The term complex manifold is variously used to mean a complex manifold in the sense above (which can be specified as an integrable complex manifold), and an almost complex manifold.
Implications of complex structure
Since holomorphic functions are much more rigid than smooth functions, the theories of smooth and complex manifolds have very different flavors: compact complex manifolds are much closer to algebraic varieties than to differentiable manifolds.
For example, the Whitney embedding theorem tells us that every smooth n-dimensional manifold can be embedded as a smooth submanifold of R2n, whereas it is "rare" for a complex manifold to have a holomorphic embedding into Cn. Consider for example any compact connected complex manifold M: any holomorphic function on it is constant by the maximum modulus principle. Now if we had a holomorphic embedding of M into Cn, then the coordinate functions of Cn would restrict to nonconstant holomorphic functions on M, contradicting compactness, except in the case that M is just a point. Complex manifolds that can be embedded in Cn are called Stein manifolds and form a very special class of manifolds including, for example, smooth complex affine algebraic varieties.
The classification of complex manifolds is much more subtle than
that of differentiable manifolds. For example, while in dimensions other
than four, a given topological manifold has at most finitely many smooth structures, a topological manifold supporting a complex structure can and often does support uncountably many complex structures. Riemann surfaces, two dimensional manifolds equipped with a complex structure, which are topologically classified by the genus,
are an important example of this phenomenon. The set of complex
structures on a given orientable surface, modulo biholomorphic
equivalence, itself forms a complex algebraic variety called a moduli space, the structure of which remains an area of active research.
Since the transition maps between charts are biholomorphic,
complex manifolds are, in particular, smooth and canonically oriented
(not just orientable: a biholomorphic map to (a subset of) Cn gives an orientation, as biholomorphic maps are orientation-preserving).
Note that there are inclusions between these as
Δ ⊆ C ⊆ Ĉ, but that there are no non-constant holomorphic maps in the other direction, by
Liouville's theorem.
Disc vs. space vs. polydisc
The
following spaces are different as complex manifolds, demonstrating the
more rigid geometric character of complex manifolds (compared to smooth
manifolds):
Concretely, this is an endomorphism of the tangent bundle whose square is −I; this endomorphism is analogous to multiplication by the imaginary number i, and is denoted J (to avoid confusion with the identity matrix I). An almost complex manifold is necessarily even-dimensional.
An almost complex structure is weaker than a complex
structure: any complex manifold has an almost complex structure, but not
every almost complex structure comes from a complex structure. Note
that every even-dimensional real manifold has an almost complex
structure defined locally from the local coordinate chart. The question
is whether this almost complex structure can be defined globally. An
almost complex structure that comes from a complex structure is called integrable, and when one wishes to specify a complex structure as opposed to an almost complex structure, one says an integrable complex structure. For integrable complex structures the so-called Nijenhuis tensor vanishes. This tensor is defined on pairs of vector fields, X, Y by
For example, the 6-dimensional sphereS6 has a natural almost complex structure arising from the fact that it is the orthogonal complement of i in the unit sphere of the octonions, but this is not a complex structure. (The question of whether it has a complex structure is known as the Hopf problem, after Heinz Hopf.[3])
Using an almost complex structure we can make sense of holomorphic maps
and ask about the existence of holomorphic coordinates on the manifold.
The existence of holomorphic coordinates is equivalent to saying the
manifold is complex (which is what the chart definition says).
Tensoring the tangent bundle with the complex numbers we get the complexified
tangent bundle, on which multiplication by complex numbers makes sense
(even if we started with a real manifold). The eigenvalues of an almost
complex structure are ±i and the eigenspaces form sub-bundles denoted by T0,1M and T1,0M. The Newlander–Nirenberg theorem shows that an almost complex structure is actually a complex structure precisely when these subbundles are involutive, i.e., closed under the Lie bracket of vector fields, and such an almost complex structure is called integrable.
Kähler and Calabi–Yau manifolds
One can define an analogue of a Riemannian metric for complex manifolds, called a Hermitian metric.
