Quantum Physics
Quantum Physics
Quantum Physics
properties of nature at the scale of atoms and subatomic particles.[2]: 1.1 It is the foundation of all
quantum physics including quantum chemistry, quantum field theory, quantum technology, and
quantum information science.
Classical physics, the collection of theories that existed before the advent of quantum mechanics,
describes many aspects of nature at an ordinary (macroscopic) scale, but is not sufficient for describing
them at small (atomic and subatomic) scales. Most theories in classical physics can be derived from
quantum mechanics as an approximation valid at large (macroscopic) scale.[3]
Quantum mechanics differs from classical physics in that energy, momentum, angular momentum, and
other quantities of a bound system are restricted to discrete values (quantization); objects have
characteristics of both particles and waves (wave–particle duality); and there are limits to how
accurately the value of a physical quantity can be predicted prior to its measurement, given a complete
set of initial conditions (the uncertainty principle).
Quantum mechanics arose gradually from theories to explain observations that could not be reconciled
with classical physics, such as Max Planck's solution in 1900 to the black-body radiation problem, and
the correspondence between energy and frequency in Albert Einstein's 1905 paper, which explained the
photoelectric effect. These early attempts to understand microscopic phenomena, now known as the
"old quantum theory", led to the full development of quantum mechanics in the mid-1920s by Niels
Bohr, Erwin Schrödinger, Werner Heisenberg, Max Born, Paul Dirac and others. The modern theory is
formulated in various specially developed mathematical formalisms. In one of them, a mathematical
entity called the wave function provides information, in the form of probability amplitudes, about what
measurements of a particle's energy, momentum, and other physical properties may yield.
Quantum mechanics allows the calculation of properties and behaviour of physical systems. It is typically
applied to microscopic systems: molecules, atoms and sub-atomic particles. It has been demonstrated to
hold for complex molecules with thousands of atoms,[4] but its application to human beings raises
philosophical problems, such as Wigner's friend, and its application to the universe as a whole remains
speculative.[5] Predictions of quantum mechanics have been verified experimentally to an extremely
high degree of accuracy.[note 1]
A fundamental feature of the theory is that it usually cannot predict with certainty what will happen, but
only give probabilities. Mathematically, a probability is found by taking the square of the absolute value
of a complex number, known as a probability amplitude. This is known as the Born rule, named after
physicist Max Born. For example, a quantum particle like an electron can be described by a wave
function, which associates to each point in space a probability amplitude. Applying the Born rule to
these amplitudes gives a probability density function for the position that the electron will be found to
have when an experiment is performed to measure it. This is the best the theory can do; it cannot say
for certain where the electron will be found. The Schrödinger equation relates the collection of
probability amplitudes that pertain to one moment of time to the collection of probability amplitudes
that pertain to another.
Another consequence of the mathematical rules of quantum mechanics is the phenomenon of quantum
interference, which is often illustrated with the double-slit experiment. In the basic version of this
experiment, a coherent light source, such as a laser beam, illuminates a plate pierced by two parallel
slits, and the light passing through the slits is observed on a screen behind the plate.[6]: 102–111 [2]:
1.1–1.8 The wave nature of light causes the light waves passing through the two slits to interfere,
producing bright and dark bands on the screen – a result that would not be expected if light consisted of
classical particles.[6] However, the light is always found to be absorbed at the screen at discrete points,
as individual particles rather than waves; the interference pattern appears via the varying density of
these particle hits on the screen. Furthermore, versions of the experiment that include detectors at the
slits find that each detected photon passes through one slit (as would a classical particle), and not
through both slits (as would a wave).[6]: 109 [7][8] However, such experiments demonstrate that
particles do not form the interference pattern if one detects which slit they pass through. Other atomic-
scale entities, such as electrons, are found to exhibit the same behavior when fired towards a double
slit.[2] This behavior is known as wave–particle duality.
When quantum systems interact, the result can be the creation of quantum entanglement: their
properties become so intertwined that a description of the whole solely in terms of the individual parts
is no longer possible. Erwin Schrödinger called entanglement "...the characteristic trait of quantum
mechanics, the one that enforces its entire departure from classical lines of thought".[11] Quantum
entanglement enables the counter-intuitive properties of quantum pseudo-telepathy, and can be a
valuable resource in communication protocols, such as quantum key distribution and superdense
coding.[12] Contrary to popular misconception, entanglement does not allow sending signals faster than
light, as demonstrated by the no-communication theorem.[12]
Another possibility opened by entanglement is testing for "hidden variables", hypothetical properties
more fundamental than the quantities addressed in quantum theory itself, knowledge of which would
allow more exact predictions than quantum theory can provide. A collection of results, most significantly
Bell's theorem, have demonstrated that broad classes of such hidden-variable theories are in fact
incompatible with quantum physics. According to Bell's theorem, if nature actually operates in accord
with any theory of local hidden variables, then the results of a Bell test will be constrained in a
particular, quantifiable way. Many Bell tests have been performed, using entangled particles, and they
have shown results incompatible with the constraints imposed by local hidden variables.[13][14]
It is not possible to present these concepts in more than a superficial way without introducing the actual
mathematics involved; understanding quantum mechanics requires not only manipulating complex
numbers, but also linear algebra, differential equations, group theory, and other more advanced
subjects.[note 2] Accordingly, this article will present a mathematical formulation of quantum
mechanics and survey its application to some useful and oft-studied examples.
