis, and every element of the Hilbert space can be written in a unique way as a sum of multiples of these basis elements. The coefficients appearing on these basis elements are sometimes known abstractly as the Fourier coefficients of the element of the space.[34] The abstraction is especially useful when it is more natural to use different basis functions for a space such as L2([0,1]). In many circumstances, it is desirable not to decompose a function into trigonometric functions, but rather into orthogonal polynomials or wavelets for instance,[35] and in higher dimensions into spherical harmonics.[36]

For instance, if en are any orthonormal basis functions of L2[0,1], then a given function in L2[0,1] can be approximated as a finite linear combination[37]
Quantum Field Theory II - 页 4 5262d8d68b09bf4374eec79f80796cc8
The coefficients {aj} are selected to make the magnitude of the difference ||ƒ − ƒn||2 as small as possible. Geometrically, the best approximation is the orthogonal projection of ƒonto the subspace consisting of all linear combinations of the {ej}, and can be calculated by[38]
Quantum Field Theory II - 页 4 2d38620bc5b1bf008c80c88ed7018baf
That this formula minimizes the difference ||ƒ − ƒn||2 is a consequence of Bessel's inequality and Parseval's formula.

In various applications to physical problems, a function can be decomposed into physically meaningful eigenfunctions of a differential operator (typically the Laplace operator): this forms the foundation for the spectral study of functions, in reference to the spectrum of the differential operator.[39] A concrete physical application involves the problem of hearing the shape of a drum: given the fundamental modes of vibration that a drumhead is capable of producing, can one infer the shape of the drum itself?[40] The mathematical formulation of this question involves the Dirichlet eigenvalues of the Laplace equation in the plane, that represent the fundamental modes of vibration in direct analogy with the integers that represent the fundamental modes of vibration of the violin string.

Spectral theory also underlies certain aspects of the Fourier transform of a function. Whereas Fourier analysis decomposes a function defined on a compact set into the discrete spectrum of the Laplacian (which corresponds to the vibrations of a violin string or drum), the Fourier transform of a function is the decomposition of a function defined on all of Euclidean space into its components in the continuous spectrum of the Laplacian. The Fourier transformation is also geometrical, in a sense made precise by the Plancherel theorem, that asserts that it is an isometry of one Hilbert space (the "time domain") with another (the "frequency domain"). This isometry property of the Fourier transformation is a recurring theme in abstract harmonic analysis, as evidenced for instance by the Plancherel theorem for spherical functions occurring in noncommutative harmonic analysis.

Quantum mechanics[edit]


Quantum Field Theory II - 页 4 220px-HAtomOrbitals

The orbitals of an electron in ahydrogen atom are eigenfunctions of the energy.


In the mathematically rigorous formulation of quantum mechanics, developed by John von Neumann,[41] the possible states (more precisely, the pure states) of a quantum mechanical system are represented by unit vectors (called state vectors) residing in a complex separable Hilbert space, known as the state space, well defined up to a complex number of norm 1 (the phase factor). In other words, the possible states are points in the projectivization of a Hilbert space, usually called the complex projective space. The exact nature of this Hilbert space is dependent on the system; for example, the position and momentum states for a single non-relativistic spin zero particle is the space of all square-integrable functions, while the states for the spin of a single proton are unit elements of the two-dimensional complex Hilbert space of spinors. Each observable is represented by a self-adjoint linear operator acting on the state space. Each eigenstate of an observable corresponds to an eigenvector of the operator, and the associated eigenvalue corresponds to the value of the observable in that eigenstate.

The time evolution of a quantum state is described by the Schrödinger equation, in which the Hamiltonian, the operator corresponding to the total energy of the system, generates time evolution.

The inner product between two state vectors is a complex number known as a probability amplitude. During an ideal measurement of a quantum mechanical system, the probability that a system collapses from a given initial state to a particular eigenstate is given by the square of the absolute value of the probability amplitudes between the initial and final states. The possible results of a measurement are the eigenvalues of the operator—which explains the choice of self-adjoint operators, for all the eigenvalues must be real. The probability distribution of an observable in a given state can be found by computing the spectral decomposition of the corresponding operator.

For a general system, states are typically not pure, but instead are represented as statistical mixtures of pure states, or mixed states, given by density matrices: self-adjoint operators of trace one on a Hilbert space. Moreover, for general quantum mechanical systems, the effects of a single measurement can influence other parts of a system in a manner that is described instead by a positive operator valued measure. Thus the structure both of the states and observables in the general theory is considerably more complicated than the idealization for pure states.

Heisenberg's uncertainty principle is represented by the statement that the operators corresponding to certain observables do not commute, and gives a specific form that thecommutator must have.

Properties[edit]


Pythagorean identity[edit]


Two vectors u and v in a Hilbert space H are orthogonal when Quantum Field Theory II - 页 4 8a947f53b980051653669e9dc792e863 = 0. The notation for this is u ⊥ v. More generally, when S is a subset in H, the notation u ⊥ S means that u is orthogonal to every element from S.
When u and v are orthogonal, one has

Quantum Field Theory II - 页 4 A9123b3d7037490ef5d5a524a4469bf1
By induction on n, this is extended to any family u1,...,un of n orthogonal vectors,
Quantum Field Theory II - 页 4 85348b1479060033850839d482849ecc
Whereas the Pythagorean identity as stated is valid in any inner product space, completeness is required for the extension of the Pythagorean identity to series. A series Σ uk oforthogonal vectors converges in H  if and only if the series of squares of norms converges, and
Quantum Field Theory II - 页 4 Fa3922c4d5f0e697b45efc9af4375b44
Furthermore, the sum of a series of orthogonal vectors is independent of the order in which it is taken.

Parallelogram identity and polarization[edit]


Quantum Field Theory II - 页 4 220px-Color_parallelogram.svg

Geometrically, the parallelogram identity asserts thatAC2 + BD2 = 2(AB2 + AD2). In words, the sum of the squares of the diagonals is twice the sum of the squares of any two adjacent sides.


By definition, every Hilbert space is also a Banach space. Furthermore, in every Hilbert space the following parallelogram identity holds:
Quantum Field Theory II - 页 4 B62e1f16f79fe1cd45d0a93632f1c40c
Conversely, every Banach space in which the parallelogram identity holds is a Hilbert space, and the inner product is uniquely determined by the norm by the polarization identity.[42] For real Hilbert spaces, the polarization identity is
Quantum Field Theory II - 页 4 463b94bba2eac1faa46c473373ff177d
For complex Hilbert spaces, it is
Quantum Field Theory II - 页 4 D94c1993850ddf0d2fc85e33a1c7cae3
The parallelogram law implies that any Hilbert space is a uniformly convex Banach space.[43]

Best approximation[edit]


If C is a non-empty closed convex subset of a Hilbert space H and x a point in H, there exists a unique point y ∈ C that minimizes the distance between x and points in C,[44]
Quantum Field Theory II - 页 4 D4cd0d054349c66733bed673feaba146
This is equivalent to saying that there is a point with minimal norm in the translated convex set D = C − x. The proof consists in showing that every minimizing sequence (dn) ⊂ D is Cauchy (using the parallelogram identity) hence converges (using completeness) to a point in D that has minimal norm. More generally, this holds in any uniformly convex Banach space.[45]

When this result is applied to a closed subspace F of H, it can be shown that the point y ∈ F closest to x is characterized by[46]
Quantum Field Theory II - 页 4 B240ecc51b769fc3f6f6916efdf90d4f
This point y is the orthogonal projection of x onto F, and the mapping PF : x → y is linear (see Orthogonal complements and projections). This result is especially significant inapplied mathematics, especially numerical analysis, where it forms the basis of least squares methods[citation needed].

In particular, when F is not equal to H, one can find a non-zero vector v orthogonal to F (select x not in F and v = x − y). A very useful criterion is obtained by applying this observation to the closed subspace F generated by a subset S of H.
A subset S of H spans a dense vector subspace if (and only if) the vector 0 is the sole vector v ∈ H orthogonal to S.

Duality[edit]


The dual space H* is the space of all continuous linear functions from the space H into the base field. It carries a natural norm, defined by
Quantum Field Theory II - 页 4 0d1e1a4074dace1be9ded5082a3b2407
This norm satisfies the parallelogram law, and so the dual space is also an inner product space. The dual space is also complete, and so it is a Hilbert space in its own right.

The Riesz representation theorem affords a convenient description of the dual. To every element u of H, there is a unique element φu of H*, defined by
Quantum Field Theory II - 页 4 9d4c25462a0f6849a9a29315bd1c49c4
The mapping Quantum Field Theory II - 页 4 A702b190939a6a81dc161754217a23d7 is an antilinear mapping from H to H*. The Riesz representation theorem states that this mapping is an antilinear isomorphism.[47] Thus to every element φof the dual H* there exists one and only one uφ in H such that
Quantum Field Theory II - 页 4 De99d884956dc3ea8c83477451b5811e
for all x ∈ H. The inner product on the dual space H* satisfies
Quantum Field Theory II - 页 4 Ddb4fade458944dc6938d1ec0da7e3c2
The reversal of order on the right-hand side restores linearity in φ from the antilinearity of uφ. In the real case, the antilinear isomorphism from H to its dual is actually an isomorphism, and so real Hilbert spaces are naturally isomorphic to their own duals.

The representing vector uφ is obtained in the following way. When φ ≠ 0, the kernel F = Ker(φ) is a closed vector subspace of H, not equal to H, hence there exists a non-zero vector v orthogonal to F. The vector u is a suitable scalar multiple λv of v. The requirement that φ(v) = ⟨vu⟩ yields
Quantum Field Theory II - 页 4 9aed3f0bedf8f6467624a10ce4fc9109
This correspondence φ ↔ u is exploited by the bra–ket notation popular in physics. It is common in physics to assume that the inner product, denoted by ⟨x|y⟩, is linear on the right,
Quantum Field Theory II - 页 4 D14a3f67145af7482ed734b36ceee44e
The result ⟨x|y⟩ can be seen as the action of the linear functional ⟨x| (the bra) on the vector  |y⟩ (the ket).

The Riesz representation theorem relies fundamentally not just on the presence of an inner product, but also on the completeness of the space. In fact, the theorem implies that the topological dual of any inner product space can be identified with its completion. An immediate consequence of the Riesz representation theorem is also that a Hilbert space His reflexive, meaning that the natural map from H into its double dual space is an isomorphism.

Weakly convergent sequences[edit]


Main article: Weak convergence (Hilbert space)

In a Hilbert space H, a sequence {xn} is weakly convergent to a vector x ∈ H when
Quantum Field Theory II - 页 4 31c5f8041bdc0c56197e49e8b4b3c9fa
for every v ∈ H.

For example, any orthonormal sequence {fn} converges weakly to 0, as a consequence of Bessel's inequality. Every weakly convergent sequence {xn} is bounded, by the uniform boundedness principle.

Conversely, every bounded sequence in a Hilbert space admits weakly convergent subsequences (Alaoglu's theorem).[48] This fact may be used to prove minimization results for continuous convex functionals, in the same way that the Bolzano–Weierstrass theorem is used for continuous functions on Rd. Among several variants, one simple statement is as follows:[49]
If f: H → R is a convex continuous function such that f(x) tends to +∞ when ||x|| tends to ∞, then f admits a minimum at some point x0 ∈ H.
This fact (and its various generalizations) are fundamental for direct methods in the calculus of variations. Minimization results for convex functionals are also a direct consequence of the slightly more abstract fact that closed bounded convex subsets in a Hilbert space H are weakly compact, since H is reflexive. The existence of weakly convergent subsequences is a special case of the Eberlein–Šmulian theorem.

Banach space properties[edit]


Any general property of Banach spaces continues to hold for Hilbert spaces. The open mapping theorem states that a continuous surjective linear transformation from one Banach space to another is an open mapping meaning that it sends open sets to open sets. A corollary is the bounded inverse theorem, that a continuous and bijective linear function from one Banach space to another is an isomorphism (that is, a continuous linear map whose inverse is also continuous). This theorem is considerably simpler to prove in the case of Hilbert spaces than in general Banach spaces.[50] The open mapping theorem is equivalent to the closed graph theorem, which asserts that a function from one Banach space to another is continuous if and only if its graph is a closed set.[51] In the case of Hilbert spaces, this is basic in the study of unbounded operators (see closed operator).

The (geometrical) Hahn–Banach theorem asserts that a closed convex set can be separated from any point outside it by means of a hyperplane of the Hilbert space. This is an immediate consequence of the best approximation property: if y is the element of a closed convex set F closest to x, then the separating hyperplane is the plane perpendicular to the segment xy passing through its midpoint.[52]

Operators on Hilbert spaces[edit]


Bounded operators[edit]


The continuous linear operators A : H1 → H2 from a Hilbert space H1 to a second Hilbert space H2 are bounded in the sense that they map bounded sets to bounded sets. Conversely, if an operator is bounded, then it is continuous. The space of such bounded linear operators has a norm, the operator norm given by
Quantum Field Theory II - 页 4 67764b2fc29d57418c46c8819274ab0a
The sum and the composite of two bounded linear operators is again bounded and linear. For y in H2, the map that sends x ∈ H1 to ⟨Axy⟩ is linear and continuous, and according to the Riesz representation theorem can therefore be represented in the form
Quantum Field Theory II - 页 4 A2fbdad640218afa757a34ea95aa9b2b
for some vector A*y in H1. This defines another bounded linear operator A*H2 → H1, the adjoint of A. One can see that A** = A.

The set B(H) of all bounded linear operators on H, together with the addition and composition operations, the norm and the adjoint operation, is a C*-algebra, which is a type ofoperator algebra.

An element A  of B(H) is called self-adjoint or Hermitian if A*A. If A  is Hermitian and ⟨Axx⟩ ≥ 0 for every x, then A  is called non-negative, written A ≥ 0; if equality holds only when x = 0, then A  is called positive. The set of self adjoint operators admits a partial order, in which A ≥ B if A − B ≥ 0. If A  has the form B*B  for some B, then A is non-negative; if B is invertible, then A  is positive. A converse is also true in the sense that, for a non-negative operator A, there exists a unique non-negative square root B such that
Quantum Field Theory II - 页 4 92a638a3d6173596bf4932646c6d83e3
In a sense made precise by the spectral theorem, self-adjoint operators can usefully be thought of as operators that are "real". An element A of B(H) is called normal if A*A = AA*. Normal operators decompose into the sum of a self-adjoint operators and an imaginary multiple of a self adjoint operator
Quantum Field Theory II - 页 4 A37eb721ad65c33a67f34c098ac838ac
that commute with each other. Normal operators can also usefully be thought of in terms of their real and imaginary parts.

An element U  of B(H) is called unitary if U  is invertible and its inverse is given by U*. This can also be expressed by requiring that U  be onto and ⟨UxUy⟩ = ⟨xy⟩ for all x andy in H. The unitary operators form a group under composition, which is the isometry group of H.