Like a Riemannian metric, a Hermitian metric consists of a smoothly
varying, positive definite inner product on the tangent bundle, which is
Hermitian with respect to the complex structure on the tangent space at
each point. As in the Riemannian case, such metrics always exist in
abundance on any complex manifold. If the skew symmetric part of such a
metric is symplectic, i.e. closed and nondegenerate, then the metric is called Kähler. Kähler structures are much more difficult to come by and are much more rigid.
Examples of Kähler manifolds include smooth projective varieties and more generally any complex submanifold of a Kähler manifold. The Hopf manifolds
are examples of complex manifolds that are not Kähler. To construct
one, take a complex vector space minus the origin and consider the
action of the group of integers on this space by multiplication by exp(n). The quotient is a complex manifold whose first Betti number is one, so by the Hodge theory, it cannot be Kähler.
Early cases of mirror symmetry were discovered by physicists.
Mathematicians became interested in this relationship around 1990 when Philip Candelas, Xenia de la Ossa, Paul Green, and Linda Parkes showed that it could be used as a tool in enumerative geometry,
a branch of mathematics concerned with counting the number of solutions
to geometric questions. Candelas and his collaborators showed that
mirror symmetry could be used to count rational curves
on a Calabi–Yau manifold, thus solving a longstanding problem. Although
the original approach to mirror symmetry was based on physical ideas
that were not understood in a mathematically precise way, some of its
mathematical predictions have since been proven rigorously.
In physics, string theory is a theoretical framework in which the point-like particles of particle physics are replaced by one-dimensional objects called strings.
These strings look like small segments or loops of ordinary string.
String theory describes how strings propagate through space and interact
with each other. On distance scales larger than the string scale, a
string will look just like an ordinary particle, with its mass, charge,
and other properties determined by the vibrational state of the string.
Splitting and recombination of strings correspond to particle emission
and absorption, giving rise to the interactions between particles.
There are notable differences between the world described by
string theory and the everyday world. In everyday life, there are three
familiar dimensions of space (up/down, left/right, and
forward/backward), and there is one dimension of time (later/earlier).
Thus, in the language of modern physics, one says that spacetime is four-dimensional. One of the peculiar features of string theory is that it requires extra dimensions of spacetime for its mathematical consistency. In superstring theory, the version of the theory that incorporates a theoretical idea called supersymmetry, there are six extra dimensions of spacetime in addition to the four that are familiar from everyday experience.
One of the goals of current research in string theory is to
develop models in which the strings represent particles observed in high
energy physics experiments. For such a model to be consistent with
observations, its spacetime must be four-dimensional at the relevant
distance scales, so one must look for ways to restrict the extra
dimensions to smaller scales. In most realistic models of physics based
on string theory, this is accomplished by a process called compactification, in which the extra dimensions are assumed to "close up" on themselves to form circles.
In the limit where these curled up dimensions become very small, one
obtains a theory in which spacetime has effectively a lower number of
dimensions. A standard analogy for this is to consider a
multidimensional object such as a garden hose. If the hose is viewed
from a sufficient distance, it appears to have only one dimension, its
length. However, as one approaches the hose, one discovers that it
contains a second dimension, its circumference. Thus, an ant crawling on
the surface of the hose would move in two dimensions.
Compactification can be used to construct models in which spacetime
is effectively four-dimensional. However, not every way of compactifying
the extra dimensions produces a model with the right properties to
describe nature. In a viable model of particle physics, the compact
extra dimensions must be shaped like a Calabi–Yau manifold. A Calabi–Yau manifold is a special space which is typically taken to be six-dimensional in applications to string theory. It is named after mathematicians Eugenio Calabi and Shing-Tung Yau.
After Calabi–Yau manifolds had entered physics as a way to
compactify extra dimensions, many physicists began studying these
manifolds. In the late 1980s, Lance Dixon, Wolfgang Lerche, Cumrun Vafa,
and Nick Warner noticed that given such a compactification of string
theory, it is not possible to reconstruct uniquely a corresponding
Calabi–Yau manifold. Instead, two different versions of string theory called type IIA string theory and type IIB can be compactified on completely different Calabi–Yau manifolds giving rise to the same physics.
In this situation, the manifolds are called mirror manifolds, and the
relationship between the two physical theories is called mirror
symmetry.
