Vous êtes sur la page 1sur 40

Quantum Field Theory

Kuhlmann, Meinard En Edward N. Zalta. The Stanford Encyclopedia of


Philosophy (Spring 2009 Edition) First published Thu Jun 22, 2006

Quantum Field Theory (QFT) is the mathematical and conceptual framework for
contemporary elementary particle physics. Since the very beginning of western philosophy
reflections about the material world which go beyond the directly observable play a central
role in philosophy. Starting with the presocratics it has always been a point of debate what
the fundamental characteristics of the material world are. Is everything constantly changing
or are there certain permanent features? What is basic and what is merely a matter of
perspective and appearance? In the course of time various answers have been given and
conflicting views have often been alternating in their predominance. QFT is presently the
best starting point for analysing the fundamental features of matter and interactions. In a
rather informal sense QFT is the extension of QM (dealing with particles) over to fields.
(See the entry on quantum mechanics.) The tools of QFT allow us to treat physical systems
that have an infinite number of degrees of freedom. Its mathematical structure allows to
analyse the creation and annihilation of particles like electrons and photons. QFT is
relativistically invariant in a way which is not possible in QM. This can easily be
demonstrated in Quantum Electrodynamics, the QFT of interactions between charged
particles and the electromagnetic field.
During the last two decades QFT became a more and more vividly discussed topic in
philosophy of physics. QFT is an attractive topic for philosophers with respect to
methodology, semantics as well as ontology. Indeed, from a methodological point of view
QFT is much more a set of formal strategies and mathematical tools than a closed theory.
Its development was accompanied by problems provoked by the application of badly
defined mathematics. Nevertheless, empirically such pragmatic approaches have been far
more successful so far than more rigorous formulations. How could such a theory work for
more than 70 years? Since mathematical reasoning dominated the heuristics of QFT, its
interpretation is open in most areas which go beyond the immediate empirical predictions.
Philosophical analysis might help to clarify its semantics. QFT taken seriously in its
metaphysical implications seems to give a picture of the world which is at variance with
central classical conceptions like particles and fields and even with some features of
quantum mechanics (QM).

1. Why Philosophy of QFT?

2. The Historial Development


o 2.1 The Early Development
o 2.2 The Emergence of Infinities
o 2.3 The Taming of Infinities

3. The Basic Structure of the Standard Formulation


o 3.1 The Lagrangian Formulation of QFT
o 3.2 Interaction
o 3.3 Gauge Invariance
o 3.4 Effective Field Theories and Renormalization
o 3.5 String Theories

4. Alternative Approaches
o 4.1 Deficiencies of the Standard Formulation of QFT
o 4.2 The Algebraic Point of View
o 4.3 Basic Ideas of AQFT

5. Philosophical Issues
o 5.1 Ontology
o 5.2 Symmetries, Heuristics, and Objectivity
o 5.3 Atomism and Reductionism

Bibliography

Other Internet Resources

Related Entries

1. Why Philosophy of QFT?


On first sight it might be surprising that the discussion on the conceptual foundations of the
quantum domain has always been primarily concerned with QM and not with QFT. After
all QFT is, in a certain sense, more comprehensive than QM and in particular
relativistically invariant in contrast to QM. There have been at least two reasons for
neglecting QFT in favour of QM regarding conceptual reflections. First, for a long time the
attitude was dominating that the decisive philosophical problems, in particular the
measurement problem, show up in QM already so that a conceptual analysis of QFT
appeared not to be necessary. It even seemed that looking at QFT would only blur the view
on the central features since QFT is much more complex and mathematically advanced than
standard QM. A second reason for neglecting QFT was the fact that QFT has not yet
reached the status of a consistent and complete theory. In addition the lack of a quantum
field theory of gravitation is felt as a pressing need. Since it cannot be excluded that the
incorporation of the fourth fundamental force might lead to deep changes of QFT as a
whole, the current version of QFT must be seen as a preliminary theory. However, if one
were to wait for its completion it is very likely that a philosophical analysis of QFT would
never even start. Interpretational reflections on the foundations of physics and its
inconsistencies might even help in the search for the final theory. In any case, some
quantum stuctures have been very steady for more than 70 years now and lead to strikingly
good predictions so that the belief is well-grounded that at least a good part of these
stuctures will remain in all improved theories.

There have been various studies on the historical development of QFT and in particular
Quantum Electrodynamics (QED). This is partly due to some charismatic figures involved,
especially Richard Feynman, and some spectacular successes of methods like
renormalization theory, Feynman diagrams and the extensive use of symmetry groups. For
the preference of historical studies on QFT over philosophical ones it might have been
more important, however, that history does not change afterwards like theories do. QFT as
an object of philosophical reflection only began to receive wider attention in the late 1980s
when the two above-mentioned arguments against QFT as a philosophical topic became
less important for the following reasons. First, careful analyses of the specifically
relativistic traits of QFT led to results, e.g., about localizability, which at least aggravate the
conceptual problems of QM severely and possibly even give rise to qualitatively new
problems. Second, due to the development of QFT and of the theory of superstrings in the
last two decades the initial hope is fading away that QFT is near to its final completion and
this fact speaks against a further postponement of philosophical analyses of QFT.

Some further arguments support the hope that a conceptual analysis of QFT will deliver
results which finally enable us to tackle problems which seemed insoluble when looking at
QM. The fundamental difficulty to find and to understand the nature of the basic entities of
the quantum regime might, looking at QFT, lead to a solution which only makes it
necessary to explain why we have the impression of elementary particles. In that case
there would be no need to take elementary particles ontologically serious. Some newer
results seem to make it almost impossible to maintain the wide-spread view that quantum
field theory is just as well a particle as a field theory, despite of its supposedly misleading
name. In particular the Reeh-Schlieder theorem and the Unruh effect seem to display
features which show that QFT is essentially a field theoryalthough not in a
straightforward classical sense. Further results like a no-go theorem by David Malament
and, e.g., Robert Wald's research on QFT in curved space-time strongly support such a
view, too.
Further Reading. One cluster of discussions on the philosophy of QFT was an
investigation by Paul Teller and Michael Redhead on different formal descriptions of many
particle systems containing identical particles, for instance in Redhead 1975, Redhead
1980, Redhead 1983, Teller 1983, Redhead 1988, Redhead & Teller 1991 and finally Teller
1995, which is the first systematic monograph on the philosophy of QFT. The anthology
Brown & Harr 1988, Philosophical Foundations of Quantum Field Theory, was the first
book to appear about this field of research. Sunny S. Auyang's book How is Quantum Field
Theory Possible? (Auyang 1995) had fewer effects than Teller's in the first time, later it
received more attention because of a new interest, first, in event ontology and, second, in
the role of symmetries in connection with the discussion on gauge theories. Cao 1999 is an

anthology with contributions by distinguished scholars from physics and philosophy (of
physics) on a wide range of general and advanced special topics, e.g., alternative
approaches, gauge invariance, renormalization, effective field theories, particle versus field
ontology etc. The anthology Kuhlmann et al. 2002 brings philosophers and philosophers of
physics together with a focus on ontological issues.

2. The Historial Development


2.1 The Early Development
The historical development of QFT is very instructive until the present day. Its first
achievement, namely the quantization of the electromagnetic field is still the paradigmatic
example of a successful quantum field theory Weinberg 1995. Ordinary QM cannot give
an account of photons which constitute the paradigmatic case of relativistic particles.
Since photons have the rest mass zero, and correspondingly travel in the vacuum at the
velocity, naturally, of light c it is ruled out that a non-relativistic theory such as ordinary
QM could give even an approximate description. Photons are implicitly contained in the
emission and absorption processes which have to be postulated, for instance, when one of
an atom's electrons makes a transition from a higher to a lower energy level or vice versa.

However, only the formalism of QFT contains an explicit description of photons. Looking
back one would say that most topics in the early development of quantum theory (1900
1927) were related with the interaction of radiation and matter and should be treated by
quantum field theoretical methods. However, the way to quantum mechanics formulated by
Dirac, Heisenberg and Schrdinger (1926/27) started from atomic spectra and did not rely
very much on problems of radiation. As soon as the conceptual framework of quantum
mechanics was developed, a small group of theoreticians immediately tried to extend the
methods to electromagnetic fields. A good example is the famous three-man paper von
M. Born, W. Heisenberg, and P. Jordan (1926). Especially P. Jordan was acquainted with
the literature on light quanta and made important contributions to QFT. The basic analogy
was that in QFT field quantities, i.e., the electric and magnetic field, should be represented
by matrices in the same way as in QM position and momentum are represented by matrices.
The ideas of QM were extended to systems having infinite degrees of freedom. The
inception of QFT is usually dated 1927 with Dirac's famous paper Dirac 1927 on The
quantum theory of the emission and absorption of radiation. Here Dirac coined the name
quantum electrodynamics (QED) which is the part of QFT that has been developed first.
Dirac supplied a systematic procedure for transferring the characteristic quantum
phenomenon of discreteness of physical quantities from the quantum mechanical treatment
of particles to a corresponding treatment of fields. Employing the quantum mechanical
theory of the harmonic oscillator, Dirac gave a theoretical description of how photons
appear in the quantization of the electromagnetic radiation field. Later, Dirac's procedure
became a model for the quantization of other fields as well. During the following three
years the first approaches to QFT were further developed. P. Jordan introduced creation

operators for fields obeying Fermi statistics. For an elementary discussion of quantum
statistics (Fermi and Bose), see the entry on quantum theory: identity and individuality.
So the methods of QFT could be applied to equations resulting from the quantum
mechanical (field like) treatment of particles like the electron (e.g., Dirac equation).
Schweber points out (Schweber 1994, p. 28) that the idea and procedure of that second
quantization goes back to Jordan, in a number of papers from 1927 (see references in
Schweber 1994, pp. 695f), while the expression itself was coined by Dirac. Some difficult
problems concerning commutation relations, statistics and Lorentz invariance could be
solved. The first comprehensive account of a general theory of quantum fields, in particular
the method of canonical quantization, was presented in Heisenberg & Pauli 1929. Whereas
the actual objects of Jordan's second quantization procedure are the coefficients of the
normal modes of the field, Heisenberg & Pauli 1929 started with the fields themselves and
subjected them to the canonical procedure. Heisenberg and Pauli thus established the basic
structure of QFT which can be found in any introduction to QFT up to the present day.
Fermi and Dirac, Fock and Podolski presented different formulations which played a
heuristic role in the following years.

