You are on page 1of 10

Hamiltonian (quantum mechanics)

In quantum mechanics, a Hamiltonian is an operator corresponding to the sum of the kinetic energies plus the potential energies for
all the particles in the system (this addition is the total energy of the system in most of the cases under analysis). It is usually denoted
by H, also Ȟ or Ĥ. Its spectrum is the set of possible outcomes when one measures the total energy of a system. Because of its close
relation to the time-evolution of a system, it is of fundamental importance in most formulations of quantum theory
.

The Hamiltonian is named afterWilliam Rowan Hamilton, who created a revolutionary reformulation ofNewtonian mechanicsthat is
now called Hamiltonian mechanicswhich is important in quantum physics.

Contents
Introduction
The Schrödinger Hamiltonian
One particle
Many particles
Schrödinger equation
Dirac formalism
Expressions for the Hamiltonian
General forms for one particle
Free particle
Constant-potential well
Simple harmonic oscillator
Rigid rotor
Electrostatic or coulomb potential
Electric dipole in an electric field
Magnetic dipole in a magnetic field
Charged particle in an electromagnetic field
Energy eigenket degeneracy, symmetry, and conservation laws
Hamilton's equations
See also
References

Introduction
The Hamiltonian is the sum of the kinetic energies of all the particles, plus the potential energy of the particles associated with the
system. The expression of the Hamiltonian can take different forms and simplifications taken into account the concrete characteristics
of the system under analysis: single or several particles in the system; interaction between particles; kind of potential energy; time
varying potential or time independent one; etc.

The Schrödinger Hamiltonian

One particle
By analogy with classical mechanics, the Hamiltonian is commonly expressed as the sum of operators corresponding to the kinetic
and potential energies of a system in the form

where

is the potential energy operator and

is the kinetic energy operator in which m is the mass of the particle, the dot denotes thedot product of vectors, and

is the momentum operator where an ∇ is the del operator. The dot product of ∇ with itself is the Laplacian ∇2. In three dimensions
using Cartesian coordinates the Laplace operator is

Although this is not the technical definition of the Hamiltonian in classical mechanics, it is the form it most commonly takes.
Combining these together yields the familiar form used in theSchrödinger equation:

which allows one to apply the Hamiltonian to systems described by a wave function Ψ(r, t). This is the approach commonly taken in
introductory treatments of quantum mechanics, using the formalism of Schrödinger's wave mechanics.

One can also make substitutions to certain variables to fit specific cases, such as some involving electromagnetic fields.

Many particles
The formalism can be extended toN particles:

where

is the potential energy function, now a function ofthe spatial configuration of the system and time (a particular set of spatial positions
at some instant of time defines a configuration) and;
is the kinetic energy operator of particle n, and ∇n is the gradient for particle n, ∇n2 is the Laplacian for particle using the
coordinates:

Combining these yields the Schrödinger Hamiltonian for theN-particle case:

However, complications can arise in the many-body problem. Since the potential energy depends on the spatial arrangement of the
particles, the kinetic energy will also depend on the spatial configuration to conserve energy. The motion due to any one particle will
vary due to the motion of all the other particles in the system. For this reason cross terms for kinetic energy may appear in the
Hamiltonian; a mix of the gradients for two particles:

where M denotes the mass of the collection of particles resulting in this extra kinetic energy. Terms of this form are known as mass
polarization terms, and appear in the Hamiltonian of many electron atoms (see below).

For N interacting particles, i.e. particles which interact mutually and constitute a many-body situation, the potential energy function V
is not simply a sum of the separate potentials (and certainly not a product, as this is dimensionally incorrect). The potential energy
function can only be written as above: a function of all the spatial positions of each particle.

For non-interacting particles, i.e. particles which do not interact mutually and move independently, the potential of the system is the
sum of the separate potential energy for each particle,[1] that is

The general form of the Hamiltonian in this case is:


where the sum is taken over all particles and their corresponding potentials; the result is that the Hamiltonian of the system is the sum
of the separate Hamiltonians for each particle. This is an idealized situation - in practice the particles are almost always influenced by
some potential, and there are many-body interactions. One illustrative example of a two-body interaction where this form would not
apply is for electrostatic potentials due to charged particles, because they interact with each other by Coulomb interaction
(electrostatic force), as shown below.

