Methodological and metaphysical guidelines
The dissipative approach to QFT is built following a set of methodological and metaphysical guidelines which we consider sound requirements to construct a consistent theoretical framework from both a formal and ontological perspective. As already stated, the principal aim of such a proposal is to formulate an alternative, effective quantum theory of fields capable of solving the major problems affecting its standard formulation starting from clear foundations. Hence, let us present the criteria we assume and employ in this essayFootnote 3:
-
We consider mathematical consistency and rigor essential requisites of any robust physical theory. On the one hand, mathematical consistency is a virtue useful in order to propose an empirically adequate physical theory, i.e. a theoretical framework able to reproduce the statistics of observed experimental findings avoiding computational deficiencies. On the other hand, it ensures that theories do not lead to contradicting results. Referring to this, it will be shown in this section that the mathematical structure of DQFT is consistent and rigorous being based on a set of clear notions and dynamical equations based on nonequilibrium thermodynamics, which guarantee that the formal machinery employed is not affected by the issues characterizing the standard formulation of QFT; it may be said, in fact, that the purpose of thermodynamics is to characterize and formulate robust equations that make mathematical sense, that is, for which the existence and uniqueness of solutions can be proven.
-
A background finite Minkowski space-time is assumed, since physical phenomena treated by QFT are usually represented as events taking place in relativistic space-time settings. This assumption entails several mathematical consequences; for instance, one can retain the inhomogeneous Lorentz transformations, and therefore, Wigner’s classifications of particles in terms of mass and spin, considering them as inherent, fundamental properties of elementary particles (cf. Section 3). It should be underlined, furthermore, that DQFT is not concerned with the inherent nature of space-time: in what follows we remain agnostic towards its ontology, whose treatment will require a deeper theory with respect to QFT. As a consequence, we consider the latter explicitly an effective theory valid only in a specific range of energy-length scales. Thus, it is possible to consider the choice of such background spacetime as a simplifying assumption.Footnote 4
-
Moving to the metaphysical principles, we aim at providing a realistic picture of the objects and processes taking place at QFT scale. More precisely, we will define precisely what are the theoretical entities representing real objects in the world and their dynamical behaviour in physical space, avoiding the metaphysical indeterminacy affecting standard QFT. So far, it is sufficient to state that such an ontology ensures that the dissipative QFT will have a precise commitment towards the existence of a well-defined set of objects, whose reality is independent of any observation and measurement. Hence, we claim, contrary to a widespread view in the philosophical literature, that it is possible to restore a realistic picture of physical processes taking place in space also in the context of QFT.
-
In order to tame the conceptual and technical problems arising from the different types of infinities occurring in standard QFT by construction, we assume that according to the dissipative approach such infinities are taken to be only potential, not actual. Therefore, in the present theory we keep the number of quantum particles always finite and countable, so that physical states can be described via a Fock space representation. As we will see in the remainder of the paper, this fact will help us to circumvent the metaphysical implications of the infinitely many inequivalent representations of the CCR. In addition, we introduce restrictions preventing the appearance of divergencies: on the one hand, we consider large but finite volumes of space, i.e. a finite universe; this fact consequently imposes a characteristic length scale and an infrared regularization. On the other hand, we take into account a dissipative mechanism which is necessary to have ultraviolet regularization. These assumptions are crucial in order to obtain an empirically adequate and well-behaved theory.Footnote 5
-
Furthermore, we assume that stochasticity naturally emerges in QFT, since there are several random events in such framework that can neither be mechanically controlled, nor precisely known, as for instance vacuum fluctuations causing electron-positron pairs which spontaneously appear and disappear. The existence of such events and our inability to known and control them should be considered as a natural source of irreversible behaviour. Hence, this fact motivates to propose an inherently stochastic dynamics for DQFT. Moreover, since the latter is based on arguments taken from nonequilibrium thermodynamics, we must underline that in such context random fluctuations are accompanied with dissipation, irreversibility and decoherence. Thus, it seems natural for a QFT based on nonequilibrium thermodynamics to implement a stochastic dynamics, which also is motivated by experimental evidence and the phenomenology of the quantum theory of fields.