Quantum electrodynamics, the historical as well as systematical entre to QFT, rests on two
pillars (see, e.g., the short and lucid Historical Introduction of Scharf's original book
Scharf 1995). The first pillar results from the quantization of the electromagmetic field, i.e.,
it is about photons as the quanta or quantized excitations of the electromagmetic field. As
Weinberg points out the photon is the only particle that was known as a field before it was
detected as a particle so that it is natural that QED began with the analysis of the radiation
field (Weinberg 1995, p. 15). The second pillar of QED consists in the relativistic theory of
the electron, with the Dirac equation in its centre.
The easiest way to quantize the electromagnetic (or: radiation) field consists of two steps.
First, one Fourier analyses the vector potential of the classical field into normal modes
(using periodic boundary conditions) corresponding to an infinite but denumerable number
of degrees of freedom. Second, since each mode is described independently by a harmonic
oscillator equation, one can apply the harmonic oscillator treatment from non-relativistic
quantum mechanics to each single mode. The result for the Hamiltonian of the radiation
field is
(2.1)

Hrad =

( ar(k)ar(k) + 1/2

where ar(k) and ar(k) are operators which satisfy the following commutation relations
(2.2)

[ar(k), as(k)]

rskk

),

[ar(k), as(k)]

[ar(k), as(k)] = 0.

with the index r labeling the polarisation. These commutation relations imply that one is
dealing with a bosonic field.
The operators ar(k) and ar(k) as well as their product ar(k)ar(k) have interesting physical
interpretations as so-called particle creation and annihilation operators. In order to see this
one has to examine the eigenvalues of the operators
(2.3) Nr(k) = ar(k)ar(k)
which are the essential parts in Hrad. Due to the commutation relations (2.2) one finds that
the eigenvalues of Nr(k) are the integers nr(k) = 0, 1, 2, and the corresponding
eigenfunctions (up to a normalisation factor) are
(2.4) |nr(k)> = [ar(k)]nr(k)|0>
where the right hand side means that ar(k) operates nr(k) times on |0>, the state vector of
the vacuum with no photons present.

The interpretation of these results is parallel to the one of the harmonic oscillator. ar(k) is
interpreted as the creation operator of a photon with momentum k and energy k (and a
polarisation which depends on r and k). That is, equation (2.4) can be understood in the
following way. One gets a state with nr(k) photons of momentum k and energy k when
the creation operator ar(k) operates nr(k) times on the vacuum state |0>. Accordingly, Nr(k)
is called the number operator and nr(k) the occupation number of the mode that is
specified by k and r, i.e., this mode is occupied by nr(k) photons. Note that Pauli's
exclusion principle is not violated since it only applies to fermions and not to bosons like
photons. The corresponding interpretation for the annihilation operator ar(k) is parallel,
when it operates on a state with a given number of photons this number is lowered by one.
It is a widespread view (see e.g., Ryder 1996, p. 131) that these results complete the
justification for interpreting N(k) as the number operator, and hence for the particle
interpretation of the quantized theory. This is a rash judgement, however. For instance, the
question of localizability or at least approximate localizability is not even touched while it
is certain that this is a pivotal criterion for something to be a particle. All that is established
so far is a certain discreteness of physical quantities which is one feature of particles.
However, this is not yet conclusive evidence for a particle interpretation of QFT. Recalling
how various entities of forerunner theories, e.g., the single particle wave function, gained a
very different meaning in QFT one should be extremely cautious to jump to one's
conclusions when only certain aspects are rediscovered in QFT. It is not clear at this stage

whether we are in fact talking about particles or about fundamentally different objects
which only have this one feature of discreteness in common with particles.

2.2 The Emergence of Infinities


Quantum field theory started with a theoretical framework that was built in analogy to
quantum mechanics. Although there was no unique and fully developed theory, quantum
field theoretical tools could be applied to concrete processes. Examples are the scattering of
radiation by free electrons (Compton scattering), the collision between relativistic
electrons or the production of electron-positron pairs by photons. Calculations to the first
order of approximation were quite successful, but most people working in the field thought
that QFT still had to undergo a major change. On the one side some calculations of effects
for cosmic rays clearly differed from measurements. On the other side and, from a
theoretical point of view more threatening, calculations of higher orders of the perturbation
series led to infinite results. The self-energy of the electron as well as vacuum fluctuations
of the electromagnetic field seemed to be infinite. The perturbation expansions did not
converge to a finite sum and even most individual terms were divergent.

The various forms of infinities suggested that the divergences were more than failures of
specific calculations. Many people tried to avoid the divergences by formal tricks
(truncating the integrals at some value of momentum, or even ignoring infinite terms) but
such rules were not reliable, violated the requirements of relativity and were not considered
as satisfactory. Some people came up with first ideas of coping with infinities by a
redefinition of the parameters of the theory and using a measured finite value (for example
of the charge of the electron) instead of the infinite bare value (renormalization).
From the point of view of philosophy of science it is remarkable that these divergences did
not give enough reason to discard the theory. The years from 1930 to the beginning of
World War II were characterized by a variety of attitudes towards QFT. Some physicists
tried to circumvent the infinities by more-or-less arbitrary prescriptions, others worked on
transformations and improvements of the theoretical framework. Most of the theoreticians
believed that QED would break down at high energies. There was also a considerable
number of proposals in favour of alternative approaches. These proposals included changes
in the basic concepts (e.g., negative probabilities), interactions at a distance instead of a
field theoretical approach, and a methodological change to phenomenological methods that
focusses on relations between observable quantities without an analysis of the
microphysical details of the interaction (the so-called S-matrix theory where the basic
elements are amplitudes for various scattering processes).

Despite the feeling that QFT was imperfect and lacking rigour, its methods were extended
to new areas of applications. In 1933 Fermi's theory of the beta decay started with
conceptions describing the emission and absorption of photons, transferred them to beta
radiation and analyzed the creation and annihilation of electrons and neutrinos (weak
interaction). Further applications of QFT outside of quantum electrodynamics succeeded in
nuclear physics (strong interaction). In 1934 a new type of fields (scalar fields) described
by the Klein-Gordon equation could be quantized (another example of second
quantization). This new theory for matter fields could be applied a decade later when new
particles, pions, were detected.

2.3 The Taming of Infinities


After the end of World War II reliable and effective methods for dealing with infinities in
QFT were developed, namely coherent and systematic rules for performing relativistic field
theoretical calculations, and a general renormalization theory. On three famous conferences
between 1947 and 1949 developments in theoretical physics were confronted with relevant
new experimental results. In the late forties there were two different ways to address the
problem of divergences. One of these was discovered by Feynman, the other one (based on
an operator formalism) by Schwinger and independently by Tomonaga. In 1949 Dyson
showed that the two approaches are in fact equivalent. Thus, Freeman Dyson, Richard P.
Feynman, Julian Schwinger and Sin-itiro Tomonaga became the inventors of
renormalization theory.

The most spectacular experimental successes of renormalization theory were the


calculations of the anomalous magnetic moment of electron and the Lamb shift in the
spectrum of hydrogen. These successes were so outstanding because the theoretical results
were in better agreement with high precision experiments than anything in physics before.
The basic idea of renormalization is to avoid divergences that appear in physical
predictions by shifting them into a part of the theory where they do not influence empirical
propositions. Dyson could show that a rescaling of charge and mass (renormalization) is
sufficient to remove all divergences in QED to all orders of perturbation theory. In general,
a QFT is called renormalizable, if all infinities can be absorbed into a redefinition of a finite
number of coupling constants and masses. A consequence is that the physical charge and
mass of the electron must be measured and cannot be computed from first principles.
Perturbation theory gives well defined predictions only in renormalizable quantum field
theories, and luckily QED, the first fully developed QFT belonged to the class of
renormalizable theories. There are various technical procedures to renormalize a theory.
One way is to cut off the integrals in the calculations at a certain value of the momentum
which is large but finite. This cut-off procedure is successful if, after taking the limit
, the resulting quantities are independent of (Part II of Peskin & Schroeder 1995 gives
an extensive description of renormalization).

Feynman's formulation of QED is of special interest from a philosophical point of view. His
so-called space-time approach is visualized by the famous Feynman diagrams that look like
depicting paths of particles. Feynman's method of calculating scattering amplitudes is based
on the functional integral formulation of field theory (for an introduction to the theory and
practice of Feynman diagrams see, e.g., chapter 4 in Peskin & Schroeder 1995). A set of
graphical rules can be derived so that the probability of a specific scattering process can be
calculated by drawing a diagram of that process and then using the diagram to write down
the mathematical expressions for calculating its amplitude. The diagrams provide an
effective way to organize and visualize the various terms in the perturbation series, and they
seem to display the flow of electrons and photons during the scattering process. External
lines in the diagrams represent incoming and outgoing particles, internal lines are
connected with virtual particles and vertices with interactions. Each of these graphical
elements is associated with mathematical expressions that contribute to the amplitude of the
respective process. The diagrams are part of Feynman's very efficient and elegant algorithm
for computing the probability of scattering processes. The idea of particles travelling from
one point to another was heuristically useful in constructing the theory, and moreover, this
intuition is useful for concrete calculations. Nevertheless, an analysis of the theoretical
justification of the space-time approach shows that its success does not imply that particle
paths have to be taken seriously. General arguments against a particle interpretation of QFT
clearly exclude that the diagrams represent paths of particles in the interaction area.
Feynman himself was not particularly interested in ontological questions.

In the beginning of the 1950s QED became a reliable theory which had left behind the
preliminary status. It took two decades from writing down the first equations until QFT
could be applied to interesting physical problems in a systematic way. The new
developments made it possible to apply QFT to new particles and new interactions. In the
following decades QFT was extended to describe not only the electromagnetic force but
also weak and strong interaction so that new Lagrangians had to be found which contain
new classes of particles or quantum fields. The research aimed at a more comprehensive
theory of matter and in the end at a unified theory of all interactions. New theoretical
concepts had to be introduced, mainly connected with non-Abelian gauge theories (see
below, section 2.3) and spontaneous symmetry breaking. See also the entry on symmetry
and symmetry breaking. Today there are trustworthy theories of the strong, weak, and
electromagnetic interactions of elementary particles which have a similar structure as QED.
A combined theory associated with the gauge group SU(3) SU(2) U(1) is considered
as the standard model of elementary particle physics which was achieved by Glashow,
Weinberg and Salam in 1962. According to the standard model there are three families of
quarks and leptons, each of them containing 15 particles/fields with spin 1/2 (for example
various quarks, the electron and its neutrino, or the muon and its neutrino). In addition it
contains terms for the photon and other spin 1 particles/fields describing the forces between
quarks and leptons. Alltogether there is good agreement to experimental data, for example
the masses of W+ and W bosons (detected in 1983) confirmed the theoretical prediction
within one per cent deviation.

Further Reading. The first chapter in Weinberg 1995 is a very good short description of
the earlier history of QFT. Detailed accounts of the historical development of QFT can be
found, e.g., in Darrigol 1986, Schweber 1994 and Cao 1997a. Various historical and
conceptual studies of the standard model are gathered in Hoddeson et al. 1997 and of
renormalization theory in Brown 1993.

3. The Basic Structure of the Standard Formulation


3.1 The Lagrangian Formulation of QFT
The crucial step towards quantum field theory is in some respects analogous to the
corresponding quantization in quantum mechanics by imposing the commutation relations.
Its starting point is the classical Lagrangian formulation of mechanics, which is a so-called
analytical formulation as opposed to the standard version of Newtonian mechanics. A
generalized notion of momentum (the conjugate or canonical momentum) is defined by
setting p = L/q, where L is the Lagrange function L = T V (T is the kinetic energy and V
the potential) and q dq/dt. This definition can be motivated by looking at the special case
of a Lagrange function with a potential V which depends only on the position so that (using
Cartesian coordinates) L/x = (/x)(
mx2/2) = mx = px. Under these conditions the
generalized momentum coincides with the usual mechanical momentum.