An element of B(H) is compact if it sends bounded sets to relatively compact sets. Equivalently, a bounded operator T is compact if, for any bounded sequence {xk}, the sequence {Txk} has a convergent subsequence. Many integral operators are compact, and in fact define a special class of operators known as Hilbert–Schmidt operators that are especially important in the study of integral equationsFredholm operators differ from a compact operator by a multiple of the identity, and are equivalently characterized as operators with a finite dimensional kernel and cokernel. The index of a Fredholm operator T is defined by
Quantum Field Theory II - 页 4 54c48ee17e425c5517e08f35b5062cd5
The index is homotopy invariant, and plays a deep role in differential geometry via the Atiyah–Singer index theorem.

Unbounded operators[edit]


Unbounded operators are also tractable in Hilbert spaces, and have important applications to quantum mechanics.[53] An unbounded operator T on a Hilbert space H is defined as a linear operator whose domain D(T) is a linear subspace of H. Often the domain D(T) is a dense subspace of H, in which case T is known as a densely defined operator.

The adjoint of a densely defined unbounded operator is defined in essentially the same manner as for bounded operators. Self-adjoint unbounded operators play the role of theobservables in the mathematical formulation of quantum mechanics. Examples of self-adjoint unbounded operators on the Hilbert space L2(R) are:[54]


  • A suitable extension of the differential operator


Quantum Field Theory II - 页 4 A17bcdbd116079316ea3369c2eb39963where i is the imaginary unit and f is a differentiable function of compact support.

  • The multiplication-by-x operator:


Quantum Field Theory II - 页 4 0eaeb3909db0cc99f933d350b14b2709
These correspond to the momentum and position observables, respectively. Note that neither A nor B is defined on all of H, since in the case of A the derivative need not exist, and in the case of B the product function need not be square integrable. In both cases, the set of possible arguments form dense subspaces of L2(R).

Constructions[edit]


Direct sums[edit]


Two Hilbert spaces H1 and H2 can be combined into another Hilbert space, called the (orthogonal) direct sum,[55] and denoted
Quantum Field Theory II - 页 4 70b1c01343b82e6a404fc0351b7c2b0e
consisting of the set of all ordered pairs (x1x2) where xi ∈ Hii = 1,2, and inner product defined by
Quantum Field Theory II - 页 4 A6b36db614ad867bd3d51a116b1be156
More generally, if Hi is a family of Hilbert spaces indexed by i ∈ I, then the direct sum of the Hi, denoted
Quantum Field Theory II - 页 4 67d2ee953568e191067e5cabe6853358
consists of the set of all indexed families
Quantum Field Theory II - 页 4 Dfa4dd976ee45fe9a22cd60ae25d8fc3
in the Cartesian product of the Hi such that
Quantum Field Theory II - 页 4 E0c34ecfd3f4d4bcb00fc9f8b58344ec
The inner product is defined by
Quantum Field Theory II - 页 4 B0a5cf4d204fcf968afe8e36b07ed49b
Each of the Hi is included as a closed subspace in the direct sum of all of the Hi. Moreover, the Hi are pairwise orthogonal. Conversely, if there is a system of closed subspaces,Vii ∈ I, in a Hilbert space H, that are pairwise orthogonal and whose union is dense in H, then H is canonically isomorphic to the direct sum of Vi. In this case, H is called the internal direct sum of the Vi. A direct sum (internal or external) is also equipped with a family of orthogonal projections Ei onto the ith direct summand Hi. These projections are bounded, self-adjoint, idempotent operators that satisfy the orthogonality condition
Quantum Field Theory II - 页 4 Ea6e21212fca8a76bb754e6ce1ef6610
The spectral theorem for compact self-adjoint operators on a Hilbert space H states that H splits into an orthogonal direct sum of the eigenspaces of an operator, and also gives an explicit decomposition of the operator as a sum of projections onto the eigenspaces. The direct sum of Hilbert spaces also appears in quantum mechanics as the Fock space of a system containing a variable number of particles, where each Hilbert space in the direct sum corresponds to an additional degree of freedom for the quantum mechanical system. In representation theory, the Peter–Weyl theorem guarantees that any unitary representation of a compact group on a Hilbert space splits as the direct sum of finite-dimensional representations.

Tensor products[edit]


Main article: Tensor product of Hilbert spaces

If H1 and H2, then one defines an inner product on the (ordinary) tensor product as follows. On simple tensors, let
Quantum Field Theory II - 页 4 50172f75b9474917fcf10599c5ae03e5
This formula then extends by sesquilinearity to an inner product on H1 ⊗ H2. The Hilbertian tensor product of H1 and H2, sometimes denoted by Quantum Field Theory II - 页 4 C9d22a983800b403012e47fb214b727a, is the Hilbert space obtained by completing H1 ⊗ H2 for the metric associated to this inner product.[56]

An example is provided by the Hilbert space L2([0, 1]). The Hilbertian tensor product of two copies of L2([0, 1]) is isometrically and linearly isomorphic to the space L2([0, 1]2) of square-integrable functions on the square [0, 1]2. This isomorphism sends a simple tensor Quantum Field Theory II - 页 4 2820db115e1f4921619ff24045e221ea to the function
Quantum Field Theory II - 页 4 1aece0b670806f941adb84de16f9adef
on the square.

This example is typical in the following sense.[57] Associated to every simple tensor product x1 ⊗ x2 is the rank one operator from H1 to H2 that maps a given Quantum Field Theory II - 页 4 9bf7751c837b05da2f0753b64d1d99e4 as
Quantum Field Theory II - 页 4 8a5c08b05975ca8b06a6982d380b40c6
This mapping defined on simple tensors extends to a linear identification between H1 ⊗ H2 and the space of finite rank operators from H*1 to H2. This extends to a linear isometry of the Hilbertian tensor product Quantum Field Theory II - 页 4 C9d22a983800b403012e47fb214b727a with the Hilbert space HS(H*1H2) of Hilbert–Schmidt operators from H*1 to H2.

Orthonormal bases[edit]


The notion of an orthonormal basis from linear algebra generalizes over to the case of Hilbert spaces.[58] In a Hilbert space H, an orthonormal basis is a family {ek}k ∈ B of elements of H satisfying the conditions:

[list="color: rgb(37, 37, 37); font-family: sans-serif; font-size: 14px; line-height: 1.6; margin-top: 0.3em; margin-right: 0px; margin-bottom: 0px; margin-left: 3.2em; padding-top: 0px; padding-right: 0px; padding-bottom: 0px; padding-left: 0px; list-style-image: none;"]
[*]Orthogonality: Every two different elements of B are orthogonal: ⟨ekej⟩= 0 for all kj in B with k ≠ j.

[*]Normalization: Every element of the family has norm 1:||ek|| = 1 for all k in B.

[*]Completeness: The linear span of the family ekk ∈ B, is dense in H.

[/list]

A system of vectors satisfying the first two conditions basis is called an orthonormal system or an orthonormal set (or an orthonormal sequence if B is countable). Such a system is always linearly independent. Completeness of an orthonormal system of vectors of a Hilbert space can be equivalently restated as:
if ⟨v, ek⟩ = 0 for all k ∈ B and some v ∈ H then v = 0.
This is related to the fact that the only vector orthogonal to a dense linear subspace is the zero vector, for if S is any orthonormal set and v is orthogonal to S, then v is orthogonal to the closure of the linear span of S, which is the whole space.

Examples of orthonormal bases include:


  • the set {(1,0,0), (0,1,0), (0,0,1)} forms an orthonormal basis of R3 with the dot product;

  • the sequence {fn : n ∈ Z} with fn(x) = exp(2πinx) forms an orthonormal basis of the complex space L2([0,1]);



In the infinite-dimensional case, an orthonormal basis will not be a basis in the sense of linear algebra; to distinguish the two, the latter basis is also called a Hamel basis. That the span of the basis vectors is dense implies that every vector in the space can be written as the sum of an infinite series, and the orthogonality implies that this decomposition is unique.

Sequence spaces[edit]


The space  2 of square-summable sequences of complex numbers is the set of infinite sequences
Quantum Field Theory II - 页 4 08b1d8b5a70427f029878554266df143
of complex numbers such that
Quantum Field Theory II - 页 4 6655c59f3aace9d0794381c869364b14
This space has an orthonormal basis:
Quantum Field Theory II - 页 4 E08267ef079950fd2fb732ca7be43b2f
More generally, if B is any set, then one can form a Hilbert space of sequences with index set B, defined by
Quantum Field Theory II - 页 4 Ffb80a7cd765804b889699888dbdd10c
The summation over B is here defined by
Quantum Field Theory II - 页 4 D10ad275e58df04a11e97a932aa5107c
the supremum being taken over all finite subsets of B. It follows that, for this sum to be finite, every element of  2(B) has only countably many nonzero terms. This space becomes a Hilbert space with the inner product
Quantum Field Theory II - 页 4 75ec175a72eaf0bbf4b1418a6a412947
for all x and y in  2(B). Here the sum also has only countably many nonzero terms, and is unconditionally convergent by the Cauchy–Schwarz inequality.

An orthonormal basis of  2(B) is indexed by the set B, given by
Quantum Field Theory II - 页 4 7e11a546c96dd6858d3fce6fb0d57056


Bessel's inequality and Parseval's formula[edit]


Let f1, …, fn be a finite orthonormal system in H. For an arbitrary vector x in H, let
Quantum Field Theory II - 页 4 677c7f7849a676fb9e5fb9252093b029
Then ⟨xfk⟩ = ⟨yfk⟩ for every k = 1, …, n. It follows that x − y is orthogonal to each fk, hence x − y is orthogonal to y. Using the Pythagorean identity twice, it follows that
Quantum Field Theory II - 页 4 6ab3c39f679d6ed54ec7461a8d5732ff
Let {fi }, i ∈ I, be an arbitrary orthonormal system in H. Applying the preceding inequality to every finite subset J of I gives the Bessel inequality[59]
Quantum Field Theory II - 页 4 6a6942673585b73758ebb11919d1fba6
(according to the definition of the sum of an arbitrary family of non-negative real numbers).

Geometrically, Bessel's inequality implies that the orthogonal projection of x onto the linear subspace spanned by the fi has norm that does not exceed that of x. In two dimensions, this is the assertion that the length of the leg of a right *** may not exceed the length of the hypotenuse.

Bessel's inequality is a stepping stone to the more powerful Parseval identity, which governs the case when Bessel's inequality is actually an equality. If {ek}k ∈ B is an orthonormal basis of H, then every element x of H may be written as
Quantum Field Theory II - 页 4 92455dc15b36a5c6aff58f019723ce99
Even if B is uncountable, Bessel's inequality guarantees that the expression is well-defined and consists only of countably many nonzero terms. This sum is called the Fourier expansion of x, and the individual coefficients ⟨x,ek⟩ are the Fourier coefficients of x. Parseval's formula is then
Quantum Field Theory II - 页 4 50b58752bfff34b2bfcf724a68818a9f
Conversely, if {ek} is an orthonormal set such that Parseval's identity holds for every x, then {ek} is an orthonormal basis.

Hilbert dimension[edit]


As a consequence of Zorn's lemmaevery Hilbert space admits an orthonormal basis; furthermore, any two orthonormal bases of the same space have the same cardinality, called the Hilbert dimension of the space.[60] For instance, since 2(B) has an orthonormal basis indexed by B, its Hilbert dimension is the cardinality of B (which may be a finite integer, or a countable or uncountable cardinal number).

As a consequence of Parseval's identity, if {ek}k ∈ B is an orthonormal basis of H, then the map Φ : H →  ℓ2(B) defined by Φ(x) = (⟨x,ek⟩)kB is an isometric isomorphism of Hilbert spaces: it is a bijective linear mapping such that
Quantum Field Theory II - 页 4 9a4ce598ef63c0c388a748b44f086aad
for all x and y in H. The cardinal number of B is the Hilbert dimension of H. Thus every Hilbert space is isometrically isomorphic to a sequence space  ℓ2(B) for some set B.
一星
一星

帖子数 : 3787
注册日期 : 13-08-07

返回页首 向下

Quantum Field Theory II - 页 4 Empty 回复: Quantum Field Theory II

帖子 由 一星 2014-08-30, 01:14

[ltr]

Separable spaces[size=13][edit]



A Hilbert space is separable if and only if it admits a countable orthonormal basis. All infinite-dimensional separable Hilbert spaces are therefore isometrically isomorphic to 2.
In the past, Hilbert spaces were often required to be separable as part of the definition.[61] Most spaces used in physics are separable, and since these are all isomorphic to each other, one often refers to any infinite-dimensional separable Hilbert space as "the Hilbert space" or just "Hilbert space".[62] Even in quantum field theory, most of the Hilbert spaces are in fact separable, as stipulated by the Wightman axioms. However, it is sometimes argued that non-separable Hilbert spaces are also important in quantum field theory, roughly because the systems in the theory possess an infinite number of degrees of freedom and any infinite Hilbert tensor product (of spaces of dimension greater than one) is non-separable.[63] For instance, a bosonic field can be naturally thought of as an element of a tensor product whose factors represent harmonic oscillators at each point of space. From this perspective, the natural state space of a boson might seem to be a non-separable space.[63] However, it is only a small separable subspace of the full tensor product that can contain physically meaningful fields (on which the observables can be defined). Another non-separable Hilbert space models the state of an infinite collection of particles in an unbounded region of space. An orthonormal basis of the space is indexed by the density of the particles, a continuous parameter, and since the set of possible densities is uncountable, the basis is not countable.[63]

Orthogonal complements and projections[edit]



If S is a subset of a Hilbert space H, the set of vectors orthogonal to S is defined by
Quantum Field Theory II - 页 4 07e405c83956ac647709a5f5a83a645b
S is a closed subspace of H (can be proved easily using the linearity and continuity of the inner product) and so forms itself a Hilbert space. If V is a closed subspace of H, thenV is called the orthogonal complement of V. In fact, every x in H can then be written uniquely as x = v + w, with v in V and w in V. Therefore, H is the internal Hilbert direct sum ofV and V.
The linear operator PV : H → H that maps x to v is called the orthogonal projection onto V. There is a natural one-to-one correspondence between the set of all closed subspaces of H and the set of all bounded self-adjoint operators P such that P2 = P. Specifically,
Theorem. The orthogonal projection PV is a self-adjoint linear operator on H of norm ≤ 1 with the property P2V = PV. Moreover, any self-adjoint linear operator E such that E2E is of the form PV, where V is the range of E. For every x in H, PV(x) is the unique element v of V, which minimizes the distance ||x − v||.
This provides the geometrical interpretation of PV(x): it is the best approximation to x by elements of V.[64]
Projections PU and PV are called mutually orthogonal if PUPV = 0. This is equivalent to U and V being orthogonal as subspaces of H. The sum of the two projections PU and PV is a projection only if U and V are orthogonal to each other, and in that case PU + PV = PU+V. The composite PUPV is generally not a projection; in fact, the composite is a projection if and only if the two projections commute, and in that case PUPV = PUV.
By restricting the codomain to the Hilbert space V, the orthogonal projection PV gives rise to a projection mapping π: H → V; it is the adjoint of the inclusion mapping
Quantum Field Theory II - 页 4 8e87f612d96be9dd5c945fc1940a5060
meaning that
Quantum Field Theory II - 页 4 8c16c7a5dda2218b3963c537f172ee1f
for all x ∈ V and y ∈ H.
The operator norm of the orthogonal projection PV onto a non-zero closed subspace V is equal to one:
Quantum Field Theory II - 页 4 075b52ce4a4b88e9f620de18256482cb
Every closed subspace V of a Hilbert space is therefore the image of an operator P of norm one such that P2 = P. The property of possessing appropriate projection operators characterizes Hilbert spaces:[65][/ltr][/size]



  • A Banach space of dimension higher than 2 is (isometrically) a Hilbert space if and only if, for every closed subspace V, there is an operator PV of norm one whose image is Vsuch that Quantum Field Theory II - 页 4 D9b1b9eea5fdb4f6d41ae90520cc6f3d


[ltr]
While this result characterizes the metric structure of a Hilbert space, the structure of a Hilbert space as a topological vector space can itself be characterized in terms of the presence of complementary subspaces:[66][/ltr]


  • A Banach space X is topologically and linearly isomorphic to a Hilbert space if and only if, to every closed subspace V, there is a closed subspace W such that X is equal to the internal direct sum V ⊕ W.