The mirror symmetry relationship is a particular example of what physicists call a physical duality. In general, the term physical duality
refers to a situation where two seemingly different physical theories
turn out to be equivalent in a nontrivial way. If one theory can be
transformed so it looks just like another theory, the two are said to be
dual under that transformation. Put differently, the two theories are
mathematically different descriptions of the same phenomena. Such dualities play an important role in modern physics, especially in string theory.
Regardless of whether Calabi–Yau compactifications of string
theory provide a correct description of nature, the existence of the
mirror duality between different string theories has significant
mathematical consequences. The Calabi–Yau manifolds used in string theory are of interest in pure mathematics, and mirror symmetry allows mathematicians to solve problems in enumerative algebraic geometry,
a branch of mathematics concerned with counting the numbers of
solutions to geometric questions. A classical problem of enumerative
geometry is to enumerate the rational curves
on a Calabi–Yau manifold such as the one illustrated above. By applying
mirror symmetry, mathematicians have translated this problem into an
equivalent problem for the mirror Calabi–Yau, which turns out to be
easier to solve.
In physics, mirror symmetry is justified on physical grounds. However, mathematicians generally require rigorous proofs
that do not require an appeal to physical intuition. From a
mathematical point of view, the version of mirror symmetry described
above is still only a conjecture, but there is another version of mirror
symmetry in the context of topological string theory, a simplified version of string theory introduced by Edward Witten, which has been rigorously proven by mathematicians. In the context of topological string theory, mirror symmetry states that two theories called the A-model and B-model are equivalent in the sense that there is a duality relating them.
Today mirror symmetry is an active area of research in mathematics, and
mathematicians are working to develop a more complete mathematical
understanding of mirror symmetry based on physicists' intuition.
History
The
idea of mirror symmetry can be traced back to the mid-1980s when it was
noticed that a string propagating on a circle of radius is physically equivalent to a string propagating on a circle of radius in appropriate units. This phenomenon is now known as T-duality and is understood to be closely related to mirror symmetry. In a paper from 1985, Philip Candelas, Gary Horowitz, Andrew Strominger,
and Edward Witten showed that by compactifying string theory on a
Calabi–Yau manifold, one obtains a theory roughly similar to the standard model of particle physics that also consistently incorporates an idea called supersymmetry.
Following this development, many physicists began studying Calabi–Yau
compactifications, hoping to construct realistic models of particle
physics based on string theory. Cumrun Vafa and others noticed that
given such a physical model, it is not possible to reconstruct uniquely a
corresponding Calabi–Yau manifold. Instead, there are two Calabi–Yau
manifolds that give rise to the same physics.
By studying the relationship between Calabi–Yau manifolds and certain conformal field theories called Gepner models, Brian Greene and Ronen Plesser found nontrivial examples of the mirror relationship.
Further evidence for this relationship came from the work of Philip
Candelas, Monika Lynker, and Rolf Schimmrigk, who surveyed a large
number of Calabi–Yau manifolds by computer and found that they came in
mirror pairs.
Mathematicians became interested in mirror symmetry around 1990
when physicists Philip Candelas, Xenia de la Ossa, Paul Green, and Linda
Parkes showed that mirror symmetry could be used to solve problems in
enumerative geometry that had resisted solution for decades or more. These results were presented to mathematicians at a conference at the Mathematical Sciences Research Institute (MSRI) in Berkeley, California
in May 1991. During this conference, it was noticed that one of the
numbers Candelas had computed for the counting of rational curves
disagreed with the number obtained by Norwegian mathematicians Geir Ellingsrud and Stein Arild Strømme using ostensibly more rigorous techniques.
Many mathematicians at the conference assumed that Candelas's work
contained a mistake since it was not based on rigorous mathematical
arguments. However, after examining their solution, Ellingsrud and
Strømme discovered an error in their computer code and, upon fixing the
code, they got an answer that agreed with the one obtained by Candelas
and his collaborators.
In 1990, Edward Witten introduced topological string theory,
a simplified version of string theory, and physicists showed that there
is a version of mirror symmetry for topological string theory.