In classical Lagrangian field theory one associates with the given field a second field,
namely the conjugate field
(3.1) = L/
where L is a Lagrangian density. The field and its conjugate field are the direct
analogues of the canonical coordinate q and the generalized (canonical or conjugate)
momentum p in classical mechanics of point particles.
In both cases, QM and QFT, requiring that the canonical variables satisfy certain
commutation relations implies that the basic quantities become operator valued. From a
physical point of view this shift implies a restriction of possible measurement values for
physical quantities some (but not all) of which can have their values only in discrete steps
now. In QFT the canonical commutation relations for a field and the corresponding
conjugate field are
(3.2)

[(x,t), (y,t)]
[(x,t), (y,t)]

=
=

i3(x y)
[(x,t), (y,t)] = 0

which are equal-time commutation relations, i.e., the commutators always refer to fields at
the same time. It is not obvious that the equal-time commutation relations are Lorentz

invariant but one can formulate a manifestly covariant form of the canonical commutation
relations. If the field to be quantized is not a bosonic field, like the Klein-Gordon field or
the electromagnetic field, but a fermionic field, like the Dirac field for electrons one has to
use anticommutation relations.
In very loose terms, the operator valuedness of quantum fields means that to each spacetime point (x,t) a field value (x,t) is assigned which is an operator. This is the fundamental
difference to classical fields because an operator valued quantum field (x,t) does not by
itself correspond to definite values of a physical quantity like the strength of the
electromagnetic field. On this background, Teller has argued in Teller 1995 that the field
interpretation of QFT is inappropriate since the alleged fields in QFT are not to be
interpreted as physical fields with definite values of some sort which are assigned to spacetime points, like in the case of the classical electromagnetic field. Rather, quantum fields
are what Teller calls determinables (p. 95), as it becomes manifest by the fact that
quantum fields are described by mappings from space-time points to operators. Operators
are mathematical entities which are defined by how they act on something. They do not
represent definite values of quantities but they specify what can be measured, therefore
Teller's expression determinables. (Below it will be discussed why this talk in terms of a
field at a point has to be refined using the notion of a smeared field (f).)

While there are close analogies between quantization in QM and in QFT there are also
important differences. Whereas the commutation relations in QM refer to a quantum object
with three degrees of freedom, so that one has a set of 15 equations, the commutation
relations in QFT do in fact comprise an infinite number of equations, namely for each 4tuple (x,t) there is a new set of commutation relations and there is, of course, a continuous
set of space-time points (x,t). This infinite number of degrees of freedom embodies the field
character of quantum field theory.
Regarding differences between QM and QFT it is important to realize that the operator
valued field (x,t) in QFT is not analogous to the wavefunction (x,t) in QM, i.e., the
quantum mechanical state in its position representation. Although in the development of
QFT there is a continuity from the wave function, i.e., the quantum mechanical state in its
position representation, to the field in QFT, it would be a misconception to understand these
two quantities as analogues. Here, the ontologically relevant formal setting has changed in
the transition from QM to QFT. While the wavefunction in QM is acted upon by
observables, i.e., by operators, it is the (operator valued) field in QFT which itself acts on
the space of states, i.e., on the states which are associated with the quantum field. In a
certain sense one can say that the single particle wave functions have been transformed, via
their reinterpretation as operator valued quantum fields, into observables. This step is
sometimes called second quantization because the single particle wave equations in
relativistic QM already came about by a quantization procedure, e.g., in the case of the

Klein-Gordon equation by replacing position and momentum by the corresponding


quantum mechanical operators. Afterwards the solutions to these single particle wave
equations, which are states in relativistic QM, are considered as classical fields which can
be subjected to the canonical quantization procedure of QFT. The term second
quantization has often been critized partly because it blurs the important fact that the single
particle wave function in relativistic QM and the operator valued quantum field are
fundamentally different kinds of entities despite their connection in the context of
discovery.
Summing up one can say that although in both, QM and QFT, the two fundamental
irreducible kinds of entities are states on the one side and observables on the other side this
fact is overshadowed by two confusing aspects in a comparison of QM and QFT. First,
quantum fields which one expects to be somehow physically concrete like classical fields
are on the side of observables although, as far as the development of theories is concerned,
they are the successors of states (in their position representation, namely wave functions),
e.g., in the Klein-Gordon equation of relativistic QM as described above. The second
confusing aspect is that states are constantly mentioned in QM, whereas most of QFT is
about quantum fields. States, which are comparatively abstract entities in QFT with no
immediate spatio-temporal meaning, seem to be an appendage. Nevertheless, both states
and observables are equally important in QM and QFT as the two fundamental kinds of
entities.

3.2 Interaction
Up to this point, the aim was to develop a free field theory. Doing so does not only neglect
interaction with other particles (fields), it is even unrealistic for one free particle because it
interacts with the field that it generates itself. For the description of interactionssuch as
scattering in particle colliderswe need certain extensions and modifications of the
formalism as so far exposed. The immediate contact between scattering experiments and
QFT is given by the scattering or S-matrix which contains all the relevant predictive
information about, e.g., scattering cross sections. In order to calculate the S-matrix the
interaction Hamiltonian is needed. The Hamiltonian can in turn be derived from the
Lagrangian density by means of the so-called Legendre transformation.
In order to discuss interactions one introduces a new representation, the so-called
interaction picture which is an alternative to the Schrdinger and the Heisenberg picture.
For the interaction picture one splits up the Hamiltonian, which is the generator of timetranslations, into two parts H = H0 + Hint, where H0 describes the free system, i.e., without
interaction, and gets absorbed in the definition of the fields and Hint is the interaction part of
the Hamiltonian, or short the interaction Hamiltonian. Using the interaction picture is
advantageous because the equations of motion as well as, under certain conditions, the
commutations relations are the same for interacting fields as for free fields. Therefore,

various results that were established for free fields can still be used in the case of
interacting fields. The central instrument for the description of interaction is again the Smatrix, which expresses the connection between in and out states by specifying the
transition amplitudes. In QED, for instance, a state |in> describes one particular
configuration of electrons, positrons and photons, i.e., it describes how many of these
particles there are and which momenta, spins and polarizations they have before the
interaction. The S-matrix supplies the probability that this state goes over to a particular |
out> state, e.g., that a particular counter responds after the interaction. Such probabilities
can be checked in experiments.
The canonical formalism of QFT as introduced in the previous section is only applicable in
the case of free fields since the inclusion of interaction leads to infinities (see the historical
part). For this reason perturbation theory makes up a large part of most publications on
QFT. The importance of perturbative methods is understandable realizing that they
establish the immediate contact between theory and experiment. Although the techniques of
perturbation theory have become ever more sophisticated it is somewhat disturbing that
perturbative methods could not be avoided even in principle. One reason for this unease is
that perturbation theory is felt to be rather a matter of (highly sophisticated) craftsmanship
than of understanding nature. Accordingly, the corpus of perturbative methods plays a small
role in the philosophical investigations of QFT. What does matter, however, is in which
sense the consideration of interaction effects the general framework of QFT. An overview is
given in Section 4.1 (Perturbation TheoryPhilosophy and Examples) of Peskin &
Schroeder 1995.

3.3 Gauge Invariance


Some theories are distinguished by the mathematical property of gauge invariance which
means that transformations, so-called gauge transformations, of certain terms do not change
the observable quantities. The requirement of gauge invariance has the mathematical
advantage that it provides an elegant way to introduce terms for interacting fields.
Moreover, requiring gauge invariance plays an important role for the selection of theories.
The prime example of an intrinsically gauge invariant theory is the theory of the
electromagnetic field. It is well-known from the classical theory that Maxwell's equations
can be stated in terms of the vector potential A and the scalar potential or in terms of the
4-vector potential A = (, A). The link to the electric field E(x,t) and the magnetic field
B(x,t) is given by
(3.3)

B
E

or covariantly
(3.4) F = A A

=
=

A
(A/t)

where F is the electromagnetic field tensor. The important point in the present context is
that given the identification (3.3) or (3.4), there remains a certain flexibility or freedom in
the choice of A and , or A. In order to see that, consider the so-called gauge
transformations
(3.5)

A
+ /t

or covariantly
(3.6) A A +
where is a scalar function (of space and time or of space-time) which can be chosen
arbitrarily. Inserting the transformed potential(s) into equation(s) (3.3), or (3.4), one can see
that the electric field E and the magnetic field B, or covariantly the electromagnetic field
tensor F, are not effected by a gauge transformation of the potential(s). Since only the
electric field E and the magnetic field B, and quantities constructed from them, are
observable, whereas the vector potential itself is not, nothing physical seems to be changed
by a gauge transformation because it leaves E and B unaltered. Note that gauge invariance
is a kind of symmetry that does not come about by space-time transformations.

In order to link the notion of gauge invariance to the Lagrangian formulation of QFT one
needs a more general form of gauge transformations which operates on the field operator
and which is supplied by
(3.7)

ei
ei*

where is an arbitrary real constant. Equations (3.7) describe a global gauge


transformation in contrast to a local gauge transformation
(3.8)

(x)

ei(x)(x)

which can vary with x. The requirement of invariance under a local gauge transformation is
essential for finding the equations describing fundamental interaction. Take for example the
Lagrangian for a free electron. The requirement that the Lagrangian should be locally
invariant under the same type of transformation can only be fulfilled by introducing
additional terms. The form of these terms is determined by the symmetry requirement,
which results in the introduction of the electromagnetic field. In a sense, the
electromagnetic field is a consequence of the local symmetry of the Lagrangian for the
electron.

This procedure can be generalized to more complex transformations (for example referring
to mixing the components of field operators) and new interactions. By requiring local gauge
invariance additional fields can be introduced. These additional fields describe the
interaction between the original fields. The gauge principle provides a general schema for
introducing interaction by constructing gauge field theories. To this end one starts with a
Lagrangian for a matter field and derives the interaction by introducing exactly those fields
that make the Lagrangian invariant under a relevant local gauge transformation. It seems
that all fundamental forces can be described by such local gauge field theories.
Gauge symmetry plays a crucial role in determining the dynamics of the theory since the
nature of gauge transformation determines the possible interaction. The structure of these
transformations are characterized by special mathematical groups: U(1) for QED, SU(2)
U(1) for electroweak interaction, SU(3) for strong interaction. The relations between these
groups are exploited in programs for the unification of the fundamental types of interaction.
There is also a strong analogy to general relativity where a local gauge symmetry is
associated with the gravitational field. From a more technical point of view gauge
symmetries are important tools in proofs of renormalizability. The upshot is that the
fulfillment of gauge invariance has an importance for the selection of theories which makes
gauge invariance a player in the same league as Lorentz invariance. Since gauge invariance
plays a pivotal role in the discovery of quantum field theories it is a paradigm case for how
a rich mathematical structure can help in the construction of theories.

General introductions to gauge theories can be found in Cao 1997a and Schweber 1994.
Auyang emphasizes the general conceptual significance of invariance principles in her
book Auyang 1995 while Redhead 2002 as well as Martin 2002 focus specifically on gauge
symmetries. Lyre 2004b is a study of the significance of gauge theories for the debate on
structural realism, with a related paper Lyre 2004a in English. The ontological significance
of gauge potentials is discussed in particular with respect to the Aharanov-Bohm effect,
e.g., in Healey 2001.