[ltr]
The orthogonal complement satisfies some more elementary results. It is a monotone function in the sense that if U ⊂ V, then Quantum Field Theory II - 页 4 4b02335aca117069d73be381353f80d2 with equality holding if and only if V is contained in the closure of U. This result is a special case of the Hahn–Banach theorem. The closure of a subspace can be completely characterized in terms of the orthogonal complement: If V is a subspace of H, then the closure of V is equal to Quantum Field Theory II - 页 4 D04203f80f97a13bd9190c29aef7b88b. The orthogonal complement is thus a Galois connection on the partial order of subspaces of a Hilbert space. In general, the orthogonal complement of a sum of subspaces is the intersection of the orthogonal complements:[67] Quantum Field Theory II - 页 4 A04d2cda6ede5ac359422949f45d39da. If the Vi are in addition closed, then Quantum Field Theory II - 页 4 37517b2912180da1269359b52da3ec4a.

Spectral theory[edit]



There is a well-developed spectral theory for self-adjoint operators in a Hilbert space, that is roughly analogous to the study of symmetric matrices over the reals or self-adjoint matrices over the complex numbers.[68] In the same sense, one can obtain a "diagonalization" of a self-adjoint operator as a suitable sum (actually an integral) of orthogonal projection operators.
The spectrum of an operator T, denoted σ(T) is the set of complex numbers λ such that T − λ lacks a continuous inverse. If T is bounded, then the spectrum is always a compact set in the complex plane, and lies inside the disc Quantum Field Theory II - 页 4 88166d2eb908e5c0b1fcc9ee5715c948 If T is self-adjoint, then the spectrum is real. In fact, it is contained in the interval [m,M] where
Quantum Field Theory II - 页 4 C2eceb704c67eb72bf87a8ac14784bb3
Moreover, m and M are both actually contained within the spectrum.
The eigenspaces of an operator T are given by
Quantum Field Theory II - 页 4 D4cd58c85714d72d73246e3eaf1cdf14
Unlike with finite matrices, not every element of the spectrum of T must be an eigenvalue: the linear operator T − λ may only lack an inverse because it is not surjective. Elements of the spectrum of an operator in the general sense are known as spectral values. Since spectral values need not be eigenvalues, the spectral decomposition is often more subtle than in finite dimensions.
However, the spectral theorem of a self-adjoint operator T takes a particularly simple form if, in addition, T is assumed to be a compact operator. The spectral theorem for compact self-adjoint operators states:[69][/ltr]


  • A compact self-adjoint operator T has only countably (or finitely) many spectral values. The spectrum of T has no limit point in the complex plane except possibly zero. The eigenspaces of T decompose H into an orthogonal direct sum:Quantum Field Theory II - 页 4 12596a3d1a8ad753b109449c5a406ad1


[ltr]
Moreover, if Eλ denotes the orthogonal projection onto the eigenspace Hλ, thenQuantum Field Theory II - 页 4 A2c9c4c45548032a913e20b2c3f13943where the sum converges with respect to the norm on B(H).
This theorem plays a fundamental role in the theory of integral equations, as many integral operators are compact, in particular those that arise from Hilbert–Schmidt operators.
The general spectral theorem for self-adjoint operators involves a kind of operator-valued Riemann–Stieltjes integral, rather than an infinite summation.[70] The spectral familyassociated to T associates to each real number λ an operator Eλ, which is the projection onto the nullspace of the operator (T − λ)+, where the positive part of a self-adjoint operator is defined by
Quantum Field Theory II - 页 4 054b4a0b5ddb932064d588f7147c3ca0
The operators Eλ are monotone increasing relative to the partial order defined on self-adjoint operators; the eigenvalues correspond precisely to the jump discontinuities. One has the spectral theorem, which asserts
Quantum Field Theory II - 页 4 C424bbea6836bf61c67a4355638cb666
The integral is understood as a Riemann–Stieltjes integral, convergent with respect to the norm on B(H). In particular, one has the ordinary scalar-valued integral representation
Quantum Field Theory II - 页 4 1d145b1cecf00f5bb124310d691382d6
A somewhat similar spectral decomposition holds for normal operators, although because the spectrum may now contain non-real complex numbers, the operator-valued Stieltjes measure dEλ must instead be replaced by a resolution of the identity.
A major application of spectral methods is the spectral mapping theorem, which allows one to apply to a self-adjoint operator T any continuous complex function f defined on the spectrum of T by forming the integral
Quantum Field Theory II - 页 4 7b8a0fc4bc11e5a8bcdc6933c9f73911
The resulting continuous functional calculus has applications in particular to pseudodifferential operators.[71]
The spectral theory of unbounded self-adjoint operators is only marginally more difficult than for bounded operators. The spectrum of an unbounded operator is defined in precisely the same way as for bounded operators: λ is a spectral value if the resolvent operator
Quantum Field Theory II - 页 4 D70cf8dfb85bb15ae7a577b88d823e08
fails to be a well-defined continuous operator. The self-adjointness of T still guarantees that the spectrum is real. Thus the essential idea of working with unbounded operators is to look instead at the resolvent Rλ where λ is non-real. This is a bounded normal operator, which admits a spectral representation that can then be transferred to a spectral representation of T itself. A similar strategy is used, for instance, to study the spectrum of the Laplace operator: rather than address the operator directly, one instead looks as an associated resolvent such as a Riesz potential or Bessel potential.
A precise version of the spectral theorem in this case is:[72]
Given a densely defined self-adjoint operator T on a Hilbert space H, there corresponds a unique resolution of the identity E on the Borel sets of R, such thatQuantum Field Theory II - 页 4 18d3e1cec23525e77ec84c905240b397for all x ∈ D(T) and y ∈ H. The spectral measure E is concentrated on the spectrum of T.
There is also a version of the spectral theorem that applies to unbounded normal operators.

See also[edit]

[/ltr]

Quantum Field Theory II - 页 4 28px-Nuvola_apps_edu_mathematics_blue-p.svgMathematics portal


[ltr]

Notes[edit]

[/ltr]


[ltr]

References[edit]

[/ltr]






一星
一星

帖子数 : 3787
注册日期 : 13-08-07

返回页首 向下

Quantum Field Theory II - 页 4 Empty 回复: Quantum Field Theory II

帖子 由 一星 2014-09-05, 03:32

Historical remarks. In 1954, the young physicists Yang and Mills tried to
generalize the Maxwell equations. They used an idea of Hermann Weyl published
in 1929.
• Weyl formulated the Maxwell equations as a gauge theory based on the commutative
Lie group U(1).
• The goal of Yang and Mills was to replace the commutative group U(1) by the
noncommutative Lie group SU(2).
To this end, they replaced the relation
Fjk = ∂jAk − ∂kAj, j= 1, 2, 3, 4
between the electromagnetic field tensor {Fij} and the 4-potential {Aj} by a modified
relation of the type
Fjk = ∂jAk − ∂kAj + AjAk −AkAj, j,k= 1, 2, 3, 4, (5.98)
where the complex 2 × 2 matrices A1,A2,A3,A4 are elements of the Lie algebra
su(2) of the symmetry group SU(2). Dianzhou Zhang asked Professor Yang in
an interview: An interesting question is whether you understood in 1954 the
tremendous importance of your joint paper with Mills on noncommutative gauge
theory. Yang answered:
No. In the 1950s we felt our work was elegant. I realized its importance in
the 1960s and its great importance to physics in the 1970s. Its relationship
to deep mathematics became only clear to me after 1974.
In the 1960s and the early 1970s, equations of the type (5.98) above were used in
order to formulate the Standard Model in elementary particle physics. Here, the
symmetry group U(1) × SU(2) × SU(3) is used.
• The crucial field tensor {Fij} describes all the 12 particles (i.e., the photon, 8
gluons, and 3 vector bosons) which are responsible for the interaction
• between the 12 basic particles (i.e., 6 quarks and 6 leptons – the electron, 3
neutrinos, the muon, and the tau) and their antiparticles.
The typical difficulty of any gauge theory is the fact that the interacting particles
are massless at the very beginning. In contrast to this, the three vector bosons
observed in nature are quite heavy – their masses equal about 100 proton masses.
One needs an additional field – the Higgs field – in order to generate the masses
of the vector bosons. This so-called Higgs mechanism will be thoroughly studied in
Vol. III.
In the late 1960s, Yang discovered the relation between Yang–Mills theory and
Riemannian geometry. He reports in the same interview as quoted above:
In the late 1960s, I began a new formulation of gauge fields, through the
approach of non-integrable phase factors. It happend that one semester, I
was teaching general relativity, and I noticed that the formula (5.93) above
in gauge theory and the formula (5.94) above in Riemannian geometry
are not just similar – they are, in fact the same if one makes the right
identification of symbols.
Yang continues:
With an appreciation of the geometrical meaning of gauge theory, I consulted
Jim Simons, a distinguished geometer, who was then the chairman
of the Mathematics Department at Stony Brook (Long Island, New York).
He said gauge theory must be related to connections on fiber bundles. I
then tried to understand fiber bundle theory from such books as Steenrod’s
The Topology of Fiber Bundles, Princeton University Press, 1951,
but I learned nothing. The language of modern mathematics is too cold
and abstract for a physicist.61
In 1975, Wu and Yang wrote a paper about global gauge theory. In this paper, they
published a quite interesting dictionary about the completely different terminology
of mathematicians and physicists concerning the same topic. For example:
connection in mathematics ⇔ potential in physics,
curvature in mathematics ⇔ field tensor (interaction) in physics,
structural equation ⇔ field tensor − potential relation,
change of bundle coordinates ⇔ gauge transformation,
structure group ⇔ gauge group.
In a long historical process, mathematicians tried to understand curvature, but
physicists studied the forces acting in the universe. Nowadays we know that mathematicians
and physicists did the same from an abstract mathematical point of view.
Observe that
The Cartan curvature 2-form is more fundamental than the Riemann curvature
tensor.
In fact, since RP (a, b, u, v) := −<F(a, b)u|v>P for all a, b, u, v ∈ TPHR, the Riemann
curvature tensor only works if the tangent space of the manifold is equipped with the
additional structure of a Hilbert space (or an indefinite Hilbert space, as in general
relativity). However, the notion of curvature in the sense of Cartan is independent of
such an additional structure. In terms of mathematics, the notion of curvature can
be introduced without using a metric, one only needs what is called a connection.
In terms of physics, this corresponds to the transport of information.

5.11 Ariadne’s Thread in Gauge Theory
In terms of physics, roughly speaking, gauge theory describes additional
internal degrees of freedom of physical systems which do not affect the
physics.
For example, the choice of the potential does not influence the forces in classical
Newtonian mechanics.
From the mathematical point of view, gauge theory studies invariants under
gauge transformations.
Only such gauge invariants can be observed in physical experiments. Therefore,
gauge theory is part of the theory of invariants. In order to explain the basic ideas
of gauge theory, let us return to the hyperbolic plane HR.

5.11.1 Parallel Transport of Physical Information – the Key to
Modern Physics
In modern physics, the Huygens principle is replaced by the parallel transport
of physical information.
Folklore

5.11.2 The Phase Equation and Fiber Bundles
In a natural way, the global mathematical description of physical fields is
based on the language of bundles.
Folklore
The main trick due to ´Elie Cartan is to reduce the parallel transport of a physical
field to a dynamical system in a principal fiber bundle (Cartan’s method of moving
frame). To this end, we will replace the differential equation of parallel transport
by the phase equation.

Summarizing, we get the following:
The global curvature form F (defined on the principal fiber bundle P =
HR × GL(2,R) with values in the Lie algebra gl(2,R)) represents an invariant
mathematical object which carries all the information on the gauge
transformations of the Cartan curvature 2-form on the base manifold HR.

5.11.4 Perspectives
The approach above was chosen in order to display the historical development from
Gauss’ theorema egregium to Cartan’s structural equation. In modern differential
geometry, the axiomatic presentation reverses the historical order.
• The starting point is the fundamental phase equation, as a differential equation
on a principal fiber bundle P.
• More precisely, one starts with a velocity field on P which splits into horizontal
and vertical velocity vectors, with respect to the fibers.
• This velocity vector field generates the dynamical system of parallel transport.
The corresponding ordinary differential equation coincides with the equation of
parallel transport, which is based on the connection 1-form A on P.
• Projection onto horizontal vector fields is used in order to replace the differential
dω by the covariant differential Dω of differential forms ω on the principal fiber
bundle P.
• This yields the curvature form
F := DA, (5.112)
and the Bianchi identity DF = 0. It turns out that (5.112) is equivalent to
Cartan’s structural equation (5.110).
The extremely elegant formula (5.112) generalizes both the theorema egregium
of Gauss and the relation between the electromagnetic field tensor F and the 4-
potential A in Maxwell’s theory of electromagnetism. Moreover, formula (5.112) lies
at the heart of both the Standard Model in elementary physics and Einstein’s theory
of general relativity. In terms of mathematics, this approach effectively describes
all kinds of differential geometries:
• The typical Lie group of the principal fiber bundle P is the symmetry group of
the geometry under consideration.
• Physicists are interested in the study of partial differential equations for physical
fields. In this general setting, physical fields ψ are sections of a vector bundle V
associated to the principal fiber bundle P.
• There exists a natural way of transplanting the parallel transport from P to V.
• The parallel transport on V generates the corresponding covariant differentiation
for physical fields, which is used in order to formulate the basic partial differential
equations for the physical field under consideration.
The point is that one has to replace product bundles by general fiber bundles.
Observe that:
General fiber bundles are obtained by gluing local product bundles together,
with the aid of a cocycle.
In particular, the choice of arbitrary Lie groups allows us to take all kind of symmetries
into account. In 1872 Felix Klein (1849–1925) formulated his Erlangen
program:
Geometry is the invariant theory of transformation groups (symmetry
groups).
Sophus Lie (1842–1899), ´Elie Cartan (1859–1951), and their successors realized
this program in differential geometry by investigating Lie groups and the invariant
calculus of differential forms. This is a fascinating chapter in the history of mathematics
and physics. We will thoroughly study this in Vol. III. We also refer to
the standard textbook by S. Kobayashi and K. Nomizu, Foundations of Differential
Geometry, Vols. 1, 2, Wiley, New York, 1963.