Schrödinger equation
The Hamiltonian generates the time evolution of quantum states. If is the state of the system at timet, then

This equation is the Schrödinger equation. It takes the same form as the Hamilton–Jacobi equation, which is one of the reasons H is
also called the Hamiltonian. Given the state at some initial time (t = 0), we can solve it to obtain the state at any subsequent time. In
particular, if H is independent of time, then

The exponential operator on the right hand side of the Schrödinger equation is usually defined by the corresponding power series in
H. One might notice that taking polynomials or power series of unbounded operators that are not defined everywhere may not make
mathematical sense. Rigorously, to take functions of unbounded operators, a functional calculus is required. In the case of the
exponential function, the continuous, or just the holomorphic functional calculus suffices. We note again, however, that for common
calculations the physicists' formulation is quite suf
ficient.

By the *-homomorphism property of the functional calculus, the operator

is a unitary operator. It is the time evolution operator, or propagator, of a closed quantum system. If the Hamiltonian is time-
independent, {U(t)} form a one parameter unitary group (more than a semigroup); this gives rise to the physical principle of detailed
balance.

Dirac formalism
However, in the more general formalism of Dirac, the Hamiltonian is typically implemented as an operator on a Hilbert space in the
following way:

The eigenkets (eigenvectors) of H, denoted , provide an orthonormal basis for the Hilbert space. The spectrum of allowed energy
levels of the system is given by the set of eigenvalues, denoted E{a}, solving the equation:

Since H is a Hermitian operator, the energy is always a real number.

From a mathematically rigorous point of view, care must be taken with the above assumptions. Operators on infinite-dimensional
Hilbert spaces need not have eigenvalues (the set of eigenvalues does not necessarily coincide with the spectrum of an operator).
However, all routine quantum mechanical calculations can be done using the physical formulation.

Expressions for the Hamiltonian


Following are expressions for the Hamiltonian in a number of situations.[2] Typical ways to classify the expressions are the number
of particles, number of dimensions, and the nature of the potential energy function - importantly space and time dependence. Masses
are denoted by m, and charges by q.

General forms for one particle

Free particle
The particle is not bound by any potential energy, so the potential is zero and this Hamiltonian is the simplest. For one dimension:

and in three dimensions:

Constant-potential well
For a particle in a region of constant potentialV = V0 (no dependence on space or time), in one dimension, the Hamiltonian is:

in three dimensions

This applies to the elementary "particle in a box" problem, and step potentials.

Simple harmonic oscillator


For a simple harmonic oscillatorin one dimension, the potential varies with position (but not time), according to:

where the angular frequency , effective spring constant k, and mass m of the oscillator satisfy:

so the Hamiltonian is:

For three dimensions, this becomes


where the three-dimensional position vectorr using cartesian coordinates is (x, y, z), its magnitude is

Writing the Hamiltonian out in full shows it is simply the sum of the one-dimensional Hamiltonian
s in each direction:

Rigid rotor
For a rigid rotor – i.e. system of particles which can rotate freely about any axes, not bound in any potential (such as free molecules
with negligible vibrationaldegrees of freedom, say due to double or triple chemical bonds), Hamiltonian is:

where Ixx, Iyy, and Izz are the moment of inertia components (technically the diagonal elements of the moment of inertia tensor), and
, and are the total angular momentum operators (components), about thex, y, and z axes respectively.

Electrostatic or coulomb potential


The Coulomb potential energy for two point charges q1 and q2 (i.e. charged particles, since particles have no spatial extent), in three
dimensions, is (in SI units - rather than Gaussian units which are frequently used inelectromagnetism):

However, this is only the potential for one point charge due to another. If there are many charged particles, each charge has a potential
energy due to every other point charge (except itself). For N charges, the potential energy of charge qj due to all other charges is (see
also Electrostatic potential energy stored in a configuration of discrete point charges):[3]

where φ(ri) is the electrostatic potential of charge qj at ri. The total potential of the system is then the sum overj:

so the Hamiltonian is:


Electric dipole in an electric field
For an electric dipole moment d constituting charges of magnitude q, in a uniform, electrostatic field (time-independent) E,
positioned in one place, the potential is:

the dipole moment itself is the operator

Since the particle is stationary, there is no translational kinetic energy of the dipole, so the Hamiltonian of the dipole is just the
potential energy:

Magnetic dipole in a magnetic field


For a magnetic dipole momentμ in a uniform, magnetostatic field (time-independent)B, positioned in one place, the potential is:

Since the particle is stationary, there is no translational kinetic energy of the dipole, so the Hamiltonian of the dipole is just the
potential energy:

[4]
For a Spin-½ particle, the corresponding spin magnetic moment is:

where gs is the spin gyromagnetic ratio (a.k.a. "spin g-factor"), e is the electron charge, S is the spin operator vector, whose
components are the Pauli matrices, hence