3.4 Effective Field Theories and Renormalization


In the 1970s a program emerged in which the theories of the standard model of elementary
particle physics are considered as effective field theories (EFTs) which have a common
quantum field theoretical framework. EFTs describe relevant phenomena only in a certain
domain since the Lagrangian contains only those terms that describe particles which are
relevant for the respective range of energy. EFTs are inherently approximative and change
with the range of energy considered. EFTs are only applicable on a certain energy scale,
i.e., they only describe phenomena in a certain range of energy. Influences from higher
energy processes contribute to average values but they cannot be described in detail. This
procedure has no severe consequences since the details of low-energy theories are largely
decoupled from higher energy processes. Both domains are only connected by altered

coupling constants and the renormalization group describes how the coupling constants
depend on the energy.
The main idea of EFTs is that theories, i.e., in particular the Langrangians, depend on the
energy of the phenomena which are analysed. The physics changes by switching to a
different energy scale, e.g., new particles can be created if a certain energy threshold is
exceeded. The dependence of theories from the energy scale distinguishes QFT from, e.g.,
Newton's theory of gravitation where the same law applies to an apple as well as to the
moon. Nevertheless, laws from different energy scales are not completely independent from
each other. A central aspect of considerations about this dependence are the consequences
of higher energy processes on the low-energy scale.
On this background a new attitude towards renormalization developed in the 1970s which
revitalizes earlier ideas that divergences result from neglecting unknown processes of
higher energies. Low-energy behaviour is thus affected by higher energy processes. Since
higher energies correspond to smaller distances this dependence is to be expected from an
atomistic point of view. According to the reductionistic program the dynamics of
constituents on the microlevel should determine processes on the macrolevel, i.e., here the
low-energy processes. However, as, for instance hydrodynamics shows, in practice theories
from different levels are not quite as closely connected because a law which is applicable
on the macrolevel can be largely independent from microlevel details. For this reason
analogies with statistical mechanics play an important role in the discussion about EFTs.

The basic idea of this new story about renormalization is that the influences of higher
energy processes are localisable in a few structural properties which can be captured by an
adjustment of parameters. In this picture, the presence of infinities in quantum field theory
is neither a disaster, nor an asset. It is simply a reminder of a practical limitationwe do
not know what happens at distances much smaller than those we can look at directly.
(Georgi 1989, p. 456) This new attitude supports the view that renormalization is the
appropriate answer to the change of fundamental interactions when the QFT is applied to
processes on different energy scales. The price one has to pay is that EFTs are only valid in
a limited domain and should be considered as approximations to better theories on higher
energy scales. This prompts the important question whether there is a last fundamental
theory in this tower of EFTs which supercede each other with rising energies. Some people
conjecture that this deeper theory could be a string theory, i.e., a theory which is not a field
theory any more. Or should one ultimately expect from physics theories that they are only
valid as approximations and in a limited domain?

3.5 String Theories


Up to now string theory is the most promising candidate for bridging the gap between QFT
and general relativity theory, thus supplying a unified theory of all natural forces, including

gravitation. The basic idea of string theory is not to take particles as fundamental objects
but strings which are very small but extended in one dimension. This assumption has the
pivotal consequence that strings interact on an extended distance and not at a point. This
difference between string theory and standard QFT is essential because it is the reason why
string theory also encompasses the gravitational force which cannot be treated in the
framework of QFT. Gravitation is so hard to be reconciled with QFT because the typical
length scale of the gravitational force is very small, namely at Planck scale, so that the
quantum field theoretical assumption of point-like interaction leads to untreatable infinities.
To put in another way, gravitation becomes significant (in particular in comparison to
strong interaction) exactly where QFT is most severely endangered by infinite quantities.
The extended interaction of strings brings it about that such infinities can be avoided. In
contrast to the entities in standard quantum physics strings are not characterized by
quantum numbers but only by their geometrical and dynamical properties. Nevertheless,
macroscopically strings look like quantum particles with quantum numbers. A basic
geometrical distinction is the one between open strings, i.e., strings with two ends, and
closed strings which are like bracelets. The central dynamical property of strings is their
mode of excitation, i.e., how they vibrate.
Reservations about string theory are mostly due to the lack of testability since it seems that
there are no empirical consequences which could be tested by the methods which are, at
least up to now, available to us. The raeson for this problem is that the length scale of
strings is in the average the same as the one of quantum gravity, namely the Planck length
of approximately 1033 centimeters which lies far beyond the accessibility of feasible
particle experiments. But there are also other peculiar features of string theory which might
be hard to swallow.

One of them is the fact that string theory implies that space-time has 10, 11 or even 26
dimensions. In order to explain the appearance of only four space-time dimensions string
theory assumes that the other dimensions are somehow folded away or compactified so
that they are no longer visible. An intuitive idea can be gained by thinking of a macaroni
which is a tube, i.e., a two-dimensional piece of pasta rolled together, but which looks from
the distance like a one-dimensional string.
Despite of the problems of string theory, physicists do not abandon this project, partly
because there seem to be no better candidates for a reconciliation of quantum physics and
general relativity theory with the possible exception of the so-called loop quantum
gravity (see the entry on quantum gravity). Correspondingly, string theory has also
received some attention within the philosophy of physics community in recent years. One
philosophical investigation of string theory is Weingard 2001 in Callender & Huggett 2001,
an anthology with further related articles. Another more recent study is Dawid 2003 (see
the Other Internet Resources section below). Dawid argues that string theory has significant
consequences for the philosophical debate about realism, namely that it speaks against the
plausibility of anti-realistic positions.

Two of the standard introductory monographs to string theory are Polchinski 2000 and
Kaku 1999. A very successful popular introduction is Greene 1999. An interactive website
with a nice elementary introduction is Stringtheory.com (see the Other Internet Resources
section below).

4. Alternative Approaches
4.1 Deficiencies of the Standard Formulation of QFT
From the 1930s onwards the problem of infinities as well as the potentially heuristic status
of the Lagrangian formulation of QFT stimulated the search for reformulations in a concise
and eventually axiomatic manner. A number of further aspects intensified the unease about
the standard formulation of QFT. The first one is that quantities like total charge, total
energy or total momentum of a field are unobservable since their measurement would have
to take place in the whole universe. Accordingly, quantities which refer to infinitely
extended regions of space-time should not appear among the observables of the theory as
they do in the standard formulation of QFT. Another problematic feature of standard QFT is
the idea that QFT is about field values at points of space-time. The mathematical aspect of
the problem is that a field at a point, (x), is not an operator in a Hilbert space. The physical
counterpart of the problem is that it would require an infinite amount of energy to measure
a field at a point of spacetime. One way to handle this situationand one of the starting
points for axiomatic reformulations of QFTis not to consider fields at a point but instead
fields which are smeared out in the vicinity of that point using certain functions, so-called
test functions. The result is a smeared field (f) = (x)f(x)dx with supp(f) O, where
supp(f) is the support of the test function f and O is a bounded open region in Minkowski
space-time.
The third important problem for standard QFT which prompted reformulations is the
existence of inequivalent representations. In the context of quantum mechanics,
Schrdinger, Dirac, Jordan and von Neumann realized that Heisenberg's matrix mechanics
and Schrdinger's wave mechanics are just two (unitarily) equivalent representations of the
same underlying abstract structure, i.e., an abstract Hilbert space H and linear operators
acting on this space. In 1931 von Neumann gave a detailed proof (of a conjecture by Stone)
that the canonical commutation relations (CCRs) for position coordinates and their
conjugate momentum coordinates in configuration space fix the representation of these two
sets of operators in Hilbert space up to unitary equivalence (von Neumann's uniqueness
theorem). This means that the specification of the purely algebraic CCRs suffices to
describe a particular physical system. In quantum field theory, however, von Neumann's
uniqueness theorem looses its validity since here one is dealing with an infinite number of
degrees of freedom. Now one is confronted with a multitude of inequivalent irreducible
representations of the CCRs and it is not obvious what this means physically and how one
should cope with it.

4.2 The Algebraic Point of View

The described situation is the background for the establishment of algebraic reformulations
of QFT. According to the algebraic point of view algebras of observables rather than
observables themselves should be taken as the basic entities in the mathematical description
of quantum physics. In the forties the mathematician i.e., Segal postulated (Segal 1947a,
1947b) that the C*-algebra generated by all bounded operators should be the basic entity in
the mathematical description of physics. It turned out that the adoption of an algebraic point
of view could be the appropriate framework in order to handle all the above-mentioned
problems of standard QFT, the representation problem for instance. In standard QM the
algebraic point of view in terms of C*-algebras makes no notable difference to the usual
Hilbert space formulation since the Hilbert space representation and the C*-algebra
formalism are equivalent. However, in QFT this is no longer the case as described above.
Since in QFT one is dealing with an infinite number of degrees of freedom there are
unitarily inequivalent irreducible representations of a C*-algebra. Sticking to the usual
Hilbert space formulation thus means to make an implicit choice of one particular
representation that is not equivalent to other available representations. Segal proposed to
take one single C*-algebra as the basic element of the mathematical description of a
quantum physical system and he dismissed the availability of inequivalent representations
as irrelevant to physics. Against this approach Haag, the most outstanding advocate of
algebraic quantum field theory (AQFT), argued that inequivalent representations can be
understood physically by a revision of Segal's approach. The decisive idea is not to take
individual C*-algebras as the basic ingredients of the description but rather a so-called net
of algebras (see below) since the relation between various algebras itself represents
important physical information. Another point where algebraic formulations are
advantageous derives from the fact that two quantum fields are physically equivalent when
they generate the same algebras of local observables. Such equivalent quantum field
theories belong to the same so-called Borchers class which entails that they lead to the
same S-matrix.

As Haag stresses, in Haag 1996, fields are only an instrument in order to coordinatize
observables, more precisely, in order to coordinatize the sets of observables with respect to
different finite space-time regions (the so-called net of local algebras in AQFT). The choice
of a field system is to a certain degree conventional, namely as long as it belongs to the
same Borchers class. From this point of view it is more appropriate to consider these
algebras as the fundamental entities in QFT rather than quantum fields.
In the fifties there was a strong tendency, in physics as well as in other fields, to
reformulate grown theories in an axiomatic manner, partly in order to remove ad hoc
features. A very prominent early attempt to axiomatise QFT is Arthur Wightman's field
axiomatics. Wightman imposed axioms on polynomial algebras P(O) of smeared fields, i.e.,
sums of products of smeared fields. A crucial point of Wightman's approach is the
replacement of the mapping x (x), which supposedly expresses what is meant by a
field, by the mapping O P(O) from finite space-time regions O to P(O). Wightman's
smeared out field operators are unbounded which makes the approach cumbersome from a
mathematical point of view and this is one of the differences to the approach I will
introduce next where only bounded operators are considered.

Algebraic Quantum Field Theory (AQFT) is arguably the most successful attempt to
reformulate QFT in an axiomatic manner. AQFT originated in the late fifties by the work of
Rudolf Haag and quickly advanced in collaboration with Huzihiro Araki and Daniel
Kastler. AQFT itself exists in two versions, concrete AQFT (Haag-Araki) and abstract
AQFT (Haag-Kastler). The concrete approach uses von Neumann algebras (or W*algebras), the abstract one C*-algebras. The adjective abstract refers to the fact that in this
approach the algebras are characterized in an abstract fashion and not by explicitly using
operators on a Hilbert space. In standard QFT, the CCRs together with the field equations
can be used for the same purpose, i.e., an abstract characterization. One common aim of
these axiomatisations of QFT is to avoid the usual approximations in standard QFT. Trying
to do this in a strictly axiomatic way, however, they only get reformulations which are not
as rich as standard QFT from a physical point of view. The algebraic approach [] has
given us a frame and a language not a theory as Haag concedes in Haag 1996 with respect
to AQFT.

4.3 Basic Ideas of AQFT


One of the main traits and possibly the most unusual one of AQFT is that so-called nets of
algebras are seen as the primary objects of study. The idea is that the physical information
in quantum field theories is not contained in individual algebras but in the mapping O
A(O) from spacetime regions O to algebras A(O) of local observables where the O's are
open and bounded regions in Minkowski spacetime. Since only finite regions are
considered, the algebras are called local algebras. Physically, the elements of an algebra
A(O) are seen as representing operations that can be performed in the region O that is
associated with the algebra.