This statement about topological string theory is usually taken as the
definition of mirror symmetry in the mathematical literature. In an address at the International Congress of Mathematicians in 1994, mathematician Maxim Kontsevich presented a new mathematical conjecture based on the physical idea of mirror symmetry in topological string theory. Known as homological mirror symmetry, this conjecture formalizes mirror symmetry as an equivalence of two mathematical structures: the derived category of coherent sheaves on a Calabi–Yau manifold and the Fukaya category of its mirror.
Also around 1995, Kontsevich analyzed the results of Candelas,
which gave a general formula for the problem of counting rational curves
on a quintic threefold, and he reformulated these results as a precise mathematical conjecture. In 1996, Alexander Givental posted a paper that claimed to prove this conjecture of Kontsevich.
Initially, many mathematicians found this paper hard to understand, so
there were doubts about its correctness. Subsequently, Bong Lian, Kefeng Liu, and Shing-Tung Yau published an independent proof in a series of papers.
Despite controversy over who had published the first proof, these
papers are now collectively seen as providing a mathematical proof of
the results originally obtained by physicists using mirror symmetry. In 2000, Kentaro Hori and Cumrun Vafa gave another physical proof of mirror symmetry based on T-duality.
Work on mirror symmetry continues today with major developments in the context of strings on surfaces with boundaries. In addition, mirror symmetry has been related to many active areas of mathematics research, such as the McKay correspondence, topological quantum field theory, and the theory of stability conditions.
At the same time, basic questions continue to vex. For example,
mathematicians still lack an understanding of how to construct examples
of mirror Calabi–Yau pairs though there has been progress in
understanding this issue.
Many of the important mathematical applications of mirror symmetry
belong to the branch of mathematics called enumerative geometry. In
enumerative geometry, one is interested in counting the number of
solutions to geometric questions, typically using the techniques of algebraic geometry. One of the earliest problems of enumerative geometry was posed around the year 200 BCE by the ancient Greek mathematician Apollonius, who asked how many circles in the plane are tangent to three given circles. In general, the solution to the problem of Apollonius is that there are eight such circles.
Enumerative problems in mathematics often concern a class of geometric objects called algebraic varieties which are defined by the vanishing of polynomials. For example, the Clebsch cubic (see the illustration) is defined using a certain polynomial of degree three in four variables. A celebrated result of nineteenth-century mathematicians Arthur Cayley and George Salmon states that there are exactly 27 straight lines that lie entirely on such a surface.
Generalizing this problem, one can ask how many lines can be
drawn on a quintic Calabi–Yau manifold, such as the one illustrated
above, which is defined by a polynomial of degree five. This problem was
solved by the nineteenth-century German mathematician Hermann Schubert,
who found that there are exactly 2,875 such lines. In 1986, geometer
Sheldon Katz proved that the number of curves, such as circles, that are
defined by polynomials of degree two and lie entirely in the quintic is
609,250.
By the year 1991, most of the classical problems of enumerative
geometry had been solved and interest in enumerative geometry had begun
to diminish. According to mathematician Mark Gross,
"As the old problems had been solved, people went back to check
Schubert's numbers with modern techniques, but that was getting pretty
stale."
The field was reinvigorated in May 1991 when physicists Philip
Candelas, Xenia de la Ossa, Paul Green, and Linda Parkes showed that
mirror symmetry could be used to count the number of degree three curves
on a quintic Calabi–Yau. Candelas and his collaborators found that
these six-dimensional Calabi–Yau manifolds can contain exactly
317,206,375 curves of degree three.
In addition to counting degree-three curves on a quintic
three-fold, Candelas and his collaborators obtained a number of more
general results for counting rational curves which went far beyond the
results obtained by mathematicians. Although the methods used in this work were based on physical intuition, mathematicians have gone on to prove rigorously
some of the predictions of mirror symmetry. In particular, the
enumerative predictions of mirror symmetry have now been rigorously
proven.[]
Theoretical physics
In
addition to its applications in enumerative geometry, mirror symmetry
is a fundamental tool for doing calculations in string theory. In the
A-model of topological string theory, physically interesting quantities
are expressed in terms of infinitely many numbers called Gromov–Witten invariants, which are extremely difficult to compute. In the B-model, the calculations can be reduced to classical integrals and are much easier.