Charged particle in an electromagnetic field


For a charged particle q in an electromagnetic field, described by the scalar potential φ and vector potential A, there are two parts to
the Hamiltonian to substitute for.[1] The momentum operator must be replaced by the canonical momentum operator, which includes
a contribution from theA field:

where is the kinetic momentum operator given as the usual momentum operator:
so the corresponding kinetic energy operator is:

and the potential energy, which is due to the φ field:

Casting all of these into the Hamiltonian gives:

Energy eigenket degeneracy, symmetry, and conservation laws


In many systems, two or more energy eigenstates have the same energy. A simple example of this is a free particle, whose energy
eigenstates have wavefunctions that are propagating plane waves. The energy of each of these plane waves is inversely proportional
to the square of its wavelength. A wave propagating in the x direction is a different state from one propagating in the y direction, but
if they have the same wavelength, then their ener
gies will be the same. When this happens, the states are said to bedegenerate.

It turns out that degeneracy occurs whenever a nontrivial unitary operator U commutes with the Hamiltonian. To see this, suppose
that is an energy eigenket. Then is an energy eigenket with the same eigenvalue, isnce

Since U is nontrivial, at least one pair of and must represent distinct states. Therefore, H has at least one pair of degenerate
energy eigenkets. In the case of the free particle, the unitary operator which produces the symmetry is the rotation operator, which
rotates the wavefunctions by some angle while otherwise preserving their shape.

The existence of a symmetry operator implies the existence of aconserved observable. Let G be the Hermitian generator ofU:

It is straightforward to show that ifU commutes with H, then so does G:

Therefore,

In obtaining this result, we have used the Schrödinger equation, as well as its
dual,

Thus, the expected value of the observable G is conserved for any state of the system. In the case of the free particle, the conserved
quantity is the angular momentum.

Hamilton's equations
Hamilton's equations in classical Hamiltonian mechanics have a direct analogy in quantum mechanics. Suppose we have a set of
basis states , which need not necessarily be eigenstates of the energy. For simplicity, we assume that they are discrete, and that
they are orthonormal, i.e.,

Note that these basis states are assumed to be independent of time. W


e will assume that the Hamiltonian is also independent of time.

The instantaneous state of the system at timet, , can be expanded in terms of these basis states:

where

The coefficients an(t) are complex variables. We can treat them as coordinates which specify the state of the system, like the position
and momentum coordinates which specify a classical system. Like classical coordinates, they are generally not constant in time, and
their time dependence gives rise to the time dependence of the system as a whole.

The expectation value of the Hamiltonian of this state, which is also the mean ener
gy, is

where the last step was obtained by expanding in terms of the basis states.

Each an(t) actually corresponds to two independent degrees of freedom, since the variable has a real part and an imaginary part. We
now perform the following trick: instead of using the real and imaginary parts as the independent variables, we use an(t) and its
complex conjugate an*(t). With this choice of independent variables, we can calculate thepartial derivative

By applying Schrödinger's equationand using the orthonormality of the basis states, this further reduces to

Similarly, one can show that

If we define "conjugate momentum" variablesπn by

then the above equations become


which is precisely the form of Hamilton's equations, with the s as the generalized coordinates, the s as the conjugate momenta,
and taking the place of the classical Hamiltonian.

See also
Hamiltonian mechanics
Operator (physics)
Bra–ket notation
Quantum state
Linear algebra
Conservation of energy
Potential theory
Many-body problem
Electrostatics
Electric field
Magnetic field
Lieb–Thirring inequality

References
1. Quantum Physics of Atoms, Molecules, Solids, Nuclei and Particles (2nd Edition), R. Resnick, R. Eisberg, John
Wiley & Sons, 1985, ISBN 978-0-471-87373-0
2. Quanta: A handbook of concepts, P
.W. Atkins, Oxford University Press, 1974,ISBN 0-19-855493-1
3. Electromagnetism (2nd edition), I.S. Grant, W
.R. Phillips, Manchester Physics Series, 2008ISBN 0-471-92712-0
4. Physics of Atoms and Molecules, B.H. Bransden, C.J.Joachain, Longman, 1983,
ISBN 0-582-44401-2

Retrieved from "https://en.wikipedia.org/w/index.php?title=Hamiltonian_(quantum_mechanics)&oldid=893563651


"

This page was last edited on 22 April 2019, at 07:37(UTC).

Text is available under theCreative Commons Attribution-ShareAlike License ; additional terms may apply. By using this
site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of theWikimedia
Foundation, Inc., a non-profit organization.

You might also like