5.12 Classification of Two-Dimensional Compact
Manifolds
The notion of manifold is of fundamental importance for both mathematics
and physics. There arises the problem of classifying manifolds in terms
of topology. This has been one of the most important research topics in
topology in the last 150 years.
Folklore
In 1863 M¨obius classified the compact orientable 2-manifolds.
In 1865 M¨obius published a strange su***ce called the
M¨obius strip nowadays. This twisted su***ce is obtained from a rectangle by gluing
together two opposite sides in a twisted manner (Fig. 5.26). If we walk along this
su***ce, then we reach both sides of the rectangle. Therefore, the M¨obius strip is a
one-sided su***ce, and hence it is not possible to define an orientation. Intuitively,
a su***ce is oriented iff the movement of a small oriented circle along a closed
curve never changes the orientation of the circle. The main theorem on classical
topological su***ce theory reads as follows:
The compact 2-manifolds M and N (without or with boundary) are homeomorphic
iff the following three conditions hold:
(i) M and N have the same genus g,
(ii) M and N have the same number of boundaries, and
(iii) both M and N are either orientable or non-orientable.
The genus attains the values g = 0, 1, 2, . . . Let us first consider a few typical
examples. Recall that the Euler characteristic of the 2-manifold M is given by
χ(M) = β0 − β1 + β2,
where β0, β1, β2 are the Betti numbers of M. Since the 2-manifold M is arcwise
connected, we always have β0 = 1. Moreover, we have β2 = 1 (resp. β2 = 0) iff M
is orientable (resp. non-orientable).
• The 2-sphere S2r
of radius r > 0 : This su***ce has the genus g = 0, the Betti
numbers
β0 = β2 = 1, β1 = 0
and the Euler characteristic χ(S2r
) = 2. The fundamental group is trivial,
π1(S2r
) = 0, that is, the sphere S2r
is simply connected.
• The 2-dimensional torus T2 : This su***ce can be obtained by identifying the
opposite boundary points of a rectangle (Fig. 5.27). Alternatively, the torus is
homeomorphic to a sphere with one handle attached to it (Fig. 5.25(a)). This
su***ce has the genus g = 1, the Betti numbers
β0 = β2 = 2, β1 = 2,
and the Euler characteristic χ(T2) = 0. For the additive fundamental group, we
get
π1(T2) = Z ⊕ Z.
This reflects the fact that there exist two different types of closed curves on the
torus which cannot be contracted to one point. For example, in Fig. 5.27(c) this
concerns the equator ABA and the meridian circle PAP. Consequently, the torus
T2 is not simply connected.
• The real 2-dimensional projective space P2 : This space is obtained from the
closed unit disc by identifying diametrically opposed boundary points of the unit
disc with each other (Fig. 5.28). The topological space P2 is compact, arcwise
connected, non-orientable, and it has the Betti numbers
β0 = 1, β1 = β2 = 0.
This yields the Euler characteristic χ(P2) = 1. The genus of P2 is given by g = 1,
as we will discuss below. The additive fundamental group of P2 is given by
π1(P2) = Z2.
This group consists of the two elements {0, 1} with 1+1 = 0. The element ”1” of
π1(P2) corresponds to the boundary curve PQP . This curve cannot be
continuously contracted to a point within P2. Thus, the projective space P2 is not
simply connected. In contrast to this, taking the identification of diametrically
opposed boundary points into account, the curve PQPQP can be continuously
contracted to the center of the unit disc. This corresponds to ”1+1=0.”
Let us now study the general case. We are given the compact 2-manifold M
of genus g. Let r = 0, 1, 2, . . . be the number of boundaries. Then there exists a
manifold N which is homeomorphic toMand which represents one of the following
normal forms:76
(I) Suppose that M is orientable.
(I-1) Let r = 0. Then the normal form N is obtained from the unit sphere S2 by
taking 2g open discs away and by attaching g handles. The genus is also equal
to the number of ‘holes’ (Fig. 5.25).
• Betti numbers of M: β0 = β2 = 1, β1 = 2g.
• Euler characteristic of M: χ = 2− 2g.
• Additive fundamental group of M: π1 = Z ⊕ . . . ⊕ Z (2g summands).
(I-2) Let r > 0. The normal form N is obtained from (I-1) by taking r open discs
away .
• Betti numbers of M: β0 = β2 = 1 and β1 = 2g + r − 1.
• Euler characteristic of M: χ = 3− 2g − r.
(II) Suppose that M is non-orientable.
(II-1) Let r = 0. The normal form N is obtained from the unit sphere S2 by taking
g open discs away, and by identifying diametral points of the boundary circles
with each other. The number g is the genus of M. For example, the su***ce
pictured in Fig. 5.30(a) is homeomorphic to the projective space P2.
• Betti numbers of M: β0 = 1, β1 = g − 1, β2 = 0. Here, g = 1, 2, . . .
• Euler characteristic of M: χ = 2− g.
(II-2) Let r > 0. The normal form is obtained from (II-1) by taking r open discs
away. Again the number g is called the genus of M.
• Betti numbers of M: β1 = 1, β1 = g + r − 1, β2 = 0.
• Euler characteristic of M: χ = 2− g − r.
In particular, the M¨obius strip corresponds to g = 1 and r = 1. Hence β1 = 1
and χ = 0.

5.13 The Poincar´e Conjecture and the Ricci Flow
The study of mathematics, like the Nil, begins in minuteness, but ends in
magnificence.
C.C. Colton, 1820
Friedman’s 1982 proof of the 4-dimensional Poincar´e hypothesis was an
extraordinary tour de force. His methods were so sharp that as to actually
provide a complete classification of all compact simply connected topological
4-manifolds, yielding many previously unknown examples of such
manifolds.
John Milnor, 1986
The Poincar´e conjecture for 3-manifolds was one of the seven Millenium Prize Problems
announced by the Clay Mathematics Institute in Cambridge, Massachusetts,
in the year 2000 (see Sect. 1.7 of Vol. I).
The topological characterization of the 2-sphere. The following result
was already known in the second half of the 19th century:
A compact simply connected 2-manifold is homeomorphic to
a 2-sphere.
This homeomorphism can be chosen as a diffeomorphism.
The topological characterization of the 3-sphere. The famous Poincar´e
conjecture claims the following:
A compact simply connected 3-manifold is diffeomorphic to
a 3-sphere.
This conjecture was proven by Grigori Perelman (born 1966) in 2003. He invented
an ingenious approach for solving this outstanding problem. The main idea comes
from physics. Here, Perelman uses the physical picture of the flow of fluid particles
in order to deform the original 3-manifold into the final 3-sphere. More precisely,
he applies the so-called Ricci flow on manifolds which was thoroughly studied in
the pioneering papers by Richard Hamilton in the 1980s and 1990s. The Ricci flow
is governed by the partial differential equation
∂g(t)/∂t= −2Ric(g(t)), t≥ 0. (5.113)
Here, the parameter t can be regarded as time. This equation describes the timedeformation
of the metric tensor g of a Riemannian manifold M. Equation (5.113)
is of the type of a diffusion process (or heat conduction process). The main difficulty
is that there may appear singularities during the time evolution. One has to control
the possible types of singularities, and one has to regularize and renormalize them
(see the hints for further reading on page 351).
The generalization of the idea of the flow of fluid particles is also crucial for
quantum physics. In Sect. 7.11.5 we will show that the Feynman path integral
corresponds to a diffusion process in imaginary time (the Feynman–Kac formula).
Moreover, the modern approach to renormalization in quantum field theory is based
on the flow generated by the renormalization group (see Chap. 3 of Vol. I).
The homotopy type of a topological space. Recall from Vol. I the following
terminology. Let X and Y be topological spaces. Let idX denote the identity map
on X. The two continuous maps f, g : X → X are called homotopic iff there exists
a continuous map H : X × [0, 1] → X such that
H(x, 0) = f(x) and H(x, 1) = g(x) for all x ∈ X.
We write f ∼ g. The map H is called a homotopy.
By definition, the space X has the same homotopy type as the space Y iff there
exist continuous maps F : X → Y and G : Y → X such that
GF ∼ idX and FG ∼ idY .
That is, the composite maps GF : X → X and FG : Y → Y are homotopic to the
corresponding identity maps. We also say that the space X is homotopy equivalent
to the space Y . For example, a topological space is called contractible iff it has the
same homotopy type as a single point (see Sect. 5.5 of Vol. I)..
By definition, the topological space X has the same topological type as the
topological space Y iff X is homeomorphic to Y . Explicitly, this means that there
exists a bijective continuous map F : X → Y such that the inverse map F−1 : Y →
X is also continuous. Setting G := F−1, we get GF = idX and FG = idY. Thus, if
X and Y have the same topological type, then they also have the same homotopy
type. However, the converse is not always true.
The topological characterization of the n-sphere. The generalized Poincar
´e conjecture reads as follows:
If an n-manifold has the same homotopy type as an n-sphere, then it has
actually the same topological type as an n-sphere.
Nowadays we know that this statement is true for all dimensions n = 1, 2, . . .
For n ≥ 5, the proof was given by Smale, and independently by Stallings and
Zeeman and by Wallace in 1960-61. For n = 4, Freedman gave the proof in 1982.
Perelman settled the most difficult case n = 3 in 2002. For their seminal results,
Smale, Freedman and Perelman were awarded the Fields medal in 1966, 1986, and
2006, respectively. Perelman refused the award. Another formulation of the Poincar´e
conjecture reads as follows:
If an n-manifold has the same fundamental group and the same
homology as the n-sphere, then it is actually homeomorphic to the
n-sphere.78
This is true for all dimensions n = 1, 2, . . . The positive answer to the Poincar´e
conjecture tells us the following highly nontrivial result:
Algebraic topology is able to detect spheres in all dimensions.

5.14 A Glance at Modern Optimization Theory
In the 1950s, modern control theory was founded by generalizing the duality between
wave fronts and light rays.
(i) Dynamic programming (generalized wave fronts): Bellman created dynamic programming
by generalizing the Hamilton–Jacobi partial differential equation to
the Bellman functional equation. In geometrical optics, the eikonal function S
measures the time needed for the propagation of light. In dynamic programming,
the function S measures the quantity to be optimized (e.g., the costs of
a production process).
(ii) Pontryagin’s maximum principle (generalized light rays): Pontryagin generalized
the Hamilton canonical equations for light rays (and the maximum principle
in geometrical optics) to the computation of optimal trajectories in modern
technology. Let us discuss some important examples.
• For the return of a spaceship from moon to earth, one has to compute a
trajectory such that the heating of the spaceship remains minimal. Of interest
in the optimal solution is the fact that the spaceship penetrates the
earth’s atmosphere rather deeply (from 120 km altitude to 50 km) and then
it climbs again to the altitude of 75 km. On the other hand, the velocity falls
almost monotonically. The computation (performed by Roland Bulirsch for
the NASA in the 1960s) can be found in Stoer and Bulirsch, Introduction
to Numerical Analysis, Springer, New York, 1993.
• For the moon landing, one needs a feed-back control program which guarantees
minimal fuel consumption of the moon landing ferry. Here, the braking
process is controlled by measuring the distance between the ferry and the
moon su***ce.
• For the flight to Mars, one needs a trajectory of the spaceship which again
guarantees minimal fuel consumption, by taking the gravitational forces of
the planets in our solar system into account.
The proof for the validity of Pontryagin’s maximum principle is highly sophisticated.
For a detailed study of (i) and (ii) in the setting of nonlinear functional
analysis, we refer to Zeidler (1986), Vol. III, quoted on page 353.
一星
一星

帖子数 : 3787
注册日期 : 13-08-07

返回页首 向下

Quantum Field Theory II - 页 4 Empty 回复: Quantum Field Theory II

帖子 由 一星 2014-09-07, 06:01

6. The Principle of Critical Action and the
Harmonic Oscillator – Ariadne’s Thread in
Classical Mechanics
Since the divine plan is the most perfect thing there is, there can be no
doubt that all actions in the universe can be determined by the calculus
of the minima and maxima from the corresponding causes.
Leonhard Euler
The history of the principle of least action has often been described. Yet
the matter is still controversial, and there seems to be no general agreement
who invented the principle, Leibniz (1646–1717), Euler (1707–1783),
or Maupertuis (1698–1759). . . We mention that the first mathematical
treatment of the action principle was given by Euler in the Additamentum
of his Methodus inveniendi.
Mariano Giaquinta and Stefan Hildebrandt, 1996
By generalizing the method of Euler in the calculus of variations, Lagrange
(1736–1813) discovered, how one can write, in a single line, the basic equation
for all problems in analytic mechanics.
Carl Gustav Jakob Jacobi (1804–1851)
When we quantize a classical theory, wave packets behave like particles. . .
A wave packet might decay into two wave packets. When two wave packets
come near to each other, they scatter and perhaps produce more wave
packets. This naturally suggests the physics of particles can be described
in these terms. . .
Quantum field theory grew out of essentially these sorts of physical ideas.
It struck me as limiting that even after some 75 years, the whole subject
of quantum field theory remains rooted in this harmonic paradigm, to use
a dreadfully pretentious word. We have not been able to get away from
the basic notions of oscillations and wave packets. Indeed, string theory,
the heir to quantum field theory, is still firmly founded on this harmonic
The aim of this and the following chapter is to explain the basic physical and
mathematical ideas of classical mechanics and quantum mechanics by considering
the so-called harmonic oscillator. In all fields of physics, one encounters oscillating
systems. Let us mention the following examples:
• electromagnetic waves and light (photons);
• laser beams (coherent states);
• oscillating molecules in a gas or a liquid;
• sound waves (phonons);
• oscillations of a crystal lattice (phonons);
• oscillations of a string (e.g., a violin string);
• waves in a plasma (plasmons);
• matter waves of elementary particles (e.g., electrons);
• gravitational waves (gravitons).
The harmonic oscillator represents the simplest oscillating system. The quantization
of the harmonic oscillator is the basis of quantum mechanics, quantum field theory,
and condensed matter physics.
System of physical units. In this chapter, we will use the international system
of units, SI (see the Appendix of Vol. I).