By applying mirror symmetry, theorists can translate difficult
calculations in the A-model into equivalent but technically easier
calculations in the B-model. These calculations are then used to
determine the probabilities of various physical processes in string
theory. Mirror symmetry can be combined with other dualities to
translate calculations in one theory into equivalent calculations in a
different theory. By outsourcing calculations to different theories in
this way, theorists can calculate quantities that are impossible to
calculate without the use of dualities.
Outside of string theory, mirror symmetry is used to understand aspects of quantum field theory, the formalism that physicists use to describe elementary particles. For example, gauge theories
are a class of highly symmetric physical theories appearing in the
standard model of particle physics and other parts of theoretical
physics. Some gauge theories which are not part of the standard model,
but which are nevertheless important for theoretical reasons, arise from
strings propagating on a nearly singular background. For such theories,
mirror symmetry is a useful computational tool.
Indeed, mirror symmetry can be used to perform calculations in an
important gauge theory in four spacetime dimensions that was studied by Nathan Seiberg and Edward Witten and is also familiar in mathematics in the context of Donaldson invariants. There is also a generalization of mirror symmetry called 3D mirror symmetry which relates pairs of quantum field theories in three spacetime dimensions.
In string theory and related theories in physics, a brane
is a physical object that generalizes the notion of a point particle to
higher dimensions. For example, a point particle can be viewed as a
brane of dimension zero, while a string can be viewed as a brane of
dimension one. It is also possible to consider higher-dimensional
branes. The word brane comes from the word "membrane" which refers to a
two-dimensional brane.
In string theory, a string may be open (forming a segment with two endpoints) or closed (forming a closed loop). D-branes
are an important class of branes that arise when one considers open
strings. As an open string propagates through spacetime, its endpoints
are required to lie on a D-brane. The letter "D" in D-brane refers to a
condition that it satisfies, the Dirichlet boundary condition.
Mathematically, branes can be described using the notion of a category. This is a mathematical structure consisting of objects, and for any pair of objects, a set of morphisms between them. In most examples, the objects are mathematical structures (such as sets, vector spaces, or topological spaces) and the morphisms are functions between these structures. One can also consider categories where the objects are D-branes and the morphisms between two branes and are states of open strings stretched between and .
In the B-model of topological string theory, the D-branes are complex submanifolds of a Calabi–Yau together with additional data that arise physically from having charges at the endpoints of strings.
Intuitively, one can think of a submanifold as a surface embedded
inside the Calabi–Yau, although submanifolds can also exist in
dimensions different from two.
In mathematical language, the category having these branes as its
objects is known as the derived category of coherent sheaves on the
Calabi–Yau.
In the A-model, the D-branes can again be viewed as submanifolds of a
Calabi–Yau manifold. Roughly speaking, they are what mathematicians call
special Lagrangian submanifolds.
This means among other things that they have half the dimension of the
space in which they sit, and they are length-, area-, or
volume-minimizing. The category having these branes as its objects is called the Fukaya category.
The derived category of coherent sheaves is constructed using tools from complex geometry, a branch of mathematics that describes geometric curves in algebraic terms and solves geometric problems using algebraic equations. On the other hand, the Fukaya category is constructed using symplectic geometry, a branch of mathematics that arose from studies of classical physics. Symplectic geometry studies spaces equipped with a symplectic form, a mathematical tool that can be used to compute area in two-dimensional examples.
The homological mirror symmetry conjecture of Maxim Kontsevich
states that the derived category of coherent sheaves on one Calabi–Yau
manifold is equivalent in a certain sense to the Fukaya category of its
mirror.
This equivalence provides a precise mathematical formulation of mirror
symmetry in topological string theory. In addition, it provides an
unexpected bridge between two branches of geometry, namely complex and
symplectic geometry.
Another approach to understanding mirror symmetry was suggested by Andrew Strominger, Shing-Tung Yau, and Eric Zaslow in 1996.