To conclude this preliminary illustration of the guiding principles of DQFT, it is worth stressing again that it is explicitly an effective theory, having a definite characteristic scale lying between 10− 20m, which is the scale of super-colliders, and 10− 35m, which is the Planck scale. Consequently, we model the physical influences due to objects and processes at higher energy scales through a heat bath. Referring to this, we heavily rely on modern renormalization methods—sharing the arguments in favor of them stated in Wallace (2006)—which are essential tools introduced to tame the already mentioned issues concerning infinities and divergencies and keeping the present theory formally well-defined.Footnote 6
The mathematical arena: Fock space representation, creation and annihilation operators and fields
Fock space \(\mathcal {F}\), a particular kind of complex vector space with inner product, is the mathematical arena in which the dissipative approach to QFT takes place. In this state space a system of independent quantum objects—whose number can vary in time—is represented by the following expression:
$$ \begin{array}{@{}rcl@{}} |n_{i}\rangle=|n_{1}, n_{2}, \dots\rangle. \end{array} $$
(1)
The states of the form written above represent an orthonormal basis vector in \(\mathcal {F}\), where the ket on the r.h.s. indicates a vector in which n1 represents the number of objects in the state 1, n2 represents the number of objects in state 2 and so on. It is worth noting that Eq. 1 only counts the number of quantum objects present in a certain state, it does not assign any label to them, i.e. these objects do not possess an inherent “thisness” or “haecceity” using Teller’s words; alternatively stated, particles of the same species in the same state are absolutely indistinguishable.
In addition, for bosons each occupation number ni is a non-negative integer; for fermions it must be 0 or 1 in virtue of Pauli’s exclusion principle, which prevents the possibility for different fermions to occupy the same state. The vacuum state |0〉 denotes a state in which all occupation numbers vanish, or more precisely, a state in which no object is present. For the sake of simplicity, we will speak about bosons and fermions, however, we will properly introduce the fundamental objects of this theory, i.e. its ontology, later on.
Exactly as in standard QFT, creation and annihilation operators for bosons and fermions are defined in \(\mathcal {F}\).Footnote 7 In the first case, the the creation operator \(a^{\dagger }_{i}\) increases the number of bosons in the state i by one,
$$ \begin{array}{@{}rcl@{}} a^{\dagger}_{i}|n_{1}, n_{2}, \dots\rangle =\sqrt{n_{i}+1}|n_{1}, n_{2}, \dots, n_{i} +1\rangle, \end{array} $$
(2)
conversely, the annihilation operator ai decreases it by one:
$$ \begin{array}{@{}rcl@{}} a_{i}|n_{1}, n_{2}, \dots\rangle =\left\{\begin{array}{ll} \sqrt{n_{i}}|n_{1}, n_{2}, \dots, n_{i} -1\rangle, \ \text{for}\ n_{i}>0, \\ \\ 0, \ \text{for}\ n_{i}=0. \end{array}\right. \end{array} $$
(3)
These operators obey the following commutation relations:
$$ \begin{array}{@{}rcl@{}} [a_{v}, a^{\dagger}_{v'}]=\delta_{vv'} \end{array} $$
(4)
and
$$ \begin{array}{@{}rcl@{}} [a_{v}, a_{v^{\prime}}]=[a^{\dagger}_{v}, a^{\dagger}_{v'}]=0 \end{array} $$
(5)
where [A,B] = AB − BA is the commutator of two generic operators A,B in \(\mathcal {F}\). Here we will not consider the definition of such operators for fermions, since these are not strictly relevant for the purposes of the present essay.Footnote 8
It is well-known in the mathematical and physical literature that a Fock space can be rigorously constructed starting from a N-particle Hilbert space.Footnote 9 The main reason for not following this route to define \(\mathcal {F}\) in DQFT is metaphysical in essence, since with the symmetrization and the anti-symmetrization of the tensor products the particles do obtain a label, which is more than what we actually need to define our ontology, as stressed a few lines above. The Hilbert space formalism, thus, “says too much” about the inherent nature of quantum particles. On the contrary, the way to define the Fock space presented above eliminates particles’ labels, providing us information concerning uniquely the particles’ numbers.
In this theory, if we consider a configuration of “particles” composed by several species, each of them is represented by an appropriate Fock space; the total configuration will be consequently represented by a single product space, obtained combining each specific Fock space of the individual particles’ species at hand. Notably, this latter space will have a unique vacuum, corresponding to the state in which there are no particle of any species. The corresponding Fock states describe an ensemble of independent particles of different kinds; however, not all the possible combinations among states are physically meaningful, as for instance superpositions of boson and fermion states, or states with different electric charges. Such limitations are known as superselection rules.
