# i.-introduction-to-qm-ii

I. Reminder of QM I

This course is about advanced quantum mechanics, trying to bridge the gap between the basic knowledge you obtained in the first course on quantum mechanics, and topics developed within the 20th century and whose understanding also help you grasp much of today's research literature. Before we dwell into the topics of the course, let us start with a reminder of what you should have learned in the previous course. I do not assume that you would know all of the following by heart. Rather, each topic should at least ring a bell, that you have heard it before, and after reading the sentences you should at least understand what they say. If not, you can go back to the introductory books and read those topics which seem hard.

Density matrices are an exception to this: they may not have been introduced in earlier courses. This is why this text includes a bit longer introduction to them.

The following consists of a brief reminder of topics in QM I (first six chapters of Griffiths' book).

# a.1-wave-function-and-probability

1.1 Wave function and probability

(See also postulates of quantum mechanics).

Crucial concepts:

  • Time dependent state vector (wave function) , where is a Hilbert space (vector space with a well-defined inner product and where sequences converge to an element inside that space). In QM, state vectors are often denoted with the Dirac notation: . The conjugate of is . The inner product between two state vectors and is a complex number, and it is denoted as .
  • Observables described by Hermitian operators acting on state vectors in : . For example, position , momentum (, ), kinetic energy , total energy operator (Hamiltonian) , angular momentum . The hat symbol is often dropped unless one wants to make a distinction between operators and complex numbers.
  • Expectation value of observables in state : .
  • Representations of the state vector. The state vector can be represented in terms of the spectrum of a given . Typically used representations are position representation , and momentum representation , . (Here and denote one of the coordinates, similar relations exist for each coordinate.)
  • Momentum and position are canonically conjugate variables, which means that the position and momentum representations are related via a Fourier transformation ( is the dimension of the vectors, the prefactor ensures normalization) This implies the gradient representation for :
  • Schrödinger equation. The time dependence of the state vector is governed by the dynamical equation We can write it also in position representation by multiplying from the left by and denoting
  • One consequence of the canonical commutation relations is the uncertainty principle: one cannot measure canonically conjugate variables simultaneously accurately, but the values of the two quantities have an intrinsic uncertainty:
  • This can be generalized to an arbitrary pair of operators and : where corresponds to the variance of the measurements of over a given state and is the commutator between operators and .
# a.2-time-independent-schrödinger-equation

1.2 Time-independent Schrödinger equation

  • If the Hamiltonian is independent of time, we can solve the time-dependent Schrödinger equation with the Ansatz , where is independent of time and satisfies The quantity is the (conserved) energy of the state , it is an eigenvalue of the Hamiltonian , and is the corresponding eigenvector.
  • The time-indepedent Schrödinger equation is often posed as a problem of finding the position representation of . In other words, it amounts to finding a function that satisfies
  • Infinite 1D square well (or a particle in a box) and discretized energies: for , otherwise. Eigenenergies and (normalized) eigenstates of are
  • Harmonic oscillator: potential energy . Define the ladder operators or raising (+) and lowering (-) operators The canonical commutation relations imply . The Hamiltonian can be written in terms of . Energy eigenvalues are , eigenstates satisfy , . The ground state is as .
  • Free particle for which is described by plane wave states. They are also position representations of states with a well-defined momentum, because if , commutes with , and thus is a conserved quantity, see also Subsection on symmetry below. Here is the wave number. Generally plane waves are not normalizable unless we include proper boundary conditions (to be discussed more in the context of time-dependent theory). Sometimes we can also use the concept of a wave packet that contains a (typically Lorentzian) weight function with several values of . Note that a state with a given corresponds to a momentum . The first term describes a wave "going" to the right and the second term a wave "going" to the left.
  • Tunneling through a -function or a square potential. Divide the Ansatz function to parts: where is the Heaviside step function. Tunneling through a square potential In each part, the ansatz consists of left- and right-going states, i.e., where . The coefficients and can be determined by matching wave functions and their derivatives at the boundaries (for the square well), or matching the difference of derivatives to the height of the -function. Finding the coefficients amounts to solving a scattering problem: a particle with amplitude coming from the left reflects with probability amplitude and transmits through with probability amplitude . Note that transmission is allowed even if the potential in the center exceeds the energy of the particle (). In that case transmission is called tunneling.
# a.3-formal-properties-of-state-vectors