According to their conjecture, now known as the SYZ conjecture, mirror
symmetry can be understood by dividing a Calabi–Yau manifold into
simpler pieces and then transforming them to get the mirror Calabi–Yau.
The simplest example of a Calabi–Yau manifold is a two-dimensional torus or donut shape.
Consider a circle on this surface that goes once through the hole of
the donut. An example is the red circle in the figure. There are
infinitely many circles like it on a torus; in fact, the entire surface
is a union of such circles.
One can choose an auxiliary circle (the pink circle in the figure) such that each of the infinitely many circles decomposing the torus passes through a point of . This auxiliary circle is said to parametrize the circles of the decomposition, meaning there is a correspondence between them and points of . The circle
is more than just a list, however, because it also determines how these
circles are arranged on the torus. This auxiliary space plays an
important role in the SYZ conjecture.
The idea of dividing a torus into pieces parametrized by an
auxiliary space can be generalized. Increasing the dimension from two to
four real dimensions, the Calabi–Yau becomes a K3 surface.
Just as the torus was decomposed into circles, a four-dimensional K3
surface can be decomposed into two-dimensional tori. In this case the
space is an ordinary sphere.
Each point on the sphere corresponds to one of the two-dimensional
tori, except for twenty-four "bad" points corresponding to "pinched" or singular tori.
The Calabi–Yau manifolds of primary interest in string theory have six dimensions. One can divide such a manifold into 3-tori (three-dimensional objects that generalize the notion of a torus) parametrized by a 3-sphere (a three-dimensional generalization of a sphere). Each point of
corresponds to a 3-torus, except for infinitely many "bad" points which
form a grid-like pattern of segments on the Calabi–Yau and correspond
to singular tori.
Once the Calabi–Yau manifold has been decomposed into simpler
parts, mirror symmetry can be understood in an intuitive geometric way.
As an example, consider the torus described above. Imagine that this
torus represents the "spacetime" for a physical theory. The fundamental objects of this theory will be strings propagating through the spacetime according to the rules of quantum mechanics. One of the basic dualities of string theory is T-duality, which states that a string propagating around a circle of radius is equivalent to a string propagating around a circle of radius in the sense that all observable quantities in one description are identified with quantities in the dual description. For example, a string has momentum
as it propagates around a circle, and it can also wind around the
circle one or more times. The number of times the string winds around a
circle is called the winding number. If a string has momentum and winding number in one description, it will have momentum and winding number in the dual description.
By applying T-duality simultaneously to all of the circles that
decompose the torus, the radii of these circles become inverted, and one
is left with a new torus which is "fatter" or "skinnier" than the
original. This torus is the mirror of the original Calabi–Yau.
T-duality can be extended from circles to the two-dimensional
tori appearing in the decomposition of a K3 surface or to the
three-dimensional tori appearing in the decomposition of a
six-dimensional Calabi–Yau manifold. In general, the SYZ conjecture
states that mirror symmetry is equivalent to the simultaneous
application of T-duality to these tori. In each case, the space provides a kind of blueprint that describes how these tori are assembled into a Calabi–Yau manifold.
Quantum indeterminacy is the apparent necessary incompleteness in the description of a physical system, that has become one of the characteristics of the standard description of quantum physics. Prior to quantum physics, it was thought that
a physical system had a determinate state which uniquely determined all the values of its measurable properties, and
conversely, the values of its measurable properties uniquely determined the state.
Quantum indeterminacy can be quantitatively characterized by a probability distribution on the set of outcomes of measurements of an observable.
The distribution is uniquely determined by the system state, and
moreover quantum mechanics provides a recipe for calculating this
probability distribution.
Indeterminacy in measurement was not an innovation of quantum
mechanics, since it had been established early on by experimentalists
that errors
in measurement may lead to indeterminate outcomes. By the later half of
the 18th century, measurement errors were well understood, and it was
known that they could either be reduced by better equipment or accounted
for by statistical error models. In quantum mechanics, however, indeterminacy is of a much more fundamental nature, having nothing to do with errors or disturbance.
Measurement
An
adequate account of quantum indeterminacy requires a theory of
measurement. Many theories have been proposed since the beginning of quantum mechanics and quantum measurement continues to be an active research area in both theoretical and experimental physics. Possibly the first systematic attempt at a mathematical theory was developed by John von Neumann.