Furthermore, it is worth stressing that creation and annihilation operators do not carry ontological weight per se: they are useful formal tools needed (i) for the definition of a variable number of particles in \(\mathcal {F}\), and (ii) to represent physical events of particle creation and destruction occurring in spacetime. Nonetheless, what is ontologically primary in DQFT are quantum particles which can be randomly created and annihilated. These operators, then, play an important functional role, i.e. to represent mathematically such physical events. As already mentioned, in DQFT the problem of the infinitely many representations of the canonical commutations relations vanishes by construction, since we have a unique representation of such relations keeping finite the number of the degrees of freedom.
Another step to the definition of DQFT is to select momentum eigenstates to represent single-particle states; as a consequence, momentum space is the fundamental representation of physical systems in this framework. More precisely, we will consider a discrete set of momentum states—this is coherent with the idea to have a Fock space with a countable dimension at any time—on a discrete d-dimensional lattice:
$$ \begin{array}{@{}rcl@{}} K^{d}=\{ \boldsymbol{{k}} = (z_{1}, \dots, z_{d})K_{L} | z_{j} \ \text{integer with}\ |z_{j}|\leq N_{L}\ \text{for all}\ j=1, \dots, d\}, \end{array} $$
(6)
where d is the finite dimension of our space, KL is a lattice constant in momentum space, which is small by assumption, and the large integer NL limits the magnitude of each component of k to NLKL. In the above equation KL,NL are truncation parameters which keep the space finite; in addition, the finite number of elements in Kd correspond to the label i of the general construction of Fock spaces.Footnote 10,Footnote 11
A further consideration about the ontology of DQFT concerns the role of fields, which do not represent physical entities in spacetime according to the present theory, being only mathematical tools introduced for heuristic reasons without a direct physical meaning. More precisely, they are useful quantities to compute collisions and relevant quantities of interest, but they do not represent physical objects in spacetime in addition to the particles.Footnote 12 Nonetheless, it is formally useful for the exposition of this theory to introduce the following field (self-adjoint) operator:
$$ \begin{array}{@{}rcl@{}} \varphi_{\boldsymbol{{x}}}=\frac{1}{\sqrt{V}}\sum\limits_{\boldsymbol{{k}} \in K^{d}}\frac{1}{\sqrt{2\omega_{k}}}\left( a^{\dagger}_{\boldsymbol{{k}}}+a_{-\boldsymbol{{k}}}\right)e^{-i\boldsymbol{{k}}\cdot \boldsymbol{{x}}} \end{array} $$
(7)
where V is the volume of our finite space, \(\omega _{k}=\sqrt {m^{2}+k^{2}}\) is a weight factor which is the relativistic energy-momentum relation for a particle of mass m.Footnote 13 Interestingly, the physical significance of the factor \(1/\sqrt {2\omega _{k}}\) becomes clear in actual computations of correlation functions of relevant quantities of interest (see Section 2.4). However, it should be stressed that such factors do not permit to interpret the above (7) as a passage from momentum eigenstates to position eigenstates. This fact entails consequences, i.e. an indispensable difficulty for DQFT to know where particles are located in space (this problem issue is tamed in the non-relativistic case, where particles have low velocity compared to c, ωk is substituted with a constant m, so that Eq. 7 can be interpreted as position eigenstates. We will come back to this issue in Section 3).
So far we have been silent about what is the ontology of this theory, i.e. its fundamental entities, however, on the one hand we have designed the Fock space in a way able to account for individual, discrete, countable objects, whose number can vary in time, on the other hand, we stated that neither creation and annihilation operators, nor fields have ontological status, these are only powerful formal tools appearing in the formal machinery of DQFT.
The dynamics of DQFT
Having defined the state space of our theory, the creation and annihilation operators and fields, let us now discuss two possible ways to describe the dynamics of the dissipative approach to QFT, the first relying on the Schrödinger picture, the second on unravelings of a quantum master equation which will be introduced below. Let us start with the former.
In the first place, it is important to underline that the complete dynamics of DQFT represented in the Schrödinger picture is composed of two contributions, the reversible and irreversible ones. Considering the reversible contribution, the dynamical evolution of a time-dependent state vector |ψt〉 in Hilbert space which is governed by the well-known unitary Schrödinger Equation (SE):
$$ \begin{array}{@{}rcl@{}} \frac{d}{dt}|\psi_{t}\rangle=-iH|\psi_{t}\rangle \end{array} $$
(8)
where H is the Hamiltonian operator, whose spectrum is assumed to be bounded from below in the context of DQFT.