The crucial point is that it is not necessary to specify observables explicitly in order to fix
physically meaningful quantities. The very way how algebras of local observables are
linked to spacetime regions is sufficient to supply observables with physical significance. It
is the partition of the so-called algebra Aloc of all local observables into subalgebras which
contains physical information about the observables, i.e., it is the net structure of algebras
which matters. The claim is that the allocation itself of observable algebras to finite spacetime regions suffices to account for the physical meaning of observables. It is not necessary
to start with any such information explicitly.
The physical justification for this approach consists in the recognition that the experimental
data for QFT are exclusively space-time localization properties of microobjects from which
other properties are inferred. The Stern-Gerlach experiment is an illuminating example. All
one gets in this experiment are certain space-time distributions of dots of detected particles
which originated from a particle source and hit a photographic plate. Only in a second step
one recognices particles with certain spin directions after having passed an inhomogeneous
magnetic field. This example might help to imagine that space-time localisation can specify
or encode all other physical properties.

Physically the most important notion of AQFT is the principle of locality which has an
external as well as an internal aspect. The external aspect is the fact that AQFT considers
only observables connected with finite regions of space-time and not global observables
like the total charge or the total energy momentum vector which refer to infinite space-time
regions. This approach was motivated by the operationalistic view that QFT is a statistical
theory about local measurement outcomes with all the experimental information coming
from measurements in finite space-time regions. Accordingly everything is expressed in
terms of local algebras of observables. The internal aspect of locality is that there is a
constraint on the observables of such local algebras: All observables of a local algebra
connected with a space-time region O are required to commute with all observables of
another algebra which is associated with a space-time region O that is space-like separated
from O. This principle of (Einstein) causality is the main relativistic ingredient of AQFT.
The basic structure upon which the assumptions or conditions of AQFT are imposed are
local observables, i.e., self-adjoint elements in local (non-commutative) von Neumannalgebras, and physical states, which are identified as positive, linear, normalized functionals
which map elements of local algebras to real numbers. States can thus be understood as
assignments of expectation values to observables. One can group the assumptions of AQFT
into relativistic axioms, such as locality and covariance, general physical assumptions, like
isotony and spectrum condition, and finally technical assumptions which are closely related
to the mathematical formulation.

As a reformulation of QFT, AQFT is expected to reproduce the main phenomena of QFT, in


particular properties which are characteristic of it being a field theory, like the existence of
antiparticles, internal quantum numbers, the relation of spin and statistics, etc. That this aim
could not be achieved within AQFT on a purely axiomatic basis is partly due to the fact that
the connection between the respective key concepts of AQFT and QFT, i.e., observables
and quantum fields, is not sufficiently clear. It turned out that the main link between the
theory of local observables and the quantum fields of standard QFT is the notion of
superselection. Superselection rules are certain restrictions on the set of all observables and
allow for classification schemes in terms of permanent or essential properties.
Comprehensive introductions to AQFT are provided by the monographs Haag 1996 and
Horuzhy 1990 as well as the overview articles Haag & Kastler 1964, Roberts 1990 and
Buchholz 1998. Early pioneering monographs on axiomatic QFT are Streater & Wightman
1964 and Bogolubov et al. 1975. Mathematical aspects are emphasized in Bratteli &
Robinson 1979.

It is only since the second half of the eighties that AQFT came into the focus of the
philosophy of physics community. Some of the most fruitful discussions were stimulated by
reexaminations of physical theorems from the sixties and seventies, in particular the ReehSchlieder theorem. Further results of interest are Haag's theorem and a lemma by Borchers.
The properties of the relativistic vacuum often play a central role in these discussions. Cf.,
e.g., Redhead 1995a, Redhead 1995b and the monograph, The Philosophy of Vacuum
(Saunders & Brown 1991). A mayor reason for this interest is that the vacuum displays
features which seem to be in conflict with standard ontological conceptions about what
things are and how things and properties are related to each other. Some central issues in
the philosophical debate about AQFT are questions about locality, localization and causality
(e.g., Saunders 1988, Redhead & Wagner 1998, Ruetsche 2003, Redei & Summers 2002 as
well as numerous fine papers by Rob Clifton, collected in Butterfield & Halvorson 2004).

5. Philosophical Issues
5.1 Ontology
Ontology is concerned with the most general features, entities and structures of being. One
can pursue ontology in a very general sense or with respect to a particular theory or a
particular part or aspect of the world. With respect to the ontology of QFT one is tempted to
more or less dismiss ontological enquiries and to adopt the following straightforward view.
There are two groups of fundamental fermionic matter constituents, two groups of bosonic
force carriers and four (including gravitation) kinds of interactions. As satisfying as this
answer might first appear, the ontological questions are, in a sense, not even touched.

Saying that, for instance the down quark is a fundamental constituent of our material world
is the starting point rather than the end of the (philosophical) search for an ontology of
QFT. The main question is what kind of entity, e.g., the down quark is. The answer does not
depend on whether we think of down quarks or muon neutrinos since the sought features
are much more general than those ones which constitute the difference between down
quarks or muon neutrinos. The relevant questions are of a different type. What are particles
at all? Can quantum particles be legitimately understood as particles any more, even in the
broadest sense, when we take, e.g., their localization properties into account? How can one
spell out what a field is and can quantum fields in fact be understood as fields? Could it
be more appropriate not to think of, e.g., quarks, whatever they are, as the most
fundamental entities at all, but rather of properties or processes or events?
5.1.1 The Particle Interpretation of QFT
Many of the creators of QFT can be found in one of the two camps regarding the question
whether particles or fields should be given priority in understanding QFT. While Dirac, the
later Heisenberg, Feynman, and Wheeler opted in favor of particles, Pauli, the early

Heisenberg, Tomonaga and Schwinger put fields first (see Landsman 1996). Today, there
are a number of arguments which prepare the ground for a proper discussion beyond mere
preferences.
It seems almost impossible to think of QFT without thinking of particles which are
accelerated and scattered in colliders. Nevertheless, it is this very interpretation which has
the most fully developed arguments against it. There still is the option to say that our
classical concept of a particle is too narrow and that we have to loosen some of its
constraints. Even in classical corpuscular theories of matter the concept of an (elementary)
particle is not as unproblematic as one might expect. For instance, if the whole charge of a
particle was contracted to a point, an infinite amount of energy would be stored in this
particle since the repulsive forces become infinitely large when two charges with the same
sign are brought together. The so-called self energy of a point particle is infinite.
Probably the most immediate trait of particles is their discreteness. Particles are countable
individuals in contrast to a liquid or a mass. Obviously this characteristic alone cannot
constitute a sufficient condition for being a particle since there are other things which are
countable as well without being particles, e.g., money or maxima and minima of the
standing wave of a vibrating string. It seems that the so-called primitive thisness or
haecceity is missing to make up a sufficient condition for a particle, i.e., it must be possible
to say that it is this or that particle which has been counted in order to account for the
fundamental difference between ups and downs in a wave pattern and particles particles
(see also the entry on quantum theory: identity and individuality). In Teller 1995 primitive
thisness as well as other possible features of the particle concept are discussed in
comparison to classical concepts of fields and waves as well as in comparison to the
concept of field quanta. A critical discussion of Teller's reasoning can be found in Seibt
2002.

There is still another feature which is commonly taken to be pivotal for the particle concept,
namely that particles are localizable in space. While it is clear from classical physics
already that the requirement of localizability need not refer to point-like localization, it will
turn out later in this section that even localizability in an arbitrarily large but still finite
region can be a strong condition for quantum particles.
Eventually, there are some potential ingredients of the particle concept which are explicitly
opposed to the corresponding (and therefore opposite) features of the field concept.
Whereas it is a core characteristic of a field that it is a system with an infinite number of
degrees of freedom, the very opposite holds for particles. A particle can for instance be
referred to by the specification of the coordinates x(t) that pertain, e.g., to its center of mass
presupposing impenetrability. A further feature of the particle concept is connected to the
last point and again explicitly in opposition to the field concept. In a pure particle ontology
the interaction between remote particles can only be understood as an action at a distance.
In contrast to that, in a field ontology, or a combined ontology of particles and fields, local
action is implemented by mediating fields. Finally, classical particles are massive and
impenetrable, again in contrast to (classical) fields.

The concept of particles has been evolving through history in accordance with the latest
scientific theories. Therefore, considering the tenability of a particle interpretation for QFT
is not quite such a straightforward issue as one might think. Confronted with features of
QFT that are in conflict with the above-mentioned features of particles one can either
abandon the particle interpretation or one adjusts the particle concept, as has been done
before.
Wigner's famous analysis of the Poincar group, Wigner 1939, is often assumed to provide
a definition of elementary particles. The main idea of Wigner's approach is the supposition
that each irreducible (projective) representation of the relevant symmetry group yields the
state space of one kind of elementary physical system, where the prime example is an
elementary particle which has the more restrictive property of being structureless. The
physical justification for linking up irreducible representations with elementary systems is
the requirement that there must be no relativistically invariant distinction between the
various states of the system (Newton & Wigner 1949). In other words the state space of an
elementary system shall have no internal structure with respect to relativistic
transformations. Put more technically, the state space of an elementary system must not
contain any relativistically invariant subspaces, i.e., it must be the state space of an
irreducible representation of the relevant invariance group. If the state space of an
elementary system had relativistically invariant subspaces then it would be appropriate to
associate these subspaces with elementary systems. The requirement that a state space has
to be relativistically invariant means that starting from any of its states it must be possible
to get to all the other states by superposition of those states which result from relativistic
transformations of the state one started with. The main part of Wigner's analysis consists in
finding and classifying all the irreducible representations of the Poincar group. Doing that
involves finding relativistically invariant quantities that serve to classify the irreducible
representations.