6.1 Prototypes of Extremal Problems
The calculus of variations has its roots in extremal problems for real-valued
functions.
Folklore
The one-dimensional problem. Let f : J → R be a smooth function on the
open interval J. Consider the minimum problem
f(x) = min!, x∈ J. (6.1)
Let us recall some standard results from classical calculus.
(i) Necessary condition for a local minimum: If x0 is a solution of (6.1), then
f'(x0) = 0 and f''(x0) ≥ 0.
(ii) Sufficient condition for a local minimum: If f'(x0) = 0 and f''(x0) > 0, then
the function f has a local minimum at the point x0. This means that there
exists a sufficiently small positive number ε such that f(x) ≥ f(x0) for all
x ∈]x0 − ε, x0 + ε[.
(iii) Sufficient condition for a global minimum: If f'(x0) = 0, f''(x0) > 0, and
f''(x) ≥ 0 on J, then the function f is convex on J, and the minimum problem
(6.1) has the unique solution x0.
Now consider the more general problem
f(x) = critical!, x0 ∈ J. (6.2)
By definition, the point x0 is a solution of (6.2) iff f'(x0) = 0. We say that x0 is
a critical point of f, and the function is critical (or stationary) at the point x0.
Intuitively, the function f is critical at the point x0 iff the tangent line of the graph
of the function f at the point x0 is horizontal (Fig. 6.1). For example, the function
f : R → R given by f(x) := x2 has a global minimum at the point x0 = 0. In
contrast to this, the function f(x) := x3 has the unique critical point x0 = 0,
but no minimal point. The function f(x) := sinx has precisely the critical points
x0 = ±π(n + 1
2) with n = 0, 1, 2, . . .

6.2 The Motion of a Particle

6.3 Newtonian Mechanics
The rise of modern science was accompanied with the replacement of authorities
or traditions by causes in explaining phenomena. One of the ultimate
goals of science is to understand the world, and this is approached by
scientific explanation, that is, by finding out causes for various phenomena.
According to Aristotle, however, there are different kinds of cause:
material, formal, efficient, and final causes. Before the rise of modern science,
teleological explanation based on the notion of final cause was a
dominant mode of explanation. With the revival of Neoplatonism, Archimedianism
and atomism in the Renaissance, there began a transformation
in basic assumptions of scientific explanation. Copernicus, Kepler, Galileo,
and Descartes, for example, believed that the underlying truth and universal
harmony of the world can be perfectly represented by simple and exact
mathematical expressions. The mathematization of nature led to a certain
degree of popularity of formal cause. But the most popular and powerful
conception of causality, in fighting, against the teleological explanation,
was a mechanical one based on the notion of efficient cause. Different from
final and formal causes, the idea of efficient cause focuses on how the cause
is transmitted to the effect, that is, on the mode of transmission. According
to the classical mechanical view, causality can be reduced to the laws
of motion of bodies in space and time. . .
Tian Yu Cao, 1998
Conceptual Developments of 20th Century Field Theories
So that we may say now that the door is opened, for the first time, to a
new method fraught with numerous and wonderful results which in future
years will command the attention of other minds.
Galileo Galilei (1564–1642)
Lex prima: A stationary body will remain motionless, and a moving body
will continue to move in the same direction with unchanging speed unless
it is acted on by some force.
Lex secunda: The time-rate-of-change of the momentum of a body is proportional
to the force.
Lex tertia: If any body exerts a force on another object, then the second
object also exerts an equal and opposite force on the first.
It remains that, from the same principles, I now demonstrate the frame of
the System of the World.
Isaac Newton (1643–1727)
Philosophiae Naturalis Principia Mathematica, London, 16877
Who, by a vigor of mind almost divine, the motions and figures of the
planets, the paths of comets, and the tides of the sea first demonstrated.8
Newton’s Epitaph, Westminster Abbey, London
When one considers all that Newton achieved, and the cultural and scientific
environment in which he achieved it, there is reason to regard him as
the greatest scientist – and perhaps the greatest genius – that ever lived.
Anthony Philip French

6.4 A Glance at the History of the Calculus of
Variations
Bees – by virtue of a certain geometrical forethought – know that the
hexagon is greater than the square and the *** and will hold more
money for the same expenditure of material.
Pappus of Alexandria, 300 B.C.
Every process in nature will occur in the shortest possible way.
Leonardo da Vinci (1452–1519)
A light ray between two points needs the shortest possible time.
Pierre de Fermat (1601–1665)
Johann Bernoulli, professor of mathematics, greets the most sophisticated
mathematicians in the world. Experience shows that noble intellectuals
are driven to work for the pursuit of the knowledge by nothing more than
being confronted with difficult and useful problems.
Six months ago, in the June edition of the Leipzig Acta eruditorum (journal
of scientists), I presented such a problem. The allotted six-month deadline
has now gone by, but no trace of a solution has appeared. Only the famous
Leibniz informed me that he had unravelled the knot of this brilliant and
outstanding problem, and he kindly asked me to extend the deadline until
next Easter. I agreed to this honorable request. . . I will repeat the problem
here once more.
Two points, at different distances from the ground, and not in a vertical
line, should be connected by such a curve that a body under the influence
of gravitational forces passes in the shortest possible way from the upper
to the lower point .
Johann Bernoulli, January 1697
This paper solves my brother’s problem, to whom I will set other problems
in return.
Jakob Bernoulli, May 1697
The Euler Calculus of Variations (Methodus inveniendi) from the year
1744 is one of the most beautiful mathematical works that has ever been
written.
Constantin Carath´eodory (1873–1950)
Read Euler, he is the master of us all.
Marquise de Pierre Simon Laplace (1749–1824)
One needs to have delved but little into the principles of differential calculus
to know the method of how to determine the greatest and least
ordinates of curves. But there are maxima or minima problems of a higher
order, which in fact depend on the same method, which however cannot
be subjected to this method. These are the problems where it is a matter
of finding the curves themselves.
The first problem of this type, which the geometers solved, is that of
the brachistochrone or the curve of fastest fall which Johann Bernoulli
proposed toward the end of the preceding century. One attained this only
in special ways, and it was only some time later and on the occasion of the
investigations concerning isoperimetric problems that the great geometer
of whom we just spoke and his extraordinary brother Jakob Bernoulli gave
some rules in order to solve several other problems of this type.
But since these rules were not of sufficient generality, the famous Euler
undertook to refer all investigations of this type to a general method.  But
even as sophisticated and fruitful as his method is, one must nevertheless
confess that it is not sufficiently simple. . . Now here one finds a method
which requires only a simple use of the principles of differential and integral
calculus.
Joseph Louis Lagrange, 1762
As I see, your analytic solution of the isoperimetric problem contains all
that one can wish for in this situation. I am very happy that this theory
which I have treated since the first attempts almost alone, has been brought
precisely by you to the highest degree of perfection.
The importance of the situation has occasioned me with the help of your
new insights to myself conceive of an analytic solution, but which I shall
not make known before you have published your deliberations, in order
not to deprive you of the least part of the fame due you.
Euler, in a letter to the young Lagrange

6.5 Lagrangian Mechanics
The mathematician is perfect only in so far as he is a perfect being, in so
far as he perceives the beauty of truth; only then will his work be thorough,
transparent, comprehensive, pure, clear, attractive, and even elegant. All
this is necessary in order to resemble Lagrange.
Johann Wolfgang von Goethe (1749–1832)
Wilhelm Meisters Wanderjahre

Jacobi’s Accessory Eigenvalue Problem
Returning to the concepts of maximum and minimum, it is a nuisance that
there reigns such confusion in these words. One says that an expression
attains a maximum or a minimum if one simply wishes to say that it is
critical (or extremal) and hence its first variation vanishes, also in the case
when neither a minimum nor a maximum occurs.
Carl Gustav Jacobi (1804–1851)

6.5.4 The Morse Index
The Morse index describes the global behavior of the action functional of
the harmonic oscillator with respect to arbitrary time intervals. This global
behavior is governed by the appearance of focal points of the trajectories
of the harmonic oscillator, which correspond to focal points in geometric
optics.
Folklore

6.6 Symmetry and Conservation Laws
Newton and his successors noticed that there exist conservation laws that simplify
the integration of the equations of motion. For example, this concerns the
conservation of the following quantities: energy, momentum, angular momentum,
Runge–Lenz vector. In 1918, Emmy Noether (1882–1935) proved a general theorem
which shows that
The symmetries of the Lagrangian are responsible for conservation laws.
For example, invariance of the Lagrangian under time translations leads to conservation
of energy. More general, we will show in Sec. 6.6.2 that smooth continuous
symmetries of the action integral imply conservation laws.

6.7 The Pendulum and Dynamical Systems
Henri Poincar´e (1854–1912) originated not only new theories, but completely
new branches of mathematics like the theory of dynamical systems,
differential topology, and algebraic topology. His ideas are so great,
his way of thinking of and looking at mathematical reality has been so
widely accepted that to us, his descendants, it seems strange that people
have thought differently, for example, that dynamical systems should be
considered on manifolds, and not only on Rn – because indeed a plane
pendulum is a motion on the circle, and a spherical pendulum is a motion
on a sphere.
Krysztof Maurin, 1999
Motions of mass point systems are frequently governed by constraints (e.g., the
oscillations of molecules). Whereas the free motion of N mass points in Euclidean
space has 3N degrees of freedom, constrained motions possess less than 3N degrees
of freedom. The Lagrangian approach to mechanics allows an elegant reduction to
the true number of degrees of freedom by considering the motion with respect to
appropriately chosen local coordinates. This corresponds to the theory of dynamical
systems on manifolds. In order to illustrate this, let us consider the motion x = x(t)
of a pendulum of mass m and length  l.
In a right-handed Cartesian coordinate system, we set x = xi + yj + zk (Fig.
6.9). The gravitational force
F = −mgk
acts on a particle of mass m. Here, g = 9.81m/s2 is the acceleration of gravity.
Letting U(x) := mgz, we get F = −grad U. Therefore, the force F has the
potential U.

6.7.3 The Phase Space of the Pendulum and Bundles
It was a great achievement of Poincar´e (1854–1912) to show that the proper
domain of analytic dynamics is the cotangent bundle TMd of the position
space M – this discovery was so fundamental that nowadays it seems to
be natural and obvious.
Krysztof Maurin, 1996
The categories of differentiable manifolds and vector bundles provide a
useful context for the mathematics needed in mechanics, especially the
new topological and qualitative results.
Ralph Abraham and Jerrold Marsden, 1978
Too often in the physical sciences, the space of states is postulated to be a
linear space when the basic problem is essentially nonlinear; this confuses
the mathematical development.
Stephen Smale, 1980
This section should help the reader to understand the intuitive roots of the language
of bundle theory in modern mathematics and physics. In mechanics, one has to
distinguish the following crucial notions:
• position space M;
• state space TM (tangent bundle of the manifold M);
• phase space TMd (cotangent bundle of M).
Let us discuss this for the prototype of a mechanical system with nontrivial topology
– the circular pendulum.
Our strategy is to introduce only such quantities which have a geometrical
meaning, that is, they are independent of the choice of local coordinates.

Perspective. For a general mechanical system of n degrees of freedom with
n = 1, 2, . . ., the following hold:
• the position space M is a real n-dimensional manifold,
• the state space TM (tangent bundle of M) is a 2n-dimensional manifold, and
• the phase space TMd (cotangent bundle) is a 2n-dimensional symplectic manifold.
The Lagrangian formulation of mechanics is based on the Lagrangian
L : TM → R,
whereas the dual Hamiltonian formulation of mechanics is based on the Hamiltonian
H : TMd → R,
which represents the energy function of the mechanical system, as a rule. In contrast
to the cotangent bundle TMd, the tangent bundle TM is not always a symplectic
manifold. Therefore, the Hamiltonian formulation of mechanics has advantages over
the Lagrangian formulation. In statistical mechanics, the Hamiltonian approach will
allow us to use the volume measure of the phase space in order to construct the key
probability measure (see Sect. 7.17.5 on page 645). Summarizing, the language of
manifolds allows us to study the global aspects of the motion of mechanical systems.
The complexity of a mechanical system is reflected by the complex topology
of the tangent bundle and the cotangent bundle.

6.8 Hamiltonian Mechanics
The Hamiltonian approach to mechanics is centered at the concept of energy. In
this setting, the following two important results are quite natural:
• Conservation of energy, and
• conservation of phase volume.
The latter property is crucial for classical statistical physics (Gibbs measure). From
the geometric point of view, the passage from Lagrangian mechanics to Hamiltonian
mechanics corresponds to a passage
• from the tangent bundle TM of the position space M (position, velocity vector)
• to the dual cotangent bundle TMd of M (position, differential form).
This transformation is called the Legendre transformation which is a contact transformation
in the sense of Lie. The cotangent bundle TMd always carries a natural
symplectic structure. Note the following:
• Lagrangian mechanics corresponds to Riemannian geometry (the metric is given
by the kinetic energy).
• Hamiltonian mechanics corresponds to symplectic geometry.
One of the great problems of mathematics and physics in the 19th century consisted
in solving the N-body problem in celestial mechanics.
The aim was to investigate the stability of our solar system.
To simplify considerations, Jacobi and his successors used transformations of time
and position which preserve the form of the canonical equations.
Such transformations are called canonical transformations.
It turns out that canonical transformations coincide with symplectic transformations.
That is, canonical transformations preserve the symplectic structure, and
hence they present the symmetry transformations of symplectic geometry. When
creating his mechanics, Hamilton (1805–1865) was motivated by an analogy between
mechanics and Huygens’ geometrical optics created in the 17th century:
• The trajectories q = q(t), p = p(t) of a particle in mechanics correspond to light
rays (solutions of the canonical ordinary differential equations), and
• the action function S = S(q, t) in mechanics corresponds to the eikonal function
in geometric optics, which determines the wave fronts (solutions of the Hamilton–
Jacobi partial differential equation).
• The duality between trajectories and wave fronts was fully established by
Carath´eodory in the framework of his ’royal road to the calculus of variations’
in 1925. This is based on the fundamental notion of geodesic fields.
The relation between geometric optics and wave optics plays a fundamental role
for understanding Schr¨odinger’s quantum mechanics which he discovered in 1926.
In the 1950s, optimal control theory was invented independently by Bellman and
Pontryagin. This theory allows many applications in technology (e.g., moon landing
and return of a spaceship to earth):
• Pontryagin’s theory generalizes Hamilton’s canonical equations (Pontryagin’s
maximum principle and light rays), whereas
• Bellman’s theory of dynamic programming generalizes the Hamilton–Jacobi partial
equation (the Hamilton–Jacobi–Bellman equation and wave fronts).
In the Bellman approach, the action function S corresponds to the cost function.
A detailed study can be found in Zeidler (1986), Vol. III (see the references on
page 1049). In what follows, let us study the main ideas on an elementary level by
considering the harmonic oscillator as a prototype.