In order to describe the full structure of the Hamiltonian, let us take into account the interaction among four colliding particles in a d dimensional space, using the φ4 theoryFootnote 14:
$$ \begin{array}{@{}rcl@{}} &&H=\sum\limits_{\boldsymbol{{k}} \in K^{d}}\omega_{k} a^{\dagger}_{\boldsymbol{{k}}}a_{\boldsymbol{{k}}}+\frac{\lambda}{96} \frac{1}{V} \sum\limits_{\boldsymbol{{k}}_{1}, \boldsymbol{{k}}_{2}, \boldsymbol{{k}}_{3}, \boldsymbol{{k}}_{4}\in K^{d}} \frac{\delta_{\boldsymbol{{k}}_{1}+\boldsymbol{{k}}_{2}+\boldsymbol{{k}}_{3}+\boldsymbol{{k}}_{4}, \textbf{0}}}{\sqrt{\omega_{k_{1}} \omega_{k_{2}} \omega_{k_{3}} \omega_{k_{4}}}}\\ &&\quad \left( a_{\boldsymbol{{k}}_{1}}a_{-\boldsymbol{{k}}_{2}}a_{-\boldsymbol{{k}}_{3}}a_{-\boldsymbol{{k}}_{4}}+4a^{\dagger}_{\boldsymbol{{k}}_{1}}a_{-\boldsymbol{{k}}_{2}}a_{-\boldsymbol{{k}}_{3}}a_{-\boldsymbol{{k}}_{4}}+6a^{\dagger}_{\boldsymbol{{k}}_{1}}a^{\dagger}_{\boldsymbol{{k}}_{2}}a_{-\boldsymbol{{k}}_{3}}a_{-\boldsymbol{{k}}_{4}}\right. \\ &&\quad \left. +4a^{\dagger}_{\boldsymbol{{k}}_{1}}a^{\dagger}_{\boldsymbol{{k}}_{2}}a^{\dagger}_{\boldsymbol{{k}}_{3}}a_{-\boldsymbol{{k}}_{4}}+a^{\dagger}_{\boldsymbol{{k}}_{1}}a^{\dagger}_{\boldsymbol{{k}}_{2}}a^{\dagger}_{\boldsymbol{{k}}_{3}}a^{\dagger}_{\boldsymbol{{k}}_{4}}\right) \\ &&\quad +\frac{\lambda^{\prime}}{4}\sum\limits_{\boldsymbol{{k}} \in K^{d}}\frac{1}{\omega_{k}}\left( a_{\boldsymbol{{k}}} a_{-\boldsymbol{{k}}}+2a^{\dagger}_{\boldsymbol{{k}}} a_{\boldsymbol{{k}}} +a^{\dagger}_{\boldsymbol{{k}}} a^{\dagger}_{-\boldsymbol{{k}}}\right)+\lambda^{\prime\prime}V. \end{array} $$
(9)
In Eq. 9δ is Kronecker’s δ and \(\lambda , \lambda ^{\prime }, \lambda ^{\prime \prime }\) are three free interaction parameters determining the strength of the quartic interaction. More precisely, λ should be regarded as the fundamental interaction parameter, whereas \(\lambda ^{\prime }, \lambda ^{\prime \prime }\) should be considered correction parameters, the former referring to the additional contribution to the square of the mass, and the latter referring to a constant background energy per unit of volume.Footnote 15 It is important to underline that in Eq. 9 momentum is conserved in collisions; this fact in turn implies the locality of such interactions, which nonetheless does not imply that DQFT has the resources needed to strictly localize particles in space-time, as mentioned above.
Interestingly, Eq. 9 can change the number of the individual particles by an even amount: 0, meaning that it leaves the number unchanged, ± 2 and ± 4 which means that the particle number can by increased or decreased by 2 and 4 respectively.
The second dynamical contribution of DQFT is inherently stochastic and here is where thermodynamical arguments—more precisely the dissipation mechanism—come properly into the scene. In what follows we describe our physical systems in terms of density matrices ρt, which can represent a number of different physical states occurring with a certain probability. In this context, density matrices are useful formal tools which enables us to treat ensembles formed by identical and indistinguishable particles, since their statistical properties are completely described in terms of ρt. It is worth noting that in DQFT density matrices do not represent physical objects in spacetime over and above the ensembles of particles which they describe; in this context they have only a functional role for the particle dynamics. Thus, they should not be compared e.g. to the ψ −function in Bohm’s original pilot-wave theory (cf. Bohm 1952), where the wave function is defined in three-dimensional space, and it is a proper physical field which guides the particles’ motion.