1.3 Formal properties of state vectors

  • Identify state vectors with -component vectors where can be (even incountably) infinite.
  • The vector representation is possible via specifying the state vector in terms of the eigenstates of some (Hermitian) operator , i.e., . In this case we can also write (sum becomes an integral in the case of an incountable set) The eigenvectors of Hermitian operators are orthogonal and can be chosen to be normalized, i.e., . They hence form an orthonormal basis. This means that the above representation is unique and that the coefficients can be found from (As for example in position/momentum representation.)
  • The above properties define the Hilbert space in which the state vector lives.
  • Basis functions can be used to generate a projection operator where is assumed to be normalized. projects a given state to state , i.e., . It also satisfies , i.e., double projection is equal to a single projection.
  • Resolution of unity. Representation of a unit operator where the sum goes over states spanning an orthonormal basis for a given Hilbert space (prove it!).
# a.4-quantum-mechanics-in-three-dimensions

1.4 Quantum mechanics in three dimensions

In spaces of dimension higher than one, the position representation of the Schrödinger equation becomes a partial differential equation with several coordinates. This typically calls for proper Ansatz functions that respect the symmetry of the problem (typically the symmetry of the potential or of the boundary conditions).

  • Special case: spherically symmetric potential, i.e., a central potential. Write the Laplacian in spherical coordinates () For the central potential , an appropriate Ansatz is . Plugging it to the Schrödinger equation gives

where is a constant. The term on the rhs of the first equation describes an effective centrifugal potential .

  • Solution for the angular part described by two integer-valued quantum numbers, (azimuthal quantum number) and (magnetic quantum number), : spherical harmonics (no need to remember the exact form) where is a constant, and is the associated Legendre function.
  • Hydrogen atom is an example of a system with a central potential . In the -equation, one gets a new quantum number, (principal quantum number). The eigenenergies are where eV.
  • Angular momentum operator , , , . The components do not commute (identify , , ): where is the completely antisymmetric tensor or Levi-Civita symbol satisfying , , and 0 otherwise. The angular momentum operators have well-defined eigenvalues: , and , where , . Note that are the spherical harmonics.
  • Spin encodes an internal degree of freedom of particles, analog to the particle revolving around itself. Spin operators behave somewhat analogously to angular momentum operators, but they permit half-integer eigenvalues: , , , .
  • Spin 1/2 is an often studied special case. It is encoded by two eigenstates and Spin operators are often represented by Pauli matrices, , (Learn by heart). Pauli matrices satisfy .
  • Presence of spin can be demonstrated in the Stern-Gerlach experiment, where applying an inhomogeneous magnetic field separates particles with different
# a.5-identical-particles

1.5 Identical particles

(We will return to this in QM II B)

  • Bosons (particles with integer spin) and fermions (half integer spins), spin statistics theorem

  • Fermions: Pauli exclusion principle

  • Atoms: shell filling, Hund's rules (not really used in this course)

  • Solids: start from free electron gas, define Fermi energy (highest allowed energy), band structure from Bloch's theorem (not used here, but in the materials physics courses)

  • Quantum statistics: chemical potential, temperature, Fermi-Dirac and Bose-Einstein distributions (discussed more in the statistical physics course)

# a.6-time-independent-perturbation-theory

1.6 Time-independent perturbation theory

(We will discuss time-dependent perturbation theory later in QM II A)

System with a Hamiltonian , where the spectrum of is somehow known, and is a (time independent) perturbation. Idea is to develop a perturbation series, both on eigenfunctions and -energies of : Here .

  • First-order theory (non-degenerate spectrum ): The term is called the matrix element of in the basis defined by .
  • Second-order energies

  • Degenerate spectrum: find the basis spanning the degenerate states, write all matrix elements of within this basis, and diagonalize. Often leads to lifting of degeneracies.