The kinds of measurements he investigated are now called projective
measurements. That theory was based in turn on the theory of projection-valued measures for self-adjoint operators which had been recently developed (by von Neumann and independently by Marshall Stone) and the Hilbert space formulation of quantum mechanics (attributed by von Neumann to Paul Dirac).
In this formulation, the state of a physical system corresponds to a vector of length 1 in a Hilbert spaceH over the complex numbers. An observable is represented by a self-adjoint (i.e. Hermitian) operator A on H. If H is finite dimensional, by the spectral theorem, A has an orthonormal basis of eigenvectors. If the system is in state ψ, then immediately after measurement the system will occupy a state which is an eigenvector e of A and the observed value λ will be the corresponding eigenvalue of the equation Ae = λe.
It is immediate from this that measurement in general will be
non-deterministic. Quantum mechanics, moreover, gives a recipe for
computing a probability distribution Pr on the possible outcomes given
the initial system state is ψ. The probability is
where E(λ) is the projection onto the space of eigenvectors of A with eigenvalue λ.
Example
In this example, we consider a single spin 1/2particle
(such as an electron) in which we only consider the spin degree of
freedom. The corresponding Hilbert space is the two-dimensional complex
Hilbert space C2, with each quantum state corresponding to a unit vector in C2
(unique up to phase). In this case, the state space can be
geometrically represented as the surface of a sphere, as shown in the
figure on the right.
are self-adjoint and correspond to spin-measurements along the 3 coordinate axes.
The Pauli matrices all have the eigenvalues +1, −1.
For σ1, these eigenvalues correspond to the eigenvectors
For σ3, they correspond to the eigenvectors
Thus in the state
σ1 has the determinate value +1, while measurement of σ3 can produce either +1, −1 each with probability 1/2. In fact, there is no state in which measurement of both σ1 and σ3 have determinate values.
There are various questions that can be asked about the above indeterminacy assertion.
Can the apparent indeterminacy be construed as in fact
deterministic, but dependent upon quantities not modeled in the current
theory, which would therefore be incomplete? More precisely, are there hidden variables that could account for the statistical indeterminacy in a completely classical way?
Can the indeterminacy be understood as a disturbance of the system being measured?
Von Neumann formulated the question 1) and provided an argument why the answer had to be no, if
one accepted the formalism he was proposing. However, according to
Bell, von Neumann's formal proof did not justify his informal
conclusion. A definitive but partial negative answer to 1) has been established by experiment: because Bell's inequalities are violated, any such hidden variable(s) cannot be local (see Bell test experiments).
The answer to 2) depends on how disturbance is understood,
particularly since measurement entails disturbance (however note that
this is the observer effect,
which is distinct from the uncertainty principle). Still, in the most
natural interpretation the answer is also no. To see this, consider two
sequences of measurements: (A) which measures exclusively σ1 and (B) which measures only σ3 of a spin system in the state ψ.
The measurement outcomes of (A) are all +1, while the statistical
distribution of the measurements (B) is still divided between +1, −1
with equal probability.
Other examples of indeterminacy
Quantum
indeterminacy can also be illustrated in terms of a particle with a
definitely measured momentum for which there must be a fundamental limit
to how precisely its location can be specified. This quantum uncertainty principle
can be expressed in terms of other variables, for example, a particle
with a definitely measured energy has a fundamental limit to how
precisely one can specify how long it will have that energy.
The units involved in quantum uncertainty are on the order of Planck's constant (defined to be 6.62607015×10−34 J⋅Hz−1).
Indeterminacy and incompleteness
Quantum
indeterminacy is the assertion that the state of a system does not
determine a unique collection of values for all its measurable
properties. Indeed, according to the Kochen–Specker theorem,
in the quantum mechanical formalism it is impossible that, for a given
quantum state, each one of these measurable properties (observables)
has a determinate (sharp) value. The values of an observable will be
obtained non-deterministically in accordance with a probability
distribution which is uniquely determined by the system state. Note that
the state is destroyed by measurement, so when we refer to a collection
of values, each measured value in this collection must be obtained
using a freshly prepared state.