Following the usual treatment of dissipative quantum systems, we introduce an inherently stochastic Quantum Master Equation (QME) for the density matrixFootnote 16, which takes the following form for the above mentioned φ4 theory:
$$ \begin{array}{@{}rcl@{}} \frac{d\rho_{t}}{dt}&=&-i[H, \rho_{t}]-\sum\limits_{\boldsymbol{{k}} \in K^{d}}\upbeta\gamma_{k}{{\int}^{1}_{0}} e^{-u\upbeta\omega_{k}} \left( [a_{\boldsymbol{{k}}}, \rho_{t}^{1-u}[a^{\dagger}_{\boldsymbol{{k}}}, \mu_{t}]{\rho^{u}_{t}}]\right.\\&&\left.+[a^{\dagger}_{\boldsymbol{{k}}}, {\rho^{u}_{t}}[a_{\boldsymbol{{k}}}, \mu_{t}] \rho_{t}^{1-u}]\right)du \end{array} $$
(12)
where \(a_{\boldsymbol {{k}}}, a^{\dagger }_{\boldsymbol {{k}}}\) are the coupling operators which model the interaction between our open system and its environment, which in the present quantum field theory is given by a heat bath of a given temperature T, representing the eliminated small-scales/high-energy degrees of freedom which directly influence and interact with our lower energy quantum particles.Footnote 17 Furthermore, the term \(e^{-u\upbeta \omega _{k}}\) “produces the proper relative weights for transitions involving the creation and annihilation of free particles” (Öttinger (2017), p. 65), β = 1/kBT represents the inverse temperature, γk denotes the decay rate, i.e. the damping coefficient describing the strength of the dissipation, which is negligible for small k and increases rapidly for large k.Footnote 18 Here the concrete form of the decay rateFootnote 19 is γk = γ0 + γk4: the factor k2 refers to the Laplace operator which causes diffusive smoothing in real space, however the presence of double commutators in Eq. 12 suggests the k4 power; the parameter γ0 is added since the state with k = 0 can be subject to dissipation. It is worth noting that the damping of the latter state k = 0 must be infinitesimally small to be consistent with the results of low energy QFT. Alternatively stated, as the parameter γ provides a UV cutoff (note that γ1/3 defines a length scale), this parameter should be sufficiently small to be in the physically inaccessible range. In the spirit of the renormalization procedure and motivated by the standard procedure in QFT, the precise value of γ does not matter here.
In order to characterize the features of the above equation, it is worth noting that the first term on the r.h.s. of Eq. 12 correspond to the dynamics given by Eq. 8. The second term appearing characterizes instead the irreversible process, which is given by the commutators involving the energy operator \(\mu _{t}=H+k_{B} T \ln \rho _{t}\)—the generator of the irreversible dynamics—and the real, non-negative rate factor \(e^{-u\upbeta \omega _{k}}\). As Breuer and Petruccione (2002) p. 129 underline, the irreversible dynamics is related to entropy production—which is non-negative and vanishes at equilibrium—in the precise sense that the latter is the “amount of entropy produced per unit of time as a result of irreversible processes”. In addition, as claimed in Öttinger (2017), pp. 62-63:
[t]he multiplicative splitting of ρt into the powers \({\rho _{t}^{u}}\) and \(\rho _{t}^{1-u}\), with an integration over u, is introduced to guarantee an appropriate interplay with entropy and hence a proper steady state or equilibrium solution. The structure of the irreversible term is determined by general arguments of nonequilibrium thermodynamics or, more formally, by a modular dynamical semigroup.
Consequently, the above mentioned QME implies the convergence to the equilibrium density matrix. Considering this concrete form of QME for a density matrix, it is important to say that temperature is naturally associated with the heat bath that consists of the unresolved, local degrees of freedom of an effective field theory. The ubiquitous loops of collisions involving high-momentum particles and occurring within short periods of time, which are the origin of divergences in QFT, become unresolvable due to the presence of dissipation, which quickly eliminates high-momentum particles and thus provides regularization. A detailed discussion of the resulting unresolvable clouds of individual particles that can be effectively seen through particle detectors can be found in Section 3. Referring to this, it is worth noting that in this theory the heat bath constituting the environment is assumed to be in thermal equilibrium, since only the slow large scale degrees of freedom can actually feel the nonequilibrium effects. This fact, in turn, follows from the fundamental assumption of a separation of time scales in nonequilibrium thermodynamics, which entails that the eliminated fine grained degrees of freedom are in equilibrium (cf. Öttinger 2009 for technical details).