Eugene Wigner's pioneering identification of types of particles with irreducible unitary


representations of the Poincar group has been exemplary until the present, as it is
emphasized, e.g., in Buchholz 1994.
Regarding the question whether Wigner has supplied a definition of particles one can say
that although Wigner has in fact found a highly valuable and fruitful classification of
particles his analysis does not contribute very much to the question what a particle is and
whether a given theory can be interpreted in terms of particles. What Wigner has given is
rather a conditional answer. If relativistic quantum mechanics can be interpreted in terms of
particles then the possible types of particles correspond to irreducible unitary
representations of the Poincar group. However, the question whether, and if yes in what
sense, at least relativistic quantum mechanics can be interpreted as a particle theory at all is
not addressed in Wigner's analysis. For this reason the discussion of the particle
interpretation of QFT is not finished with Wigner's analysis as one might be tempted to say.
For instance the pivotal question of the localizability of particle states is still open.
5.1.1.1 Localization Problems

The observed particle traces, e.g., on photographic plates of bubble chambers, seem to be
a clear indication for the existence of particles. On the other hand, however, the theory
which has been built on the basis of these scattering experiments, viz., QFT, turns out to
have considerable problems to account for the observed `particle trajectories'. Not only are
sharp trajectories excluded by Heisenberg's uncertainty relations for position and
momentum coordinates which hold for non-relativistic quantum mechanics already. More
advanced theoretical examinations in AQFT which will be described and scrutinized below,
show that quantum particles which behave according to the principles of relativity theory
cannot be localized in any bounded region of space-time, no matter how large, a result
which excludes even tube-like tajectories. From this theoretical point of view it thus
appears to be impossible that our world is composed of particles when we assume that
localizability in some finite region of space-time is a necessary ingredient of the particle
concept. Surprisingly, the very theory which excludes localizability is remarkably good in
predicting experiments which apparently involve localizable particles. For the working
physicist this contradiction is not an important issue because it does not cause any
problems, neither for the theoretical nor the experimental physicist, as long as conceptual
questions as such are not at stake.
So far there is no unquestioned proof against the possibility of a particle interpretation for
QFT. Although there are N-particle states among the possible states of QFT it is not clear
how these states relate to N particles. The pieces of circumstantial evidence against a
particle interpretation seem to be strong. The core of these pieces consists in problems to
localize particle states in any sensible way. Reeh & Schlieder, Hegerfeldt, Malament and
Redhead all gained mathematical results, or formalized their interpretation, which prove
that certain sets of assumptions, which are taken to be essential for the particle concept,
lead to contradictions. However, it is a point of debate what exactly has been shown by
these no-go theorems in the end and how the different results relate to one another.
The Reeh-Schlieder theorem (Reeh & Schlieder 1961) is a central analyticity result in
AQFT. From a physical point of view the Reeh-Schlieder theorem is based on vacuum
correlations. What the Reeh-Schlieder theorem asserts is that acting on the vacuum state
with elements of the von Neumann algebra R(O), containing observables associated with
space-time region O, one can approximate as closely as one likes any state in Hilbertspace
H, in particular one that is very different from the vacuum in some space-like separated
region O. The Reeh-Schlieder theorem is thus clearly exploiting long distance correlations
of the vacuum. Or one can express the result by saying that local measurements do not
allow for a distinction between an N-particle state and the vacuum state. The technical
statement of the theorem together with introductory and interpretive comments can be
found, e.g., in Redhead 1995a. Redhead's interpretation of the Reeh-Schlieder theorem is
that local measurements can never decide whether one observes an N-particle state since a
projection operator P which corresponds to an N-particle state can never be an element
of a local algebra R(O) (cf. Redhead 1995a). The consequences of the Reeh-Schlieder
theorem for the issue of entanglement are discussed in Clifton & Halvorson 2001.
Malament's no-go theorem, Malament 1996, is another consequence of analyticity which
rests on a lemma of Borchers (1967). In short it says that a relativistic quantum theory of a
fixed number of particles predicts a zero probability for finding a particle in any spatial set

if four conditions are satisfied, namely concerning translation covariance, energy,


localizability and locality. The localizability condition is the essential ingredient of the
particle concept: A particlein contrast to a fieldcannot be found in two disjoint spatial
sets at the same time. The locality condition is the main relativistic part of Malament's
assumptions, it requires that the statistics for measurements in one space-time region must
not depend on whether or not a measurement has been performed in a space-like related
second space-time region. Malament's proof has the weight of a no-go theorem provided
that we acccept his four conditions as natural assumptions for a particle interpretation. A
relativistic quantum theory of a fixed number of particles, satisfying in particular the
localizability and the locality condition, has to assume a world devoid of particles (or at
least a world in which particles can never be detected) in order not to contradict itself.
Malament's no-go theorem thus seems to show that there is no middle ground between QM
and QFT, i.e., no theory which deals with a fixed number of particles (like in QM) and
which is relativistic (like QFT) without running into the localizability problem of the no-go
theorem. One is forced towards QFT which, as Malament is convinced, can only be
understood as a field theory. Nevertheless, whether or not a particle interpretation of QFT is
in fact ruled out by Malament's result is a point of debate. At least prima facie Malament's
no-go theorem alone cannot supply a final answer since it assumes a fixed number of
particles, an assumption that is not valid in the case of QFT.
The results about non-localizability which have been explored above may appear to be not
very astonishing in the light of the following facts about ordinary QM: Quantum
mechanical wave functions (in position representation) are usually smeared out over all 3,
so that everywhere in space there is a non-vanishing probability for finding a particle. This
is even the case arbitrarily close after a sharp position measurement due to the
instantaneous spreading of wave packets over all space. Note, however, that ordinary QM is
non-relativistic.

A conflict with SRT would thus not be very surprising although it is not yet clear whether
the above-mentioned quantum mechanical phenomena can actually be exploited to allow
for superluminal signalling. QFT, on the other side, has been designed to be in accordance
with special relativity theory (SRT). The local behaviour of phenomena is one of the
leading principles upon which the theory was built. This makes non-localizability within
the formalism of QFT a much severer problem for a particle interpretation.
Malament's reasoning has come under attack in Fleming & Butterfield 1999 and Busch
1999. Both argue to the effect that there are alternatives to Malament's conclusion. The
main line of thought in both criticisms is that Malament's mathematical result might just
as well be interpreted as evidence that the assumed concept of a sharp localization operator
is flawed and has to be modified either by allowing for unsharp localization (Busch 1999)
or for so-called hyperplane dependent localization (Fleming & Butterfield 1999). In
Saunders 1995 a different conclusion from Malament's (as well as from similar) results is
drawn. Rather than granting Malament's four conditions and deriving a problem for a
particle interpretation Saunders takes Malament's proof as further evidence that one can not
hold on to all four conditions. According to Saunders it is the localizability condition which

might not be a natural and necessary requirement on second thought. Stressing that
relativity requires the language of events, not of things Saunders argues that the
localizability condition loses its plausibility when it is applied to events: It makes no sense
to postulate that the same event can not occur at two disjoint spatial sets at the same time.
One can only require for the same kind of event not to occur at both places. For Saunders
the particle interpretation as such is not at stake in Malament's argument. The question is
rather whether QFT speaks about things at all. Saunders considers Malament's result to give
a negative answer to this question. A kind of meta paper on Malament's theorem is Clifton
& Halvorson 2002. Various objections to the choice of Malament's assumptions and his
conclusion are considered and rebutted. Moreover, Clifton and Halvorson establish two
further no-go theorems which preserve Malament's theorem by weakening tacit
assumptions and showing that the general conclusion still holds. One thing seems to be
clear. Since Malament's mathematical result appears to allow for various different
conclusions it cannot be taken as conclusive evidence against the tenability of a particle
interpretation of QFT and the same applies to Redhead's interpretation of the ReehSchlieder theorem.
5.1.1.2 Further Problems for a Particle Interpretation of QFT

The standard definition for the vacuum state |0> is that it is the energy ground state, i.e., the
eigenstate of the energy operator with the lowest eigenvalue. It is a notable result in
ordinary non-relativistic QM that the ground state energy of e.g., the harmonic oscillator is
not zero in contrast to its analogue in classical mechanics. Not only is the same true for the
vacuum state in QFT, the relativistic vacuum of QFT displays even more striking features.

The expectation values for various quantities do not vanish for the vacuum state. The label
|0> does not indicate that the energy is zero in the vacuum state. It rather stems from the
assumption that there are no particles present in the vacuum state: an N-particle state can be
built up from the vacuum state by the N-fold application of a creation operator. Nonvanishing vacuum expectation values prompt the question what it is that has these values or
gives rise to them if the vacuum is taken to be the state with no particles present. Since the
vacuum state |0> is closely linked to N-particle states where N is not zero, properties of the
vacuum state have a great impact on the particle interpretation as such. If particles were the
basic objects about which QFT speaks how can it be that there are physical phenomena
even if nothing is there according to this very ontology?
An even greater related challenge for a particle interpretation of QFT is the Unruh effect.
The Unruh effect is a surprising result which seems to show that the concept of a particle is
observer dependent. The Unruh effect is the striking phenomenon that a uniformly
accelerated observer in a Minkowski vacuum will detect a thermal bath of particles, the socalled Rindler quanta (Unruh 1976 and Unruh & Wald 1984). Whereas the number of

particles in the Minkowski vacuum is 0, an accelerated observer suddenly detects a thermal


bath of particles. A mere change of the frame of reference thus leads to a change of the
number of particles. Since basic features of a theory should be invariant under
transformations of the referential frame the Unruh effect constitutes a severe challenge to
the concept of particles as basic objects of QFT. Teller tries to show, in Teller 1995, that the
Unruh effect is not a fundamental problem for a particle interpretation. Further extensive
studies have been done by H. Halvorson, partly together with R. Clifton, reprinted in
Butterfield & Halvorson 2004 together with various other important papers.
Eventually, studies of QFT in curved space-time show that the particle concept hinges on
Poincar symmetry. This result indicates that the existence of a particle number operator
might be a contingent property of the flat Minkowski space-time. In flat space-time
Poincar symmetry is used to pick out a preferred representation of the canonical
commutation relations which is equivalent to picking out a preferred vacuum state. This
leads to the standard notion of a particle. However, neither the existence of global families
of inertial observers nor the Poincar tranformations which relate between these families
can be generalized to curved space-time. QFT in curved space-time can actually teach us
something about standard QFT (in flat space-time). Since QFT in flat space-time is a
special case of QFT in curved space-time, QFT in curved space-time can help us to see
what is contingent in QFT in flat space-time.
5.1.1.3 Results

On the one side, the adoption of a particle interpretation of QFT would make the
importance of particle experiments and the predominance of speaking in terms of particles
comprehensible. It could explain why charge only exists in discrete amounts which is a
typical feature of particles and not continuously which is characteristic for fields. On the
other side, we saw that there are various problems for a particle interpretation.

Some results indicate that particle states cannot be localized in any finite region of spacetime no matter how large it is. Other results show that the particle number might not be an
objective feature. Nevertheless, it turned out that most arguments need to be seen in relative
terms. At this stage of research it can only be recorded that there are various potential
threats for a particle interpretation.
5.1.2 The Field Interpretation of QFT
Classical Newtonian mechanics is formulated as a theory about bodies and forces with
action at a distance. It is only stated which force bodies exert on each other while nothing
is said about how these effects are mediated since it is assumed that there is an
instantaneous interaction between two massive bodies. It turned out that the
electromagnetic interaction between charged bodies cannot be described within this
framework. A mediating field, the electromagnetic field, had to be introduced which
accounts for the local transmission of electromagnetic forces. The systematic and efficient
formulation of the theory of electromagnetism with Maxwell's equations at its core revealed