  • Hydrogen atom: lifting of degeneracies due to magnetic field due to spin-orbit coupling (more in QMII B), Zeeman effect (more later in this course) and hyperfine splitting due to coupling of proton and electron spins.

# subs:symmetry

1.7 Addition: symmetry in QM

System described by a Hamiltonian has a given (discrete or continuous) symmetry if the symmetry operator commutes with : . This commutation is equivalent with

  • The value for the observable related to is time independent, i.e., constant of motion.

  • and share the same eigenstates (but naturally not usually the same eigenvalues)

# a.8-postulates-of-quantum-mechanics

1.8 Postulates of quantum mechanics

With the knowledge above, let us state the theory of quantum mechanics in terms of postulates, i.e., things that really cannot be derived, but which we more or less use as starting points.

  1. Information as precise as possible of the physical system at a given time is contained in the state vector . The state vectors are elements of a linear vector space which possesses an inner product.
  1. Corresponding to each observable, i.e., a physical measurable quantity , there is a linear and hermitian operator , which operates on the state vectors of the state vector space. [In this case has an eigenvalue equation , real eigenvalues , and orthogonal (orthonormal) eigenvectors which form a basis.]
  1. In the measurement of an observable , only the eigenvalues of the operator are possible measurement results.
  1. If the system is in a state and an observable is measured, the probability to obtain a value is , i.e., the expectation value of an observable is .
  1. Canonical quantization. The operators and , which correspond to the classical quantities for location and momentum, fulfill the canonical commutation rules
  1. The time evolution of the system state obeys the Schrödinger equation The Hamilton operator is formed from the classical Hamilton function via canonical quantization.

1.9 Density matrix

This may be a new topic to many of you

Above we describe the quantum phenomena using the state vector or wave function description. This is adequate when dealing with closed, isolated systems. Often, however, we have access only to part of the relevant system, or we have a limited knowledge of the state of the system. For example, let us assume that the system is with probability in state (). In this case it is described with a density operator where and are normalized: , . If there are only a finite number of states involved, can be represented with a matrix, and therefore the term density matrix is also often used.

With this definition, the expectation value of observables can be expressed straigthforwardly, Expectation value can hence be calculated from a trace over the operator product of and the density operator : In addition, the density operator has three important properties:

  1. It is Hermitian:

  2. It has a unit trace: .

  3. is a positive operator: for any state , .

Property (i) is easy to prove from the definition of : because is a probability and not a probability amplitude.

Property (ii) comes from the normalization:

Moreover, the positivity condition is proven as This also means that all eigenvalues of are bound to be larger than or equal to zero (they are real because is Hermitian).

# iii.8.1-pure-and-mixed-states

1.9b Pure and mixed states

Let us consider the density matrix corresponding to the superposition state where and are some normalized eigenstates. The density matrix is In the second line we express the states with the two-component vector, This density matrix should be contrasted to that of a mixed state that cannot be represented in terms of a single state vector. For example where both are non-zero.

Pure states, which can be represented in terms of a single state vector, satisfy

A state described by is pure if and only if .

First, let us show that implies a pure state.

Let be the spectral decomposition of . If , we have Because are orthonormal, this means for any . Therefore, or . But because , for some , and . Therefore is a pure state described by the state vector .

Conversely, if ,

We can in fact simplify this requirement to consider only the trace:

is pure iff .

You can prove this as an exercise.

Entropy

John von Neumann showed that besides purity, the pure and mixed states can be distinguished by their entropy, defined as The trace of an operator is independent of the basis in which it is calculated. Therefore, the above can also be written as (check) where is the probability of the density matrix being in state .

It is straightforward to show that the entropy of a pure state vanishes, whereas it is non-vanishing (and ) for a mixed state.

The following sections are strictly speaking outside the course contents, but if you are interested to understand how density matrices are actually used in describing open quantum systems, you may study also them.

1.9c System and a bath, thermal density matrix

1.9d Reduced density operator

1.9e Liouville-von Neumann equation and quantum master equation

These are the current permissions for this document; please modify if needed. You can always modify these permissions from the manage page.