Finally, it is important to stress that the QME is one of the most efficient ways to represent the interaction between a quantum system and a heat bath, since an exact treatment of the high energy degrees of freedom would require the solution of a too complex system of coupled equations of motion. In the second place, the evolution of the heat bath’s degrees of freedom can be neither known, nor mechanically controlled, thus, one has to simplify the description of such physical situation taking into consideration a restricted set of relevant quantities accounting for this influence. Referring to this, the short-time correlations with the heat bath allow one to neglect memory effects on the dynamics and to define a stochastic Markov process on the state space of the system, given that such times are much smaller than the characteristic time scale of the system’s evolution, as clearly stated by Breuer and Petruccione (2002), pp. 115-122.
To conclude this section let us briefly underline the crucial role of renormalization group methods in the context of DQFT. As repeatedly stressed, the fast degrees of freedom are eliminated from our theory, these form the environment with which individual particles interact. This scaling is dependent on the friction parameter present in the QME (therefore, also the notion of interacting particle depends on such scaling): increasing the length scale is equivalent to increasing the parameter γk in Eq. 12, with the consequence of increasing in the entropy production rate.
Finally, in DQFT the entities that can be subject to detection and observations are clouds of particles emerging from the collisions and interactions of more fundamental and faster degrees of freedom which are instead inaccessible; it should be noted that “the dissipative coupling to the bath is very weak, except at short length scales. In other words, the dissipative coupling erases the short-scale features very rapidly, whereas it leaves large-scale features basically unaffected” (Öttinger 2017, p. 29). Furthermore, since there is no a clear cut decoupling among the various high and low-energy processes we assume self-similarity, meaning that although the faster degrees of freedom are eliminated and not directly treated by DQFT, we stipulate that they behave in a similar way with respect to the slower degrees of freedom. Rigorous arguments to justify this claim are contained in Öttinger (2009).Footnote 20
Quantities of interest
The quantities of interest one may want to compute rest on subjective decisions; however, in this subsection we will provide the most general class of multi-time correlation functions associated to measurable quantities which connect the general abstract formalism of the theory presented so far with experimental evidence.Footnote 21
Firstly, it is worth noting that here we deal uniquely with statistical quantities, hence, it is natural to work with density matrices, as previously anticipated. The formal expression of a multi-time correlation function is given as follows:
$$ \begin{array}{@{}rcl@{}} \text{tr}\{ \mathcal{N}_{n}A_{n}\mathcal{E}_{t_{n}-t_{n-1}} ({\dots} \mathcal{N}_{2}A_{2}\mathcal{E}_{t_{2}-t_{1}} (\mathcal{N}_{1}A_{1}\mathcal{E}_{t_{1}-t_{0}}(\rho_{0})A^{\dagger}_{1})A^{\dagger}_{2}\dots)A^{\dagger}_{n}\}. \end{array} $$
(13)
This formula must be read from the inside to the outside: we start from a density matrix ρ0 at time t0, the evolution super-operator \(\mathcal {E}\) is obtained by solving the QME over a definite interval of time t, Aj represent linear operators associated with times tj with \(t_{0} < t_{1} < \dots , < t_{n}\), finally the normalization factors \(\mathcal {N}\) guarantee that after every step the evolution continues with the density matrix. Importantly, the experimental outcomes of a time series of different measurements is contained in the normalization factors.Footnote 22
Unraveling of the quantum master equation
Another possibility to represent the dynamics of the dissipative approach to quantum field theory is based on the notion of unraveling of the quantum master equation.Footnote 23 Specifically, instead of formulating the dynamics of DQFT using quantum master equations for density matrices, it is possible to represent it in terms of a stochastic process in the system’s state space. Thus, the fundamental idea at play is to re-write the dynamics of the presented theory obtaining a time-dependent density matrix ρt solving a QME as second moment or expectation ρt = E(|ψt〉〈ψt|), where |ψt〉 is a stochastic process in the relevant Fock space of the open system at hand consisting of periods of continuous Schrödinger-type evolution interrupted by random quantum jumps. We underline that the unravelings are not unique, and here we explain only the most basic ideas behind unravelings for the simplest case of a non-interacting theory (for more general developments see Öttinger (2017) and references therein). We first fill in some details on the one-process unravelings considered above and then motivate and develop the idea of two-process unravelings. Here and for all generalizations, we consider unravelings in which the state vector at any time t is a complex multiple of one of the base vectors of \(\mathcal {F}\), where interactions need to be expressed as jumps. This restriction, which can be regarded as a superselection rule, has important consequences: at any time t, the system has a well-defined particle content and superpositions do not play any role in our unravelings (cf. also Pashby and Öttinger 2021). The practical advantages of this restriction for numerical simulations is discussed in Section 3.3.