6.9 Poissonian Mechanics
The Poissonian formulation of mechanics has the following two advantages:
• Conservation laws can be expressed in terms of Poisson brackets.
• Quantum mechanics can be obtained from classical mechanics by replacing the
Poisson bracket {A,B} with the commutator
1/ih· [A,B]−
where the Lie bracket [A,B]− is equal to AB−BA. This was implicitly discovered
by Heisenberg in 1925. The general quantization rule was formulated by Dirac
in 1926.

6.10 Symplectic Geometry
A deeper understanding of classical mechanics is based on symplectic geometry.
By definition, a quantity belongs to symplectic geometry iff it is invariant under
symplectic transformations. Let us explain the basic ideas.

6.11 The Spherical Pendulum
Two-dimensional spheres are the simplest curved su***ces. They serve as
prototypes for the geometry and analysis of manifolds.
Folklore

6.13.2 Lagrangian Submanifolds in Symplectic Geometry
In geometrical optics, one wants to construct wave fronts by means of families of
light rays. The point is that only special families of light rays allow this construction.
This is intimately related to the notion of Lagrange brackets, which were introduced
by Lagrange in the 18th century in order to simplify computations in celestial mechanics
in the framework of perturbation theory. The point is that the Lagrange
brackets of a family of light rays are constant in time along the light rays (i.e.,
they are first integrals of the Hamilton canonical equations). In modern symplectic
geometry, the Lagrange brackets are reformulated as Lagrangian submanifolds of
a symplectic manifold. Replacing light rays by trajectories of particles in classical
mechanics, we will study
• the construction of a solution of the Hamilton–Jacobi partial differential equation
• by the help of suitable families of trajectories which satisfy the Hamilton canonical
system of ordinary differential equations.


由一星于2014-09-19, 03:34进行了最后一次编辑,总共编辑了7次
一星
一星

帖子数 : 3787
注册日期 : 13-08-07

返回页首 向下

Quantum Field Theory II - 页 4 Empty 回复: Quantum Field Theory II

帖子 由 一星 2014-09-13, 04:19

[ltr]

[size=35]分类“橢圓函數”中的页面

本分类中包含以下9个页面,共9个页面。
[/ltr][/size]

A


G


T


W


Θ



椭 续









椭圆积分[编辑]





[ltr]在积分学中,椭圆积分最初出现于椭圆弧长有关的问题中。Guilio Fagnano欧拉是最早的研究者。现代数学将椭圆积分定义为可以表达为如下形式的任何函数 Quantum Field Theory II - 页 4 18f63800376271ee4b0efe1545744cd6的积分
Quantum Field Theory II - 页 4 8341d1326b0ddd46ba7b80c83ad0c02f
其中Quantum Field Theory II - 页 4 153fc2a5a0a49d52dda62d96ae0a293f是其两个参数的有理函数Quantum Field Theory II - 页 4 8a140337171d690f8dd0eebd94448bf0是一个无重根的Quantum Field Theory II - 页 4 670ba4ac59ff057abf9a9714ea1523d7Quantum Field Theory II - 页 4 D611e96a4df456427237ae8f7bb44c73多项式的平方根,而Quantum Field Theory II - 页 4 08163b03d3a58471d7f88fc4e581a282是一个常数。
通常,椭圆积分不能用基本函数表达。这个一般规则的例外出现在Quantum Field Theory II - 页 4 8a140337171d690f8dd0eebd94448bf0有重根的时候,或者是Quantum Field Theory II - 页 4 153fc2a5a0a49d52dda62d96ae0a293f,Quantum Field Theory II - 页 4 43bed6d55f204dfc7251cacf37f11c91没有Quantum Field Theory II - 页 4 Ec9ff0a12771e750c2685d3b89a37c79的奇数幂时。但是,通过适当的简化公式,每个椭圆积分可以变为只涉及有理函数和三个经典形式的积分。(也即,第一,第二,和第三类的椭圆积分)。
除下面给出的形式之外,椭圆积分也可以表达为勒让德形式Carlson对称形式。通过对施瓦茨-克里斯托费尔映射的研究可以加深对椭圆积分理论的理解。历史上,椭圆函数是作为椭圆积分的逆函数被发现的,特别是这一个:Quantum Field Theory II - 页 4 4d2cdf056b6c49751170145fec650e78其中Quantum Field Theory II - 页 4 349b7a418ed90f5397b78a263ea8a3ac雅可比椭圆函数之一。[/ltr]



[ltr]

目录



  [隐藏[/ltr]




[ltr]

记法[编辑]



椭圆积分通常表述为不同变量的函数。这些变量完全等价(它们给出同样的椭圆积分),但是它们看起来很不相同。很多文献使用单一一种标准命名规则。在定义积分之前,先来检视一下这些变量的命名常规:[/ltr]



[ltr]
上述三种常规完全互相确定。规定其中一个和规定另外一个一样。椭圆积分也依赖于另一个变量,可以有如下几种不同的设定方法:[/ltr]



[ltr]
规定其中一个决定另外两个。这样,它们可以互换地使用。注意Quantum Field Theory II - 页 4 61efad693efe8e0ffd7d7bc042b427ef也依赖于Quantum Field Theory II - 页 4 79dd9720ffa5bbe026e23afc9ab4df3c。其它包含Quantum Field Theory II - 页 4 61efad693efe8e0ffd7d7bc042b427ef的关系有
Quantum Field Theory II - 页 4 26b4239f521b89a498f9278548f132b4

Quantum Field Theory II - 页 4 198eee8f6e800cff7f0a0fc94090b5b3
后者有时称为δ幅度并写作Quantum Field Theory II - 页 4 819c75cb0b597815150c36586bf1614f。有时文献也称之为补参数,补模或者补模角。这些在四分周期中有进一步的定义。

第一类不完全椭圆积分[编辑]



第一类不完全椭圆积分 Quantum Field Theory II - 页 4 Bc352fc10ca296a872b51d91a1132127定义为
Quantum Field Theory II - 页 4 541503a9bdd9d601332f66354277235e
与此等价,用雅可比的形式,可以设 Quantum Field Theory II - 页 4 80f8b79e5fb2210c77e6ec4be7d8449a;则
Quantum Field Theory II - 页 4 7493458d889c6f7b06585c0f1e6386b8
其中,假定任何有竖直条出现的地方,紧跟竖直条的变量是(如上定义的)参数;而且,当反斜杠出现的时候,跟着出现的是模角。 在这个意义下,Quantum Field Theory II - 页 4 27d22a4a503885e9202959fe515e6d28,这里的记法来自标准参考书Abramowitz and Stegun。使用限界符;Quantum Field Theory II - 页 4 25f2c517edfb6f7d8f3f2d10650c20f6 Quantum Field Theory II - 页 4 Fa0d646a4f60f9fec797957680ad6430是椭圆积分中的传统做法。
但是,还有许多不同的常规用于椭圆积分的记法。取值为椭圆积分的函数没有(象平方根正弦误差函数那样的)标准和唯一的名字。甚至关于该领域的文献也常常采用不同的记法。Gradstein, Ryzhik[1]Quantum Field Theory II - 页 4 4af5d4b3507909ab404709274db145d6.(8.111)]采用Quantum Field Theory II - 页 4 1ab2381bd99ac12b47a028dee343e106。该记法和这里的Quantum Field Theory II - 页 4 153b4ef667caa10e3c552677934f6193;以及下面的Quantum Field Theory II - 页 4 8a34f21acb2325cde8cd6395aab2c59c等价。
和上面的不同对应的是,如果从Mathematica语言翻译代码到Maple语言,必须将EllipticK函数的参数用它的平方根代替。反过来,如果从Maple翻到Mathematica,则参数应该用它的平方代替。Maple中的EllipticK(Quantum Field Theory II - 页 4 9dd4e461268c8034f5c8564e155c67a6)几乎和Mathematica中的EllipticK[Quantum Field Theory II - 页 4 C66452631491acdbf8e5ed69dfd19681]相等;至少当Quantum Field Theory II - 页 4 7cbced71029dbc2a89e0d8127b8245ec时是相等的。
注意
Quantum Field Theory II - 页 4 32b6a5fdd1ebc774c852250254319fb7
其中Quantum Field Theory II - 页 4 61efad693efe8e0ffd7d7bc042b427ef如上文所定义:由此可见,雅可比椭圆函数是椭圆积分的逆。

加法公式[编辑]



Quantum Field Theory II - 页 4 1a9802a6713d65ab12d3d059251280a3
此公式成立是有条件的。参见《第一、二类椭圆积分加法公式的成立条件》http://zuijianqiugen.blog.163.com/blog/static/126524062201422111572376/

性质[编辑]



Quantum Field Theory II - 页 4 943e2a2a34bfefe344c55a5a0bed2f5bQuantum Field Theory II - 页 4 Bbf8cc27a8c25de831c2341c8e76d5dbQuantum Field Theory II - 页 4 6d9f2af69416beca65d1537a530c6bf3Quantum Field Theory II - 页 4 B4a3fd65b014d6d39778efba6555b275Quantum Field Theory II - 页 4 B098d8cb0089ff9af21cc6e2c3df0f1eQuantum Field Theory II - 页 4 Eb24bcf124a00150705d3b938913ee95Quantum Field Theory II - 页 4 08294ebeb0cf7eccd35f4cfb240871acQuantum Field Theory II - 页 4 6646c4fa12baff406b105059fb929641

第一类不完全椭圆积分的导数[编辑]



Quantum Field Theory II - 页 4 7b6a2a256327c2f220282aa1be9c4274Quantum Field Theory II - 页 4 599a6cc55f534460a7b4365ef172ab1d

第二类不完全椭圆积分[编辑]



第二类不完全椭圆积分 Quantum Field Theory II - 页 4 8e04ea2fbed21e24f37b273140d25ad4
Quantum Field Theory II - 页 4 37cb0839b30642482ac04b346a52b6a7
与此等价,采用另外一个记法(作变量替换Quantum Field Theory II - 页 4 Fa645dbaf2fdf0eea331095a02a1112b),
Quantum Field Theory II - 页 4 57558adaeb21176d72b6fccb269d7fbe
其它关系包括
Quantum Field Theory II - 页 4 7973c47985a52848eee36417ff2008e5Quantum Field Theory II - 页 4 9adca1f945cda816ff91104578209027

加法公式[编辑]



Quantum Field Theory II - 页 4 903ca45ec8d974f56fea249e98078f4bQuantum Field Theory II - 页 4 54717bca498287f5aafc99c4e0de8889
此公式成立是有条件的。参见《第一、二类椭圆积分加法公式的成立条件》http://zuijianqiugen.blog.163.com/blog/static/126524062201422111572376/

性质[编辑]



Quantum Field Theory II - 页 4 119bec12142fbce5f78073e7c485ba1bQuantum Field Theory II - 页 4 E6b466705689d8638c898749ffbbd34e

第二类不完全椭圆积分的导数[编辑]



Quantum Field Theory II - 页 4 5c0919ef0af63c6af2ee6ba5337c4fcdQuantum Field Theory II - 页 4 852f18547394365a03a6b8eab066de53Quantum Field Theory II - 页 4 630d7ec30e089f816c40524e046fbb47

第三类不完全椭圆积分[编辑]



第三类不完全椭圆积分Quantum Field Theory II - 页 4 6c16c9c351f4d98fb70fe3a6f1963b02
Quantum Field Theory II - 页 4 1499aafee6314d0ebcec0a0ffc22dca4
或者
Quantum Field Theory II - 页 4 6cf0821e27605da8e8ff593caaa95eaa
或者
Quantum Field Theory II - 页 4 214d28dddf1fb7f4f2961d7437714fe7
数字Quantum Field Theory II - 页 4 A957404c96e59f1746f97ab668c8e1f8称为特征数,可以取任意值,和其它参数独立。但是要注意Quantum Field Theory II - 页 4 943daf84f9af8a89422bc07dd00e07b5对于任意Quantum Field Theory II - 页 4 D4f5579278053dcc711fa0e6e45244fa是无穷的。

加法公式[编辑]



Quantum Field Theory II - 页 4 432567e822457212d4dd50b7709e2b82

第三类不完全椭圆积分的导数[编辑]



Quantum Field Theory II - 页 4 86b87e1421096c9dc719e3074ab0d0b5Quantum Field Theory II - 页 4 0f681a2bd79fa34b53138e47cc7f0c98Quantum Field Theory II - 页 4 7d887ebcc6a8a6d8f61d4b2f3e25d160Quantum Field Theory II - 页 4 759cdbbe0282ab36b83c0e27ed1fd575

特殊值[编辑]



Quantum Field Theory II - 页 4 9b794ba394f8e050b1ecd7d9c64c94f1Quantum Field Theory II - 页 4 99e75ce1445fcc8c00f1d1da25cacb9bQuantum Field Theory II - 页 4 1959815de053d03ddaae65ef2be3b378Quantum Field Theory II - 页 4 547172581564ce3013c3054dbee04ca3Quantum Field Theory II - 页 4 99e75ce1445fcc8c00f1d1da25cacb9bQuantum Field Theory II - 页 4 A12395cf4d43ba14874bc773fd355058Quantum Field Theory II - 页 4 A5e962bfac2577c2f82963a8837f0cd6Quantum Field Theory II - 页 4 72685e9657c5fefc883d6efcf6edebc9

第一类完全椭圆积分[编辑]



如果幅度为Quantum Field Theory II - 页 4 51a0d168b7a1bb8e3bb8f4983049d4a2或者Quantum Field Theory II - 页 4 C744814ed171fdf17ec17b589337b525,则称椭圆积分为完全的。 第一类完全椭圆积分Quantum Field Theory II - 页 4 D0e1b8571128845c03a4cfac00d43b66可以定位为
Quantum Field Theory II - 页 4 67f9ec4c5ef1ee9bf0dd6fe3d49bfa7a
或者
Quantum Field Theory II - 页 4 19a1fb56b7d84178107644b5f4b462e3
它是第一类不完全椭圆积分的特例:
Quantum Field Theory II - 页 4 7ff88a21b49c987da45e296a9922133f
这个特例可以表达为幂级数
Quantum Field Theory II - 页 4 F7ef1dc81de81c7bddbbf5149e424e43
它等价于
Quantum Field Theory II - 页 4 5a3f0d2d1e308865af0730f877bd5666
其中Quantum Field Theory II - 页 4 8519304401464f2492f48988c9931bcc表示双阶乘。采用高斯的超几何函数,第一类完全椭圆积分可以表达为
Quantum Field Theory II - 页 4 E529e697aebbeb8733fcd81d9cdb2d20
第一类完全椭圆积分有时称为四分周期。它可以采用算术几何平均值计算。

复数值[编辑]



Quantum Field Theory II - 页 4 9d59026f8e7a00f1b1470922d73dfcabQuantum Field Theory II - 页 4 415389075f27b11f8b8e3383b2fa5d30

特殊值[编辑]