Mathematical formulation
In the mathematically rigorous formulation of quantum mechanics, the state of a quantum mechanical
system is a vector
{\mathcal {H}}. This vector is postulated to be normalized under the Hilbert space inner product, that is,
it obeys
�
⟩
1
{\displaystyle \langle \psi ,\psi \rangle =1}, and it is well-defined up to a complex number of modulus 1
(the global phase), that is,
�\psi and
�{\displaystyle e^{i\alpha }\psi } represent the same physical system. In other words, the possible
states are points in the projective space of a Hilbert space, usually called the complex projective space.
The exact nature of this Hilbert space is dependent on the system – for example, for describing position
and momentum the Hilbert space is the space of complex square-integrable functions
{\displaystyle L^{2}(\mathbb {C} )}, while the Hilbert space for the spin of a single proton is simply the
space of two-dimensional complex vectors
Physical quantities of interest – position, momentum, energy, spin – are represented by observables,
which are Hermitian (more precisely, self-adjoint) linear operators acting on the Hilbert space. A
quantum state can be an eigenvector of an observable, in which case it is called an eigenstate, and the
associated eigenvalue corresponds to the value of the observable in that eigenstate. More generally, a
quantum state will be a linear combination of the eigenstates, known as a quantum superposition.
When an observable is measured, the result will be one of its eigenvalues with probability given by the
Born rule: in the simplest case the eigenvalue
→
,
�
⟩
→{\displaystyle {\vec {\lambda }}} is its associated eigenvector. More generally, the eigenvalue is
degenerate and the probability is given by
�P_{\lambda } is the projector onto its associated eigenspace. In the continuous case, these formulas
give instead the probability density.
�
,
⟩{\displaystyle P_{\lambda }\psi /{\sqrt {\langle \psi ,P_{\lambda }\psi \rangle }}}, in the general case.
The probabilistic nature of quantum mechanics thus stems from the act of measurement. This is one of
the most difficult aspects of quantum systems to understand. It was the central topic in the famous
Bohr–Einstein debates, in which the two scientists attempted to clarify these fundamental principles by
way of thought experiments. In the decades after the formulation of quantum mechanics, the question
of what constitutes a "measurement" has been extensively studied. Newer interpretations of quantum
mechanics have been formulated that do away with the concept of "wave function collapse" (see, for
example, the many-worlds interpretation). The basic idea is that when a quantum system interacts with
a measuring apparatus, their respective wave functions become entangled so that the original quantum
system ceases to exist as an independent entity. For details, see the article on measurement in quantum
mechanics.[17]
�
ℏ
�
)
Here
H denotes the Hamiltonian, the observable corresponding to the total energy of the system, and
ℏi\hbar is introduced so that the Hamiltonian is reduced to the classical Hamiltonian in cases where the
quantum system can be approximated by a classical system; the ability to make such an approximation
in certain limits is called the correspondence principle.
)
.
The operator
ℏ{\displaystyle U(t)=e^{-iHt/\hbar }} is known as the time-evolution operator, and has the crucial
property that it is unitary. This time evolution is deterministic in the sense that – given an initial
quantum state
Some wave functions produce probability distributions that are independent of time, such as
eigenstates of the Hamiltonian. Many systems that are treated dynamically in classical mechanics are
described by such "static" wave functions. For example, a single electron in an unexcited atom is
pictured classically as a particle moving in a circular trajectory around the atomic nucleus, whereas in
quantum mechanics, it is described by a static wave function surrounding the nucleus. For example, the
electron wave function for an unexcited hydrogen atom is a spherically symmetric function known as an
s orbital (Fig. 1).
Analytic solutions of the Schrödinger equation are known for very few relatively simple model
Hamiltonians including the quantum harmonic oscillator, the particle in a box, the dihydrogen cation,
and the hydrogen atom. Even the helium atom – which contains just two electrons – has defied all
attempts at a fully analytic treatment.
However, there are techniques for finding approximate solutions. One method, called perturbation
theory, uses the analytic result for a simple quantum mechanical model to create a result for a related
but more complicated model by (for example) the addition of a weak potential energy. Another method
is called "semi-classical equation of motion", which applies to systems for which quantum mechanics
produces only small deviations from classical behavior. These deviations can then be computed based
on the classical motion. This approach is particularly important in the field of quantum chaos.