One-process unravelings:
The main idea can be explained more conveniently by considering the zero-temperature master equation for the non-interacting theory:
$$ \begin{array}{@{}rcl@{}} {\frac{d}{dt}|\psi_{t}\rangle=-iH_{\text{free}}|\psi_{t}\rangle- \sum\limits_{k\in K^{d}}\gamma_{k}(1-|\psi_{t}\rangle\langle\psi_{t}|)a^{\dagger}_{k}a_{k} |\psi_{t}\rangle}. \end{array} $$
(14)
The above equation contains a dissipative term, and the continuous Schrödinger-type evolution (14) is interrupted by jumps of the form:
$$ \begin{array}{@{}rcl@{}} {|\psi_{t}\rangle\rightarrow\frac{a_{k}|\psi_{t}\rangle}{\Vert a_{k}|\psi_{t}\rangle\Vert}} \end{array} $$
(15)
occurring with rate \(2\langle \psi _{t}|a^{\dagger }_{k}a_{k}|\psi _{t}\rangle \). For a clear explanation let us consider Fig. 1:
Here we consider a decay of a three-particle state, particles are then removed until one gets the vacuum |0〉. Interestingly, at any time t one can calculate the probability to find any state that can be generated by removing one of the particles from the initial Fock space. Looking at Fig. 1 we start at the top vertex of the hexagon with three particles, by removing one of them we can obtain three different states represented in the second line; by reiterating the process we obtain three different one-particle states and, eventually, one can reach the vacuum annihilating the last particle. Interestingly, it should be underlined that at any finite time, it is possible to compute the probability to find any state that can be obtained by the removal of a number of particles from the initial Fock state. Since in QFT we have to do with real events of creation and annihilation of quantum objects, we interpret these unraveling as real physical processes in space.
Two-process unravelings:
In the one-process unraveling any change in |ψt〉 affects and modifies in the same way both the bra and ket component of |ψt〉〈ψt|. However, for interacting theories or when we are interested in more general correlation functions than those listed in Eq. 13, we need to decouple the bra and ket components. In this case, one should use the two-process unraveling, which are based on the following representation of the density matrix of our system ρt = E(|ϕt〉〈ψt|), where |ϕt〉 and |ψt〉 are two random trajectories in Fock space, i.e. two different lists of individual particles, with potentially different jumps.Footnote 24 For the example of the free theory at zero temperature, the two-process unraveling introduces two simultaneous jumps:
$$ \begin{array}{@{}rcl@{}} {|\phi_{t}\rangle\rightarrow\frac{a_{k}|\phi_{t}\rangle \Vert |\phi_{t}\rangle\Vert}{\Vert a_{k}|\phi_{t}\rangle\Vert}} \\ \\ {|\psi_{t}\rangle\rightarrow\frac{a_{k}|\psi_{t}\rangle \Vert |\psi_{t}\rangle\Vert}{\Vert a_{k}|\psi_{t}\rangle\Vert}} \end{array} $$
with rate 2ik(|ϕt〉,|ψt〉)γk, where
$$i_{k}(|\phi_{t}\rangle, |\psi_{t}\rangle)=\frac{\Vert a_{k}|\phi_{t}\rangle\Vert \Vert a_{k}|\psi_{t}\rangle \Vert}{\Vert |\phi_{t}\rangle\Vert \Vert |\psi_{t}\rangle\Vert},$$
and two unitary evolution equations:
$$ \begin{array}{@{}rcl@{}} {\frac{d}{dt}|\phi_{t}\rangle=-iH_{\text{free}}|\phi_{t}\rangle-\sum\limits_{k}\gamma_{k}\left[a^{\dagger}_{k}a_{k}-i_{k}(|\phi_{t}\rangle, |\psi_{t}\rangle)\right]|\phi_{t}\rangle} \\ {\frac{d}{dt}|\psi_{t}\rangle=-iH_{\text{free}}|\psi_{t}\rangle-\sum\limits_{k}\gamma_{k}\left[a^{\dagger}_{k}a_{k}-i_{k}(|\phi_{t}\rangle, |\psi_{t}\rangle)\right]|\psi_{t}\rangle.} \end{array} $$
In this case jumps can take place only if both the two vectors |ϕt〉 and |ψt〉 contain a particle with the same momentum k.