Quantum Field Theory II - 页 4 8e5b8e89d0cfcf330045a83b33b11e7dQuantum Field Theory II - 页 4 D9bf52f54dd3d5f8fbf8ecf5b78c09adQuantum Field Theory II - 页 4 0c0ddfa8d07401f038991a745fcf9c26Quantum Field Theory II - 页 4 46924583835dff61081b5155cb6a2fb3Quantum Field Theory II - 页 4 108a777e588936c99a3f096bbbdb6aa1Quantum Field Theory II - 页 4 32d3dd97971020bf52d33fbb534e9378Quantum Field Theory II - 页 4 8767162f3a5f1b5b59be37b5a8074861Quantum Field Theory II - 页 4 5d7c66ee5b6dd40d6e0272167bd69233Quantum Field Theory II - 页 4 6059a3ef64ff20fb7bd05ecbda80fe68Quantum Field Theory II - 页 4 2ad28b40896d1e1b2878961d1893bca3Quantum Field Theory II - 页 4 4227ae70f297a9c04943650a67164bdd
其中
Quantum Field Theory II - 页 4 E3c8a35fe8d6ec004ccd7f2d362d58feQuantum Field Theory II - 页 4 B74beba6b00e39b138f92cd8630fa52c
第一类完全椭圆积分满足
Quantum Field Theory II - 页 4 Dea8d2dc2fb8fe210a94547fc35ac034

第一类完全椭圆积分的导数[编辑]



Quantum Field Theory II - 页 4 597d8bbd5bbfc832dede7f4699fc3531

第二类完全椭圆积分[编辑]



第二类完全椭圆积分 Quantum Field Theory II - 页 4 4b88f47f80273fd5788e1e20aa81c38a可以定义为
Quantum Field Theory II - 页 4 Df5634731743e429875d20c8a8a9971d
或者
Quantum Field Theory II - 页 4 Feae0e6445b83e6b7a8be3c3a9712a13
它是第二类不完全椭圆积分的特殊情况:
Quantum Field Theory II - 页 4 820104ebdff332ac2ef6b3830a1da60a
它可以用幂级数表达
Quantum Field Theory II - 页 4 8da804c4cce41a5ffe9f83b23b5678b2
也就是
Quantum Field Theory II - 页 4 F1a9328c5213fdc7e9947b9bcc998913
高斯超几何函数表示的话,第二类完全椭圆积分可以写作
Quantum Field Theory II - 页 4 44fa0b14007bf85ae2890cc0e38af1fb
有如下性质
Quantum Field Theory II - 页 4 Dbf4d36d1ae6f018a6a558ceaaa8e591Quantum Field Theory II - 页 4 6d9f2af69416beca65d1537a530c6bf3

复数值[编辑]



Quantum Field Theory II - 页 4 11dd80c52e85fe72715c3ffb1a5628b2

特殊值[编辑]



Quantum Field Theory II - 页 4 D2ec2e3bd745baee33be9d612232a396Quantum Field Theory II - 页 4 Ab3ceae68e66ec40504230cc8ba3c826Quantum Field Theory II - 页 4 697eb8cdac9f15f467cf3058931f9c43Quantum Field Theory II - 页 4 158520e21a37f6f3d62639bebe9ec745Quantum Field Theory II - 页 4 42f90751e0abf89bc28852890f4093a8Quantum Field Theory II - 页 4 630c4434373ea6c6a6c9dc4a8cad829dQuantum Field Theory II - 页 4 Ba0c6634756d863a95c81e439cf796f0Quantum Field Theory II - 页 4 7f6abd4980f298c7a5d7c14a065ec9afQuantum Field Theory II - 页 4 48b6f2359e660c1205dbe054e5fcf57cQuantum Field Theory II - 页 4 13869a2059c9989ed15faea7f8a8d399Quantum Field Theory II - 页 4 1998c6d203aeead59d83dabb48cdc79eQuantum Field Theory II - 页 4 5babecbb12c8d9d6db42ec2a8c3093de

其中
Quantum Field Theory II - 页 4 D6e0418c1abe1f19449a682b25ee8212Quantum Field Theory II - 页 4 61dfe74ae4bdebc8170b8d690488133aQuantum Field Theory II - 页 4 03733cb4fe4e8be8ea88a58108ffe3b4Quantum Field Theory II - 页 4 7e98c534c54de312218e444d0e18c9b9

第二类完全椭圆积分的导数和不定积分[编辑]



Quantum Field Theory II - 页 4 3e9985cf61ca1c434ef2606e8168875bQuantum Field Theory II - 页 4 B1df1392f47b511785cf33f0e522f0c5

第三类完全椭圆积分[编辑]



第三类完全椭圆积分Quantum Field Theory II - 页 4 206932cf05272e53cc4b051c110e58f0可以定义为
Quantum Field Theory II - 页 4 4b6625c5f992f449bc24aa23568d7448
注意有时第三类椭圆积分被定义为带相反符号的Quantum Field Theory II - 页 4 A957404c96e59f1746f97ab668c8e1f8,也即
Quantum Field Theory II - 页 4 1dff291ee4c5a3ccb74d98a10c2dfc80
用Appell hypergeometric Function表示为
Quantum Field Theory II - 页 4 Babfe579b6fc3c63beaf5fb19c54b20e
第三类完全椭圆积分和第一类椭圆积分之间的关系
Quantum Field Theory II - 页 4 E04c75723ca7157f5333cea17e9204a5
Quantum Field Theory II - 页 4 9abc3c60517a3ea9213ca7363bf60b6d

Quantum Field Theory II - 页 4 D1eadb6e46c0795c02bba9c80f18f208
Quantum Field Theory II - 页 4 6e9ff28c498cdc99811b4e029be92e5d

第三类完全椭圆积分的导数[编辑]



Quantum Field Theory II - 页 4 04eba29438824e83fdd8558b70e42354Quantum Field Theory II - 页 4 B5f25e8511f93db8377d0d47f55a838f

特殊值[编辑]



Quantum Field Theory II - 页 4 53a48f6ef4641924e7c45683ebeff62fQuantum Field Theory II - 页 4 5aeb2c40f33e504e5f5ad2956578ca7fQuantum Field Theory II - 页 4 35164667e34bbce2d667ec45cc4a59f0Quantum Field Theory II - 页 4 5cec09117f5ca3be58563bb22e26a9ebQuantum Field Theory II - 页 4 56a74cc9d155ea2e6f14145879a090f1Quantum Field Theory II - 页 4 6d7189578a6609d5505dcc17b0c216c2Quantum Field Theory II - 页 4 79e2d541f266ff5643f27d18f3f02d1b

参看[编辑]

[/ltr]



[ltr]

参考[编辑]

[/ltr]



一星
一星

帖子数 : 3787
注册日期 : 13-08-07

返回页首 向下

Quantum Field Theory II - 页 4 Empty 回复: Quantum Field Theory II

帖子 由 一星 2014-09-19, 06:08

7. Quantization of the Harmonic Oscillator –
Ariadne’s Thread in Quantization
Whoever understands the quantization of the harmonic oscillator can understand
everything in quantum physics.
Folklore
Almost all of physics now relies upon quantum physics. This theory was
discovered around the beginning of this century. Since then, it has known
a progress with no analogue in the history of science, finally reaching a
status of universal applicability.
The radical novelty of quantum mechanics almost immediately brought a
conflict with the previously admitted corpus of classical physics, and this
went as far as rejecting the age-old representation of physical reality by
visual intuition and common sense. The abstract formalism of the theory
had almost no direct counterpart in the ordinary features around us, as,
for instance, nobody will ever see a wave function when looking at a car
or a chair. An ever-present randomness also came to contradict classical
determinism.
Roland Omn`es, 1994
Quantum mechanics deserves the interest of mathematicians not only because
it is a very important physical theory, which governs all microphysics,
that is, the physical phenomena at the microscopic scale of 10
−10m, but
also because it turned out to be at the root of important developments of
modern mathematics.
Franco Strocchi, 2005
In this chapter, we will study the following quantization methods:
• Heisenberg quantization (matrix mechanics; creation and annihilation operators),
• Schr¨odinger quantization (wave mechanics; the Schr¨odinger partial differential
equation),
• Feynman quantization (integral representation of the wave function by means of
the propagator kernel, the formal Feynman path integral, the rigorous infinitedimensional
Gaussian integral, and the rigorous Wiener path integral),
• Weyl quantization (deformation of Poisson structures),
• Weyl quantization functor from symplectic linear spaces to C

-algebras,
• Bargmann quantization (holomorphic quantization),
• supersymmetric quantization (fermions and bosons).
We will choose the presentation of the material in such a way that the
reader is well prepared for the generalizations to quantum field theory to
be considered later on.
Formally self-adjoint operators. The operator A : D(A) → X on the complex
Hilbert space X is called formally self-adjoint iff the operator is linear, the domain
of definition D(A) is a linear dense subspace of the Hilbert space X, and we have
the symmetry condition
<χ|Aϕ >=< Aχ|ϕ> for all χ, ψ ∈ D(A).
Formally self-adjoint operators are also called symmetric operators. The following
two observations are crucial for quantum mechanics:
• If the complex number λ is an eigenvalue of A, that is, there exists a nonzero
element ϕ ∈ D(A) such that Aϕ = λϕ, then λ is a real number. This follows
from λ =< ϕ|Aϕ> =< Aϕ|ϕ >= λ†.
• If λ1 and λ2 are two different eigenvalues of the operator A with eigenvectors ϕ1
and ϕ2, then ϕ1 is orthogonal to ϕ2. This follows from
(λ1 − λ2)<ϕ1|ϕ2 >= <Aϕ1|ϕ2> − <ϕ1|Aϕ2> = 0.
In quantum mechanics, formally self-adjoint operators represent formal observables.
For a deeper mathematical analysis, we need self-adjoint operators, which
are called observables in quantum mechanics.
Each self-adjoint operator is formally self-adjoint. But, the converse is not true. For
the convenience of the reader, on page 683 we summarize basic material from functional
analysis which will be frequently encountered in this chapter. This concerns
the following notions: formally adjoint operator, adjoint operator, self-adjoint operator,
essentially self-adjoint operator, closed operator, and the closure of a formally
self-adjoint operator. The reader, who is not familiar with this material, should
have a look at page 683. Observe that, as a rule, in the physics literature one does
not distinguish between formally self-adjoint operators and self-adjoint operators.
Peter Lax writes:
The theory of self-adjoint operators was created by John von Neumann to
fashion a framework for quantum mechanics. The operators in Schr¨odinger’s
theory from 1926 that are associated with atoms and molecules
are partial differential operators whose coefficients are singular at certain
points; these singularities correspond to the unbounded growth of the force
between two electrons that approach each other. . . I recall in the summer
of 1951 the excitement and elation of von Neumann when he learned that
Kato (born 1917) has proved the self-adjointness of the Schr¨odinger operator
associated with the helium atom.
And what do the physicists think of these matters? In the 1960s Friedrichs5
met Heisenberg and used the occasion to express to him the deep gratitude
of the community of mathematicians for having created quantum mechanics,
which gave birth to the beautiful theory of operators in Hilbert space.
Heisenberg allowed that this was so; Friedrichs then added that the mathematicians
have, in some measure, returned the favor. Heisenberg looked
noncommittal, so Friedrichs pointed out that it was a mathematician, von
Neumann, who clarified the difference between a self-adjoint operator and
one that is merely symmetric.“What’s the difference,” said Heisenberg.
As a rule of thumb, a formally self-adjoint (also called symmetric) differential operator
can be extended to a self-adjoint operator if we add appropriate boundary
conditions. The situation is not dramatic for physicists, since physics dictates the
‘right’ boundary conditions in regular situations. However, one has to be careful.
In Problem 7.19, we will consider a formally self-adjoint differential operator which
cannot be extended to a self-adjoint operator.
The point is that self-adjoint operators possess a spectral family which allows
us to construct both the probability measure for physical observables
and the functions of observables (e.g., the propagator for the quantum dynamics).
In general terms, this is not possible for merely formally self-adjoint operators.
The following proposition displays the difference between formally self-adjoint and
self-adjoint operators.
Proposition 7.1 The linear, densely defined operator A : D(A) → X on the complex
Hilbert space X is self-adjoint iff it is formally self-adjoint and it always follows
from
<ψ|Aϕ >=< χ|ϕ>
for fixed ψ, χ ∈ X and all ϕ ∈ D(A) that ψ ∈ D(A).
Therefore, the domain of definition D(A) of the operator A plays a critical role.

7.1 Complete Orthonormal Systems
A complete orthonormal system of eigenstates of an observable (e.g., the
energy operator) cannot be extended to a larger orthonormal system of
eigenstates.
Folklore
Basic question. Let H : D(H) → X be a formally self-adjoint operator on the
infinite-dimensional separable complex Hilbert space X. Physicists have invented
algebraic methods for computing eigensolutions of the form
Hϕn = Enϕn, n= 0, 1, 2, . . . (7.1)
The idea is to apply so-called ladder operators which are based on the use of commutation
relations (related to Lie algebras or super Lie algebras).We will encounter
this method several times. In terms of physics, the operator H describes the energy
of the quantum system under consideration. Here, the real numbers E0,E1,E2, . . .
are the energy values, and ϕ0, ϕ1, ϕ2, . . . are the corresponding energy eigenstates.
Suppose that ϕ0, ϕ1, ϕ2, . . . is an orthonormal system, that is,
<ϕk|ϕn> = δkn, k,n= 0, 1, 2, . . .
There arises the following crucial question.

7.2 Bosonic Creation and Annihilation Operators
Whoever understands creation and annihilation operators can understand
everything in quantum physics.
Folklore

7.3 Heisenberg’s Quantum Mechanics
Quantum mechanics was born on December 14, 1900, when Max Planck
delivered his famous lecture before the German Physical Society in Berlin
which was printed afterwards under the title “On the law of energy distribution
in the normal spectrum.” In this paper, Planck assumed that the
emission and absorption of radiation always takes place in discrete portions
of energy or energy quanta hν, where ν is the frequency of the emitted or
absorbed radiation. Starting with this assumption, Planck arrived at his
famous formula

for the energy density of black-body radiation at temperature T.
Barthel Leendert van der Waerden, 1967
The present paper seeks to establish a basis for theoretical quantum mechanics
founded exclusively upon relationships between quantities which
in principle are observable.
Werner Heisenberg, 1925
The recently published theoretical approach of Heisenberg is here developed
into a systematic theory of quantum mechanics with the aid of mathematical
matrix theory. After a brief survey of the latter, the mechanical
equations of motions are derived from a variational principle and it is
shown that using Heisenberg’s quantum condition, the principle of energy
conservation and Bohr’s frequency condition follow from the mechanical
equations. Using the anharmonic oscillator as example, the question of
uniqueness of the solution and of the significance of the phases of the
partial vibrations is raised. The paper concludes with an attempt to incorporate
electromagnetic field laws into the new theory.
Max Born and Pascal Jordan, 1925
There exist three different, but equivalent approaches to quantum mechanics,
namely,
(i) Heisenberg’s particle quantization from the year 1925 and its refinement by
Born, Dirac, and Jordan in 1926,
(ii) Schr¨odinger’s wave quantization from 1926, and
(iii) Feynman’s statistics over classical paths via path integral from 1942.
In what follows we will thoroughly discuss these three approaches in terms of the
harmonic oscillator. Let us start with (i).