Uncertainty principle
One consequence of the basic quantum formalism is the uncertainty principle. In its most familiar form,
this states that no preparation of a quantum particle can imply simultaneously precise predictions both
for a measurement of its position and for a measurement of its momentum.[19][20] Both position and
momentum are observables, meaning that they are represented by Hermitian operators. The position
operator
^\hat{P} do not commute, but rather satisfy the canonical commutation relation:
[
�
ℏ
Given a quantum state, the Born rule lets us compute expectation values for both
X and
P, and moreover for powers of them. Defining the uncertainty for an observable by a standard deviation,
we have
⟨
�
⟩
�
⟩
ℏ
2
Either standard deviation can in principle be made arbitrarily small, but not both simultaneously.[21]
This inequality generalizes to arbitrary pairs of self-adjoint operators
A and
{\displaystyle [A,B]=AB-BA,}
and this provides the lower bound on the product of standard deviations:
�
≥
Another consequence of the canonical commutation relation is that the position and momentum
operators are Fourier transforms of each other, so that a description of an object according to its
momentum is the Fourier transform of its description according to its position. The fact that
dependence in momentum is the Fourier transform of the dependence in position means that the
momentum operator is equivalent (up to an
ℏ{\displaystyle i/\hbar } factor) to taking the derivative according to the position, since in Fourier
analysis differentiation corresponds to multiplication in the dual space. This is why in quantum
equations in position space, the momentum
p_{i} is replaced by
�
ℏ
∂
∂
{\displaystyle -i\hbar {\frac {\partial }{\partial x}}}, and in particular in the non-relativistic Schrödinger
equation in position space the momentum-squared term is replaced with a Laplacian times
-\hbar ^{2}.[19]
When two different quantum systems are considered together, the Hilbert space of the combined
system is the tensor product of the Hilbert spaces of the two components. For example, let A and B be
two quantum systems, with Hilbert spaces
{\displaystyle {\mathcal {H}}_{B}}, respectively. The Hilbert space of the composite system is then
�
⊗
.
{\displaystyle {\mathcal {H}}_{AB}={\mathcal {H}}_{A}\otimes {\mathcal {H}}_{B}.}
{\displaystyle \psi _{A}} and the state for the second system is
�
⊗
{\displaystyle {\mathcal {H}}_{AB}} can be written in this form, however, because the superposition
principle implies that linear combinations of these "separable" or "product states" are also valid. For
example, if
�
A, and likewise
B, then
�
⊗
�
⊗
{\displaystyle {\tfrac {1}{\sqrt {2}}}\left(\psi _{A}\otimes \psi _{B}+\phi _{A}\otimes \phi _{B}\right)}
is a valid joint state that is not separable. States that are not separable are called entangled.[22][23]
If the state for a composite system is entangled, it is impossible to describe either component system A
or system B by a state vector. One can instead define reduced density matrices that describe the
statistics that can be obtained by making measurements on either component system alone. This
necessarily causes a loss of information, though: knowing the reduced density matrices of the individual
systems is not enough to reconstruct the state of the composite system.[22][23] Just as density matrices
specify the state of a subsystem of a larger system, analogously, positive operator-valued measures
(POVMs) describe the effect on a subsystem of a measurement performed on a larger system. POVMs
are extensively used in quantum information theory.[22][24]
There are many mathematically equivalent formulations of quantum mechanics. One of the oldest and
most common is the "transformation theory" proposed by Paul Dirac, which unifies and generalizes the
two earliest formulations of quantum mechanics – matrix mechanics (invented by Werner Heisenberg)
and wave mechanics (invented by Erwin Schrödinger).[26] An alternative formulation of quantum
mechanics is Feynman's path integral formulation, in which a quantum-mechanical amplitude is
considered as a sum over all possible classical and non-classical paths between the initial and final
states. This is the quantum-mechanical counterpart of the action principle in classical mechanics.
The Hamiltonian
H is known as the generator of time evolution, since it defines a unitary time-evolution operator
�
−
U(t) and
H will be conserved: its expectation value will not change over time. This statement generalizes, as
mathematically, any Hermitian operator
A, any observable
�
A will be conserved. Moreover, if
A, then
B. This implies a quantum version of the result proven by Emmy Noether in classical (Lagrangian)
mechanics: for every differentiable symmetry of a Hamiltonian, there exists a corresponding
conservation law.
Examples
Free particle
Position space probability density of a Gaussian wave packet moving in one dimension in free space
The simplest example of a quantum system with a position degree of freedom is a free particle in a
single spatial dimension. A free particle is one which is not subject to external influences, so that its
Hamiltonian consists only of its kinetic energy:
−
ℏ