another famous feature. There is a limiting velocity for the transmission of signals, namely
the velocity of light. In classical electromagnetism the existence of a limiting velocity for
the transmission of signals simply emerged from this theory which rests on observed
electromagnetic phenomena. Einstein established this feature as a requirement for any
physical theory, hence the term Einstein causality.
While the concept of fields as mediators for the transmission of forces is intuitively helpful,
the standard definition of a field is somewhat different. A field is generally defined as a
system with an infinite number of degrees of freedom for which certain field equations
must hold. A point particle, in contrast, can be described by its position x(t) which changes
as the time t progresses so that, in a three-dimensional space, there are three degrees of
freedom for the motion of a point particle corresponding to the three coordinates of the
particle's position. In the case of a field the description is more complex since one needs a
specification of a field value for each point x in space where this specification can change
as the time t progresses. A field is therefore specified by (x,t), i.e., a (time-dependent)
mapping from each point of space to a field value. Whereas the intuitive notion of a field is
that it is something transient and fundamentally different from matter, it is perfectly normal
in physics to ascribe energy and even momentum to a pure field where no particles are
present. This surprising feature shows how gradual the distinction between fields and
matter can be.
There are two lines of argumentation which are often taken to show that an ontology of
fields is the appropriate construal of the most fundamental entities to which QFT refers.
The first argumentation rests on the fact that so-called field operators are at the base of the
mathematical formalism of QFT. The other line of argumentation is indirect. Since various
arguments seem to exclude a particle interpretation, the allegedly only alternative, namely a
field interpretation, must be the right conception.
The transition from a classical field theory (like electromagnetism) to quantum field theory
can be characterized by the transition from the field (x,t) to the quantum field ( x,t), and a
corresponding transition for its conjugate field for both of which a certain specification of
canonical commutation relations holds. In difference to a classical field (x,t), the basic
fields ( x,t) of QFT are operator-valued fields since to each point of space and time an
operator is attached. There is a formal analogy between classical and quantum fields: field
values are attached to space-time points where these values are real-valued in the case of
classical fields and operator-valued in the case of quantum fields. In technical terms the
analogy reads as one between the mappings x
(x,t), x 3, and x
( x,t), x 3.
This formal analogy between classical and quantum fields is one reason why QFT is taken
to be a field theory. However, it has to be examined whether this formal analogy actually
justifies this conclusion.
In his paper What the quantum field is not (Teller 1990, also chapter five in Teller 1995),
Teller criticizes this conclusion. Teller argues that quantum fields lack an essential feature
of classical field theories so that the expression quantum field is only justified on a
perverse reading of the notion of a field. His reason for this conclusion is that in the case
of quantum fieldsin contrast to classical fieldsthere are no definite physical values
whatsoever assigned to space-time points. Instead, the assigned quantum field operators

represent the whole spectrum of possible values so that they rather have the status of
observables (Teller: determinables) or general solutions. Something physical emerges
only when the state of the system or when initial and boundary conditions are supplied.
Teller's criticism of the standard gloss about operator-valued quantum fields has one
justified and one unjustified aspect. The justified aspect is that quantum fields actually
differ considerably from classical fields since the field values which are attached to spacetime points have no direct physical significance in the case of a quantum field. However, it
was not to be expected anyway that one would only encounter definite values for physical
quantities in QFT since it is, like QM, an inherently probabilistic theory after all and is
equally confronted with the measurement problem.
Nevertheless, even taking the probabilistic character of QFT into account there still is the
problem that one needs quantum fields as well as state vectors in order to fix probabilistic
properties. But one should not conclude that quantum fields are not physically significant at
all since physical significance of field quantities in QFT cannot be judged along classical
distinctions. There is considerable physical information encoded in quantum fields but the
field character of QFT is by no means as obvious as it first seems. The formal analogy
between classical and quantum fields as such is not a fully convincing argument for a field
interpretation of QFT. If a field interpretation should actually yield the appropriate ontology
for QFT than it seems that those objects which are called quantum fields are not already
the fundamental entities one is looking for, at least not alone. Teller's own proposal is an
ontology of QFT in terms of field quanta. Teller argues that the Fock space representation
or occupation number representation suggests this conception with objects (quanta)
which can be counted or aggregated but which cannot be numbered. The number of objects
is given by the degree of excitation of a certain mode of the underlying field.

Particle labels like the ones in the Schrdinger many-particle formalism do not occur any
more. Teller has been criticized to draw such far-reaching ontological conclusions from one
particular representation, in particular since the Fock space representation cannot be
appropriate in general because it is only valid for free particles.
5.1.3 Alternative Ontologies
On the background of various problems with particle as well as field interpretations of QFT
there are a number of proposals for alternative ontological approaches. In Auyang 1995 and
Dieks 2002 different versions of event ontologies are proposed. Seibt 2002 and Httich
2004 defend process-ontological accounts of QFT. A critical examination of a process
ontological understanding of QFT is given in Kuhlmann 2000 and also in Kuhlmann 2002
where an interpretation in terms of so-called tropes (i.e., properties understood as
particulars) is proposed as a better alternative. Cao argues, e.g., in Cao 1997b, that the best
ontological access to QFT is gained by concentrating on structural properties rather than on
any particular category of entities.

5.2 Symmetries, Heuristics, and Objectivity


Symmetries play a central role in QFT. In order to characterize a special symmetry one has
to specify transformations T and features, which remain unchanged during these
transformations (invariants I), symmetries are thus pairs {T, I}. The basic idea is that the
transformation change elements of the mathematical description (the Lagrangians for
instance) whereas the empirical content of the theory is unchanged. There are space-time
transformations and so-called internal transformations. Whereas space-time symmetries are
universal, i. e., they are valid for all interactions, internal symmetries characterize special
sorts of interaction (strong, electromagnetic or weak interaction). Symmetry
transformations define properties of particles/quantum fields which are conserved if the
symmetry is not broken. The invariance of a system defines a conservation law, e.g., if a
system is invariant under translations the linear momentum is conserved, if it is invariant
under rotation the angular momentum is conserved. Inner transformations such as so-called
gauge transformations are connected with more abstract properties.
Symmetries are not only defined for Lagrangians but they can also be found in empirical
data and phenomenological descriptions. Symmetries can thus bridge the gap between
descriptions which are close to empirical results (phenomenology) and the more abstract
general theory which is a most important reason for their heuristic force. If a conservation
law is found one has some knowledge about the system even if details of the dynamics are
unknown. The analysis of many high energy collision experiments led to the assumption of
special conservation laws for abstract properties like baryon number or strangeness.
Evaluating experiments in this way allowed for a classification of particles. This
phenomenological classification was good enough to predict new particles which could be
found in the experiments.
Free places in the classification could be filled even if the dynamics of the theory (for
example the Lagrangian of strong interaction) was yet unknown. As the history of QFT for
strong interaction shows considerable constraints on the construction of dynamics follow
from symmetries found in the phenomenological description: The Lagrangian should not
exhibit less symmetries than the phenomenology. Arguments from group theory played a
decisive role in the unification of fundamental interactions. In addition, symmetries bring
about substantial technical advantages. For example, by using gauge transformations one
can bring the Lagrangian into a form which makes it easy to prove the renormalizability of
the theory. See also the entry on symmetry and symmetry breaking.
Symmetries are not only instruments of provisional physics used in not yet fully developed
theories. Symmetries also supply some sort of justification, they are often used in the
beginning of a chain of explanation. To a remarkable degree the present theories of
elementary particle interactions can be understood by deduction from general principles.
Under these principles symmetry requirements play a crucial role in order to determine the
Lagrangian. For example, the only Lorentz invariant and gauge invariant renormalizable
Lagrangian for photons and electrons is precisely the original Dirac Lagrangian. In this way
symmetry arguments acquire an explanatory power and help to minimize the unexplained
basic assumptions of a theory. Heisenberg concludes that in order to find the way to a real

understanding of the spectrum of particles it will therefore be necessary to look for the
fundamental symmetries and not for the fundamental particles. (Blum et al. 1995, p. 507).
Since symmetry operations change the perspective of an observer but not the physics an
analysis of the relevant symmetry group can yield very general information about those
entities which are unchanged by transformations. Such an invariance under a symmetry
group is a necessary (but not sufficient) requirement for something to belong to the
ontology of the considered physical theory. Hermann Weyl propagated the idea that
objectivity is associated with invariance (see, e.g., his authoritative work Weyl 1952, p.
132). Symmetries help to separate objective facts from the conventions of descriptions (see
Kosso in Brading & Castellani 2003).
Symmetries are typical examples for such abstract mathematical structures that show much
more continuity in scientific change than assumptions about the entities of a theory (light as
particles, as waves, as quantum fields). For that reason structural realists consider abstract
structures as the best candidate for what is true about a physical theory (Redhead 1999,
p. 34). Physical entities (like electrons) are similar to fictions and, in the end, should not be
taken seriously. In the epistemic variant of structural realism structure is all we know about
nature whereas the objects which are related by structures might exist but they are not
accessible to us. In the ontic variant of structural realism nature seems to be reduced to
mathematical objects. In the world, there is nothing but structure.
Symmetry considerations are of central importance in Auyang 1995 where the connection
between properties of physically relevant symmetry groups and ontological questions is
stressed. A recent anthology with various philosophical studies about symmetries in physics
is Brading & Castellani 2003. For an account and defence of structural realism see
Ladyman 1998.

5.3 Atomism and Reductionism


There is a coherent tradition of philosophical thinking about nature from early ancient times
to our days. One indication for this coherence is the fact that some of the most outstanding
twentieth century physicists like Schrdinger and Heisenberg put considerable emphasis on
the linkage between quantum physics and ancient Greek philosophy. The monographs
Schrdinger 1954 and Heisenberg 1958 are just two of their explictly philosophical works.
Atomism can be seen as one form of reductionism with its assertion that everything can be
reduced to some basic building blocks. In order to clarify possible ways to understand this
assertion and some of its consequences it is helpful to distinguish different notions of
reductionism. One reductionist position is that all scientific theories can and should be
reduced to a fundamental theory which is generally taken to be found in physics. Another
reductionist position is more modest. It agrees with the first reductionist view that the
reduction of higher-level theories to lower-level theories is possible in principle. However,
the modest reductionist thinks that a reduction to the lowest possible level is often neither
practically feasible nor even desirable. The question whether QFT is an atomistic theory
hinges on the understanding of atomism. The answer could be yes, if atomism is
understood as the availability of reductionist explanations, and no, if atomism is

understood as the claim that all there is are particles and the void. However, looking at
effective field theories, even the yes with respect to reductionism could turn into a no
since there is no fundamental level of particles/fields any more (see Cao & Schweber
1993). Further discussions of the significance of effective field theories for the question of
reductionism can be found in Hartmann 2001 and Castellani 2002.

Bibliography

Auyang, S. Y., 1995, How is Quantum Field Theory Possible?, Oxford-New York:
Oxford University Press.

Blum, W., Drr, H.-P., and Rechenberg, H., eds., 1985, Werner Heisenberg:
Collected Works, The Philosophical Background of Modern Physics, Mnchen:
Piper.

Bogolubov, N. N., Logunov, A. A., and Todorov, I. T., 1975, Introduction to


Axiomatic Quantum Field Theory, Reading, MA: Benjamin.

Born, M., Heisenberg, W., and Jordan, P., 1926, "Zur Quantenmechanik II",
Zeitschr. fur Physik 35, 557.

Brading, K. and Castellani, E., eds., 2003, Symmetries in Physics: Philosophical


Reflections, Cambridge: Cambridge University Press.

Bratteli, O. and Robinson, D. W., 1979, Operator Algebras and Quantum Statistical
Mechanics 1: C* and W*-Algebras, Symmetry Groups, Decomposition of States,
New York-Heidelberg-Berlin: Springer

Brown, H. R. and Harr, R., eds., 1988, Philosophical Foundations of Quantum


Field Theory, Oxford: Clarendon Press.

Brown, L. M., ed., 1993, Renormalization: from Lorentz to Landau (and Beyond),
Berlin-New York: Springer.

Buchholz, D., 1994, On the manifestations of particles, in R. N. Sen and A.


Gersten, eds., Mathematical Physics Towards the 21st Century, Beer-Sheva: BenGurion University Press.

Buchholz, D., 1998, Current trends in axiomatic qantum field theory, in P.


Breitenlohner and D. Maison, eds, Quantum Field Theory. Proceedings of the
Ringberg Workshop Held at Tegernsee, Germany, 21-24 June 1998 On the
Occasion of Wolfhart Zimmermann's 70th Birthday, Lecture Notes in Physics, Vol.
558, pp. 43-64, Berlin-Heidelberg: Springer.