7.3.1 Heisenberg’s Equation of Motion
In a recent paper, Heisenberg puts forward a new theory which suggests
that it is not the equations of classical mechanics that are in any way at
fault, but that the mathematical operations by which physical results are
deduced from them require modification. All the information supplied by
the classical theory can thus be made use of in the new theory . . . We make
the fundamental assumption that the difference between the Heisenberg
products is equal to i times their Poison bracket
xy − yx = i{x, y}. (7.18)
It seems reasonable to take (7.18) as constituting the general quantum
conditions.
Paul Dirac, 1925
The general quantization principle.We are looking for a simple principle which
allows us to pass from classical mechanics to quantum mechanics. This principle
reads as follows:
• position q(t) and momentum p(t) of the particle at time t become operators,
• and Poisson brackets are replaced by Lie brackets,
{A(q, p),B(q, p)} ⇒ 1/ih[A(q, p),B(q, p)]−.
Recall that [A,B]− := AB − BA. Using this quantization principle, the classical
equation of motion (7.17) passes over to the equation of motion for the quantum
harmonic oscillator
ihq˙(t) = [q(t),H(q(t), p(t))]−,
ihp˙(t) = [p(t),H(q(t), p(t))]−                                                            (7.19)
together with
[q(t), p(t)]− = ihI. (7.20)
The latter equation is called the Heisenberg–Born–Jordan commutation relation.

The famous Heisenberg uncertainty inequality for the quantum harmonic oscillator
tells us that the state ϕn has the sharp energy En, but it is impossible
to measure sharply both position and momentum of the quantum particle at the
same time. Thus, there exists a substantial difference between classical particles
and quantum particles.
It is impossible to speak of the trajectory of a quantum particle.

7.3.3 Quantization of Energy
I have the best of reasons for being an admirer of Werner Heisenberg.
He and I were young research students at the same time, about the same
age, working on the same problem. Heisenberg succeeded where I failed. . .
Heisenberg - a graduate student of Sommerfeld - was working from the
experimental basis, using the results of spectroscopy, which by 1925 had
accumulated an enormous amount of data20. . .
Paul Dirac, 1968
The measured spectrum of an atom or a molecule is characterized by two quantities,
namely,
• the wave length λnm of the emitted spectral lines (where n,m = 0, 1, 2, . . . with
n > m), and
• the intensity of the spectral lines.
In Bohr’s and Sommerfeld’s semi-classical approach to the spectra of atoms and
molecules from the years 1913 and 1916, respectively, the spectral lines correspond
to photons which are emitted by jumps of an electron from one orbit of the atom or
molecule to another orbit. If E0 < E1 < E2 < .. . are the (discrete) energies of the
electron corresponding to the different orbits, then a jump of the electron from the
higher energy level En to the lower energy level Em produces the emission of one
photon of energy En − Em. According to Einstein’s light quanta hypothesis from
1905, this yields the frequency
νnm =(En − Em)/h, n>m (7.26)
of the emitted photon, and hence the wave length λnm = c/νnm of the corresponding
spectral line is obtained. The intensity of the spectral lines depends on the transition
probabilities for the jumps of the electrons. In 1925 it was Heisenberg’s philosophy
to base his new quantum mechanics only on quantities which can be measured in
physical experiments, namely,
• the energies E0,E1, . . . of bound states and
• the transition probabilities for changing bound states.

7.3.5 The Wightman Functions
Both the Wightman functions and the correlation functions of the quantized
harmonic oscillator are the prototypes of general constructions used
in quantum field theory.
Folklore

We define the n-point
Wightman function of the quantized harmonic oscillator by setting
Wn(t1, t2, . . . , tn) := 0|q(t1)q(t2) · · · q(tn)|0 (7.32)
for all times t1, t2, . . . , tn ∈ R. This is the vacuum expectation value of the operator
product q(t1)q(t2) · · · q(tn). In contrast to the operator function (7.31), the
Wightman functions are classical complex-valued functions. It turns out that
The Wightman functions know all about the quantized harmonic oscillator.
Using the Wightman functions, we avoid the use of operator theory in Hilbert space.
This is the main idea behind the introduction of the Wightman functions.

Perspectives. In 1956 Wightman showed that it is possible to base quantum
field theory on the investigation of the vacuum expectation values of the products of
quantum fields. These vacuum expectation values are called Wightman functions.
The crucial point is that the Wightman functions are highly singular objects in
quantum field theory. In fact, they are generalized functions. However, they are
also boundary values of holomorphic functions of several complex variables. This
simplifies the mathematical theory. Using a similar construction as in the proof
of the Gelfand–Naimark–Segal (GNS) representation theorem for C∗-algebras in
Hilbert spaces, Wightman proved a reconstruction theorem which shows that the
quantum field (as a Hilbert-space valued distribution) can be reconstructed from
its Wightman distributions.

7.3.6 The Correlation Functions
In contrast to the Wightman functions, the correlation functions reflect
causality.
Folklore
Parallel to (7.32), we now define the n-point correlation function (also called the
n-point Green’s function) by setting
Cn(t1, t2, . . . , tn) :=< 0|T (q(t1)q(t2) · · · q(tn))|0>        (7.37)
for all times t1, t2, . . . , tn ∈ R. Here, the symbol T denotes the time-ordering operator,
that is, we define
T (q(t1)q(t2) · · · q(tn)) := q(tπ(1))q(tπ(2)) · · · q(tπ(n))
where the permutation π of the indices 1, 2, . . . , n is chosen in such a way that
tπ(1) ≥ tπ(2) ≥ . . . ≥ tπ(n).

7.4 Schr¨odinger’s Quantum Mechanics
In particular, I would like to mention that I was mainly inspired by the
thoughtful dissertation of Mr. Louis de Broglie (Paris, 1924). The main
difference here lies in the following. De Broglie thinks of travelling waves,
while, in the case of the atom, we are led to standing waves. . . I am most
thankful to Hermann Weyl with regard to the mathematical treatment of
the equation of the hydrogen atom.
Erwin Schr¨odinger, 1926

Freeman Dyson writes in his foreword to Odifreddi’s book:
One of the most profound jokes of nature is the square root of −1 that the
physicist Erwin Schr¨odinger put into his wave equation in 1926 . . . The
Schr¨odinger equation describes correctly everything we know about the behavior
of atoms. It is the basis of all of chemistry and most of physics. And
that square root of −1 means that nature works with complex numbers.
This discovery came as a complete surprise, to Schr¨odinger as well as to
everybody else. According to Schr¨odinger, his fourteen-year-old girlfriend
Itha Junger said to him at the time: “Hey, you never even thought when
you began that so much sensible stuff would come out of it.” All through
the nineteenth century, mathematicians from Abel to Riemann and Weierstrass
had been creating a magnificent theory of functions of complex variables.
They had discovered that the theory of functions became far deeper
and more powerful if it was extended from real to complex numbers. But
they always thought of complex numbers as an artificial construction, invented
by human mathematicians as a useful and elegant abstraction from
real life. It never entered their heads that they had invented was in fact
the ground on which atoms move. They never imagined that nature had
got there first.

7.4.6 Heisenberg’s Uncertainty Principle
In 1927 Heisenberg discovered that there exists a deep difference between classical
mechanics and quantum mechanics. He derived the following fundamental result
in quantum physics:
The classical notion of the trajectory of a particle, which has a precise
position and a precise velocity at the same time, is not meaningful anymore
in quantum mechanics.
Explicitly, for the operators Q, P : S(R) → L2(R) called position operator Q and
momentum operator P, we have the Heisenberg commutation relation
(QP − PQ)ϕ = iϕ for all ϕ ∈ S(R). (7.53)
Let ϕ ∈ S(R) be a normalized state in the Hilbert space L2(R). We claim that
ΔxΔp ≥ h/2. (7.54)
This means that it is impossible to measure precisely the position and the momentum
of the quantum particle in the state ϕ at the same time. The uncertainty
inequality (7.54) follows from (7.53) as a special case of Theorem 10.4 on page 524
of Vol. I.

7.4.7 Unstable Quantum States and the Energy-Time
Uncertainty Relation
In particle accelerators, many particles are unstable; such so-called resonances
only live a very short time.
Folklore
We are going to show that wave packets are unstable in quantum mechanics. There
exists a fundamental inequality between the life-time of the wave packet and its
mean energy fluctuation which is called the energy–time uncertainty relation.

7.5 Feynman’s Quantum Mechanics
It is a curious historical fact that quantum mechanics began with two
quite different mathematical formulations: the differential equation of
Schr¨odinger, and the matrix algebra of Heisenberg. The two, apparently
dissimilar approaches, were proved to be mathematically equivalent. These
two points of view were destined to complement one another and to be ultimately
synthesized in Dirac’s transformation theory.
This paper will describe what is essentially a third formulation of nonrelativistic
quantum theory. This formulation was suggested by some of
Dirac’s remarks concerning the relation of classical action to quantum
mechanics. A probability amplitude is associated with an entire motion of
a particle as a function of time, rather than simply with a position of the
particle at a particular time.
The formulation is mathematically equivalent to the more usual formulations.
There are, therefore, no fundamentally new results. However, there
is a pleasure in recognizing old things from a new point of view. Also, there
are problems for which the new point of view offers a distinct advantage.40
Richard Feynman, 1948
The calculations that I did for Hans Bethe, using the Schr¨odinger equation,
took me several months of work and several hundred sheets of paper.
Dick Feynman (1918–1988) could get the same answer, calculating on a
blackboard, in half an hour.
Freeman Dyson, 1979

The idea is to
pass from time t to imaginary time it and to use analytic continuation in order to
translate well-known results from diffusion processes to quantum processes. This is
called the Euclidean strategy in quantum physics. The following golden rule holds:
Apply analytic continuation only to such quantities that you can measure
in physical experiments.
Analytic continuation of functions depending on energy plays a crucial role in studying
the following subjects:
• scattering processes,
• the energies energies of bound states, and
• the energies of unstable particles having finite lifetime (called resonances).

7.5.1 Main Ideas
The basic idea of Feynman’s approach to quantum mechanics is
• to describe the time-evolution of a quantum system by an integral formula, which
is equivalent to the Schr¨odinger differential equation,
• and to represent the kernel K(x, t; y, t0) of the integral formula by a path integral.
From the physical point of view, Feynman emphasized that
The description of quantum particles becomes easier if we use probability
amplitudes as basic quantities, but not transition probabilities.
The reason is that, in contrast to transition probabilities, probability amplitudes
satisfy a simple composition rule (also called product rule) which is at the heart
of Feynman’s approach to quantum theory. In terms of finite-dimensional Hilbert
spaces, the following hold:
• Feynman’s probability amplitudes are precisely the complex-valued Fourier coefficients
c1, c2, . . . , cn of a state vector.
• Feynman’s composition rule for probability amplitudes coincides with the Parseval
equation (7.81) for Fourier coefficients in mathematics.

According to Feynman, the passage from classical mechanics to quantum
mechanics corresponds to a statistics over all possible classical
paths where the statistical weight exp(iS[q]/h) depends on the classical action.
This is a highly intuitive interpretation of the quantization of classical processes.
Let us discuss the intuitive background.

7.6 Von Neumann’s Rigorous Approach
Rigorous propagator theory is based on von Neumann’s operator calculus
for functions of self-adjoint operators.
Folklore
As a preparation for the rigorous propagator theory to be considered in the next
section, let us summarize von Neumann’s operator calculus. In this section, we consider
an arbitrary complex separable Hilbert space X of finite or infinite dimension.
The inner product on X is denoted by ψ|ϕ for all ϕ, ψ ∈ X.

The Fourier–Laplace
transform is also briefly called the Laplace transform.
Interestingly enough, both retarded (i.e., causal) propagators and advanced
(i.e., non-causal) propagators play a crucial role in quantum field theory.
From the mathematical point of view, the reason is that the relevant perturbation
theory depends on quantities which are constructed by using both retarded and
advanced propagators. Physicists interpret this by saying that
• the interaction between elementary particles is governed by virtual particles
(which are graphically represented by the internal lines of the Feynman diagrams),
and
• the virtual particles violate basic laws of physics (e.g., the relation between energy
and momentum or causality).

7.6.4 The Free Quantum Particle as a Paradigm of Functional
Analysis
Extend the pre-Hamiltonian to a self-adjoint operator on an appropriate
Hilbert space X of quantum states, and use costates related to a Gelfand
triplet with respect to X.
The golden rule
The modern theory of differential and integral equations is based on functional
analysis, which was created by Hilbert (1862–1943) in the beginning of the 20th
century. The development of functional analysis was strongly influenced by the
questions arising in quantum mechanics and quantum field theory. In this section,
we want to study thoroughly how the motion of a free quantum particle on the real
line is related to fundamental notions in functional analysis.
This is Ariadne’s thread in functional analysis.
This way, the formal considerations from Sect. 7.5.3 will obtain a sound basis for
the free quantum particle.
The main idea of the modern strategy in mathematics and physics consists in
describing differential operators and integral operators by abstract operators related
to generalized integral kernels.
(i) In the language used by physicists, this concerns the Dirac calculus based on
Dirac’s delta function and Green’s functions (also called Feynman propagators).
(ii) In the language used by mathematicians this is closely related to:
• Lebesgue’s passage from the Riemann integral to the Lebesgue integral based
on measure theory in about 1900;
• von Neumann’s passage from formally self-adjoint operators to self-adjoint
operators and his generalization of the classical Fourier transform via spectral
theory in the late 1920s;
• Laurent Schwartz’s theory of generalized functions including the kernel theorem
in the 1940s;
• the generalization of von Neumann’s spectral theory by Gelfand and Kostyuchenko
in 1955 (based on quantum costates as generalized functions and
the corresponding Gelfand triplets);
• the extension of the Gelfand–Kostyuchenko approach to general nuclear
spaces by Maurin in 1959.
Tempered Distributions
In order to translate the very elegant, but formal Dirac calculus into mathematics,
one has to leave the Hilbert space of states used by von Neumann
in about 1930. Folklore
In what follows, we will use
• the space S(R) of smooth test functions ϕ : R → C which decrease rapidly at
infinity,
• and the space S
(R) of tempered distributions introduced on page 615 of Vol. I.
Our basic tools will be
• the Fourier transform and
• the language of tempered distributions, and Gelfand triplets.
一星
一星

帖子数 : 3787
注册日期 : 13-08-07

返回页首 向下

4页/共4 上一页  1, 2, 3, 4

返回页首


 
您在这个论坛的权限:
不能在这个论坛回复主题