This indeterminacy might be regarded as a kind of essential
incompleteness in our description of a physical system. Notice however,
that the indeterminacy as stated above only applies to values of
measurements not to the quantum state. For example, in the spin 1/2
example discussed above, the system can be prepared in the state ψ by
using measurement of σ1 as a filter which retains only those particles such that σ1
yields +1. By the von Neumann (so-called) postulates, immediately
after the measurement the system is assuredly in the state ψ.
However, Einstein believed that quantum state cannot be a
complete description of a physical system and, it is commonly thought,
never came to terms with quantum mechanics. In fact, Einstein, Boris Podolsky and Nathan Rosen
showed that if quantum mechanics is correct, then the classical view of
how the real world works (at least after special relativity) is no
longer tenable. This view included the following two ideas:
A measurable property of a physical system whose value can be
predicted with certainty is actually an element of (local) reality (this
was the terminology used by EPR).
Effects of local actions have a finite propagation speed.
This failure of the classical view was one of the conclusions of the EPR thought experiment in which two remotely located observers, now commonly referred to as Alice and Bob, perform independent measurements of spin on a pair of electrons, prepared at a source in a special state called a spin singlet state. It was a conclusion of EPR, using the formal apparatus of quantum theory, that once Alice measured spin in the x direction, Bob's measurement in the x
direction was determined with certainty, whereas immediately before
Alice's measurement Bob's outcome was only statistically determined.
From this it follows that either value of spin in the x direction is not an element of reality or that the effect of Alice's measurement has infinite speed of propagation.
Indeterminacy for mixed states
We have described indeterminacy for a quantum system which is in a pure state. Mixed states are a more general kind of state obtained by a statistical mixture of pure states. For mixed states
the "quantum recipe" for determining the probability distribution of a measurement is determined as follows:
Let A be an observable of a quantum mechanical system. A is given by a densely
defined self-adjoint operator on H. The spectral measure of A is a projection-valued measure defined by the condition
for every Borel subset U of R. Given a mixed state S, we introduce the distribution of A under S as follows:
This is a probability measure defined on the Borel subsets of R which is the probability distribution obtained by measuring A in S.
Logical independence and quantum randomness
Quantum
indeterminacy is often understood as information (or lack of it) whose
existence we infer, occurring in individual quantum systems, prior to
measurement. Quantum randomness is the statistical manifestation
of that indeterminacy, witnessable in results of experiments repeated
many times. However, the relationship between quantum indeterminacy and
randomness is subtle and can be considered differently.[4]
In classical physics, experiments of chance, such as
coin-tossing and dice-throwing, are deterministic, in the sense that,
perfect knowledge of the initial conditions would render outcomes
perfectly predictable. The ‘randomness’ stems from ignorance of physical
information in the initial toss or throw. In diametrical contrast, in
the case of quantum physics, the theorems of Kochen and Specker, the inequalities of John Bell, and experimental evidence of Alain Aspect, all indicate that quantum randomness does not stem from any such physical information.
In 2008, Tomasz Paterek et al. provided an explanation in mathematical information. They proved that quantum randomness is, exclusively, the output of measurement experiments whose input settings introduce logical independence into quantum systems.
Logical independence is a well-known phenomenon in Mathematical Logic.
It refers to the null logical connectivity that exists between
mathematical propositions (in the same language) that neither prove nor
disprove one another.
In the work of Paterek et al., the researchers demonstrate a link connecting quantum randomness and logical independence
in a formal system of Boolean propositions. In experiments measuring
photon polarisation, Paterek et al. demonstrate statistics correlating
predictable outcomes with logically dependent mathematical propositions,
and random outcomes with propositions that are logically independent.
In 2020, Steve Faulkner reported on work following up on the
findings of Tomasz Paterek et al; showing what logical independence in
the Paterek Boolean propositions means, in the domain of Matrix
Mechanics proper. He showed how indeterminacy's indefiniteness
arises in evolved density operators representing mixed states, where
measurement processes encounter irreversible 'lost history' and
ingression of ambiguity.