Busch, P., 1999, Unsharp localization and causality in relativistic quantum theory,
Journal of Physics A: Mathematics General, 32: 6535.

Butterfield, J. and Halvorson, H., eds., 2004, Quantum Entanglements Selected


Papers Rob Clifton, Oxford: Oxford University Press.

Butterfield, J. and Pagonis, C., eds., 1999, From Physics to Philosophy, Cambridge:
Cambridge University Press.

Callender, C. and Huggett, N., eds., 2001, Physics Meets Philosophy at the Planck
Scale, Cambridge: Cambridge University Press.

Cao, T. Y., 1997a, Conceptual Developments of 20th Century Field Theories,


Cambridge: Cambridge University Press.

Cao, T. Y., 1997b, Introduction: Conceptual issues in QFT, in Cao 1997a, pp. 127.

Cao, T. Y., ed., 1999, Conceptual Foundations of Quantum Field Theories,


Cambridge: Cambridge University Press.

Cao, T. Y., and Schweber, S. S., 1993, The conceptual foundations and the
philosophical aspects of renormalization theory, Synthese, 97: 33-108.

Castellani, E., 2002, Reductionism, emergence, and effective field theories,


Studies in History and Philosophy of Modern Physics, 33: 251-267.

Clifton, R., ed., 1996, Perspectives on Quantum Reality: Non-Relativistic,


Relativistic, and Field-Theoretic, The University of Western Ontario Series in
Philosophy of Science, Dordrecht-Boston-London: Kluwer Academic Publishers.

Clifton, R. and Halvorson, H., 2001, Entanglement and open systems in algebraic
quantum field theory, Studies in History and Philosophy of Modern Physics, 32: 131; reprinted in Butterfield & Halvorson 2004.

Clifton, R. and Halvorson, H., 2002, No place for particles in relativistic quantum
theories? Philosophy of Science, 69: 1-28; reprinted in Butterfield & Halvorson
2004 and in Kuhlmann et al. 2002.

Darrigol, O., 1986, The origin of quantized matter waves, Historical Studies in
the Physical and Biological Sciences, 16: 197-253.

Davies, P., ed., 1989, The New Physics, Cambridge: Cambridge University Press.

Dieks, D., 2002, Events and covariance in the interpretation of quantum field
theory, in Kuhlmann et al. 2002, pp. 215-234.

Dirac, P. A. M., 1927, The quantum theory of emission and absorption of


radiation, Proceedings of the Royal Society of London, A 114: 243-256.

Faye, J., Scheffler, U., and Urchs, M., eds., 2000, Events, Facts, and Things,
Volume 72 of the series Pozna Studies in the Philosophy of the Sciences and
Humanities, Amsterdam: Rodopi.

Fleming, G. N. and Butterfield, J., 1999, Strange positions, in Butterfield &


Pagonis 1999, pp. 108-165.

Georgi, H., 1989, Effective quantum field theories, in Davies 1989, pp. 446-457.

Greene, B., 1999, The Elegant Universe. Superstrings, Hidden Dimensions and the
Quest for the Ultimate Theory, New York: W. W. Norton and Company.

Haag, R., 1996, Local Quantum Physics: Fields, Particles, Algebras, 2nd edition,
Berlin-Heidelberg-New York: Springer.

Haag, R. and Kastler, D., 1964, An algebraic approach to quantum field theory,
Journal of Mathematical Physics, 5: 848-861.

Hartmann, S., 2001, Effective field theories, reductionism, and explanation,


Studies in History and Philosophy of Modern Physics, 32: 267-304.

Httich, F., 2004, Quantum Processes A Whiteheadian Interpretation of Quantum


Field Theory, Mnster: agenda Verlag.

Healey, R., 2001, On the reality of gauge potentials, Philosophy of Science, 68/4:
432-455.

Heisenberg, W., 1958, Physics and Philosophy. The Revolution in Modern Science,
Introduction by F. S. C. Northrop, Epilogue by Ruth Nanda Anshen, New York:
Harper and Row.

Heisenberg, W. and Pauli, W., 1929, Zur Quantendynamik der Wellenfelder,


Zeitschrift fr Physik, 56: 1-61.

Hoddeson, L., Brown, L., Riordan, M., and Dresden, M., eds., 1997, The Rise of the
Standard Model: A History of Particle Physics from 1964 to 1979, Cambridge:
Cambridge University Press.

Horuzhy, S. S., 1990, Introduction to Algebraic Quantum Field Theory, 1st edition,
Dordrecht-Boston-London: Kluwer Academic Publishers.

Kaku, M., 1999, Introduction to Superstrings and M-Theory, New York: Springer.

Kastler, D., ed., 1990, The Algebraic Theory of Superselection Sectors: Introduction
and Recent Results, Singapore-London-Hackensack, NJ: World Scientific
Publishing Company.

Kuhlmann, M., 2000, Processes as objects of quantum field theory, in Faye et al.
2000, pp. 365-388.

Kuhlmann, M., 2002, Analytical ontologists in action: A comment on Seibt and


Simons, in Kuhlmann et al. 2002, pp. 99-109.

Kuhlmann, M., Lyre, H., and Wayne, A., eds., 2002, Ontological Aspects of
Quantum Field Theory,. Singapore-London-Hackensack, NJ: World Scientific
Publishing Company.

Ladyman, J., 1998, What is structural realism? Studies in History and Philosophy
of Science, 29: 409-424.

Landsman, N. P., 1996, Local quantum physics, Studies in History and


Philosophy of Modern Physics, 27: 511-525.

Lyre, H., 2004a, Holism and structuralism in U(1) gauge theory, Studies in
History and Philosophy of Modern Physics, 35/4: 643-670.

Lyre, H., 2004b, Lokale Symmetrien und Wirklichkeit. Eine naturphilosophische


Studie ber Eichtheorien und Strukturenrealismus, Paderborn: mentis Verlag.

Malament, D., 1996, In defense of dogma: Why there cannot be a relativistic


quantum mechanics of (localizable) particles, in Clifton 1996, pp. 1-10.

Martin, C. A., 2002, Gauge principles, gauge arguments and the logic of nature,
Philosophy of Science, 69/3: 221-234.

Newton, T. D. and Wigner, E. P., 1949, Localized states for elementary particles,
Reviews of Modern Physics, 21/3: 400-406.

Peskin, M. E. and Schroeder, D. V., 1995, Introduction to Quantum Field Theory,


Cambridge, MA: Perseus Books.

Polchinski, J., 2000, String Theory, 2 volumes, Cambridge: Cambridge University


Press.

Redei, M. and Summers, S. J., 2002, Local primitive causality and the common
cause principle in quantum field theory, Foundations of Physics, 32: 335-355.

Redhead, M. L. G., 1975, Symmetry in intertheory relations, Synthese, 32: 77112.

Redhead, M. L. G., 1980, Some philosophical aspects of particle physics, Studies


in History and Philosophy of Science, 11: 279-304.

Redhead, M. L. G., 1983, Quantum field theory for philosophers, in Asquith, P. D.


and Nickles, T., eds., 1983, Proceedings of the Biennial Meeting of the Philosophy
of Science Association: PSA 1982, East Lansing, MI: Philosophy of Science
Association, vol. 2, pp. 57-99.

Redhead, M. L. G., 1988, A philosopher looks at quantum field theory, in Brown


& Harr 1988, pp. 9-23.

Redhead, M. L. G., 1995a, More ado about nothing, Foundations of Physics, 25:
123-137.

Redhead, M. L. G., 1995b, The vacuum in relativistic quantum field theory, in


Hull et al. 1994 (vol. 2), pp. 88-89.

Redhead, M. L. G., 2002a, The interpretation of gauge symmetry, in Kuhlmann et


al. 2002, pp. 281-301.

Redhead, M. L. G., 2002b, Quantum field theory and the philosopher, in Cao
1999, pp 34-40.

Redhead, M. L. G. and Teller, P., 1991, Particles, particle labels, and quanta: the
toll of unacknowledged metaphysics, Foundations of Physics, 21: 43-62.

Redhead, M. L. G. and Wagner, F., 1998, Unified treatment of EPR and Bell
arguments in Algebraic Quantum Field Theory, Foundations of Physics Letters, 11:
111-125.

Reeh, H. and Schlieder, S., 1961, Bemerkungen zur Unitrquivalenz von


Lorentzinvarianten Feldern, Nuovo Cimento, 22: 1051-1068.

Roberts, J. E., 1990, Lectures on algebraic quantum field theory, in Kastler 1990,
pp. 1-112.

Ruetsche, L., 2003, A matter of degree: Putting unitary equivalence to work,


Philosophy of Science, 70/5: 1329-1342.

Ryder, L. H., 1996, Quantum Field Theory, 2nd edition, Cambridge: Cambridge
University Press.

Saunders, S., 1988, The algebraic approach to quantum field theory, in Brown and
Harr 1988, pp. 149-183.

Saunders, S., 1995, A dissolution of the problem of locality, in Hull, M. F. D.,


Forbes, M., and Burian, R. M., eds., 1995, Proceedings of the Biennial Meeting of
the Philosophy of Science Association: PSA 1994, East Lansing, MI: Philosophy of
Science Association, vol. 2, pp. 88-98.

Saunders, S. and Brown, H. R., eds., 1991, The Philosophy of Vacuum, Oxford:
Clarendon Press.

Scharf, G., 1995, Finite Quantum Electrodynamics: the Causal Approach, 2nd
edition, Berlin-Heidelberg-New York: Springer.

Schrdinger, E., 1954, Nature and the Greeks, Cambridge: Cambridge University
Press, Cambridge. (Shearman Lectures, delivered at University College, London,
May 1948.)

Schweber, S. S., 1994, QED and the Men Who Made It, Princeton: Princeton
University Press.

Segal, I. E., 1947a, Irreducible representations of operator algebras, Bulletin of


the American Mathematical Society, 53: 73-88.

Segal, I. E., 1947b, Postulates for general quantum mechanics, Annals of


Mathematics, 48/4: 930-948.

Seibt, J., 2002, The matrix of ontological thinking: Heuristic preliminaries for an
ontology of QFT, in Kuhlmann et al. 2002, pp. 53-97.

Streater, R. F. and Wightman, A. S., 1964, PCT, Spin and Statistics, and all that,
New York: Benjamin.

Teller, P., 1983, Quantum physics, the identity of indiscernibles and some
unanswered questions, Philosophy of Science, 50: 309-319.

Teller, P., 1990, What the quantum field is not, Philosophical Topics, 18: 175-186.

Teller, P., 1995, An Interpretive Introduction to Quantum Field Theory, Princeton:


Princeton University Press.

Unruh, W. G., 1976, Notes on black hole evaporation, Physical Review D, 14:
870-92.

Unruh, W. G. and Wald, R. M., 1984, What happens when an accelerating observer
detects a Rindler particle? Physical Review D, 29: 1047-1056.

Weinberg, S., 1995, The Quantum Theory of Fields Foundations, Volume 1.


Cambridge: Cambridge University Press.

Weingard, R., 2001, A philosopher looks at string theory, in Callender & Huggett
2001, pp. 138-151.

Weyl, H., 1952, Symmetry, Princeton: Princeton University Press.

Wigner, E. P., 1939, On unitary representations of the inhomoneneous Lorentz


group, Annals of Mathematics, 40: 149-204.

Vous aimerez peut-être aussi