The two-process unraveling is helpful in the calculation of multi-time correlation functions of a more general type than listed in Eq. 13, as illustrated in Fig. 2 below:
Considering an initial ensemble of states |ϕ0〉 and |ψ0〉 representing the system’s density matrix ρ0, they evolve from time t0 to t1 according to the two-process unraveling. The operators Ai,Bj are then introduced via the jumps of |ϕj〉 and |ψj〉, at times tj, between these jumps the states and their trajectories in Fock space evolve according to the two-process unraveling. At the final time one gets the final states |ϕf〉 and |ψf〉, which allow us to evaluate the multi-time correlation function as follows:
$$ \begin{array}{@{}rcl@{}} {tr\bigg\{A_{n}\mathcal{E}_{t_{n}-t_{n-1}} \bigg({\dots} A_{2}\mathcal{E}_{t_{2}-t_{1}}\left( A_{1}\mathcal{E}_{t_{1}-t_{0}}(\rho_{0})B^{\dagger}_{1}\right) B^{\dagger}_{2} {\dots} \bigg) B^{\dagger}_{n} \bigg\}=E[\langle\psi_{f}|\phi_{f}\rangle].} \end{array} $$
This discussion can be extended incorporating various forms of the unravelings and concrete examples applied to the φ4 theory; a fully detailed picture of these processes are given in Öttinger (2017), Section 1.2.8. However, a such technical discussion is beyond the introductory scope of the present essay.
Dissipation mechanism: more than another UV regularization scheme
To conclude our introduction to DQFT let us play the role of devil’s advocate. Considering the dissipation mechanism of DQFT, one might regard it just as another ultraviolet regularization scheme such as, for example, lattice regularization, momentum cut-off, dimensional regularization, or Pauli–Villars regularization. Thus, one would conclude, DQFT would simply retrace the road of standard QFT in order to avoid unwelcome results as those summarized in Section 1. Contrary to this potential objection, in this subsection we are going to explain why dissipation should not be considered another merely formal regularization scheme. In what follows, then, we summarize a number of arguments showing that there is much more to dissipation; some of them will be elaborated in more detail in the following section.
-
1.
It is worth noting that regularization is deeply related to renormalization, that is, to the elimination of degrees of freedom. Moreover, whenever degrees of freedom are eliminated one should expect entropy and dissipation to play a role, i.e. one should expect to enter the realm of irreversible thermodynamics. The occurrence of irreversibility should be considered natural since the infamous divergences in QFT arise from spontaneous particle creation and annihilation, processes that are far beyond our mechanistic control being too fast and too local. This is the motivation which led us to assume that stochasticity naturally emerges in QFT in Section 2.2.
-
2.
Despite the stochastic character of fundamental interactions, they are described via Hamiltonian dynamics (without much critical questioning) which has a pure reversible structure. The equations of irreversible thermodynamics possess a mathematical structure that generalizes Hamiltonian dynamics. Nonequilibrium thermodynamics, indeed, not only provides robust evolution equations, but also important additional features, such as a fluctuation-dissipation relation characterizing the thermal fluctuations accompanying a dissipation mechanism at nonzero temperature (see Öttinger et al. 2021 and references therein). Hence, the dissipation mechanism seems to be more appropriate to represent fundamental interactions.
-
3.
In the effective field theories of electro-weak and strong interactions, the strength of the dissipation in DQFT is a variable parameter, very much like a lattice spacing or a momentum cutoff, requiring a renormalization treatment. However, unlike these merely computational tools, dynamic dissipative smearing provides a more appealing option for a physical theory at some fundamental scale, namely the Planck scale. Dissipative smearing may be interpreted as the origin of the limit of resolution at the Planck scale and must hence associated with gravity. An alternative theory of gravity that could be treated by means of DQFT has been proposed and elaborated in Öttinger (2020a, 2020b). This higher derivative theory of gravity effectively selects a small subset of solutions from the Yang-Mills theory based on the Lorentz group via constraints. As a result, all fundamental interactions would be unified by DQFT in terms of constrained irreversible dynamic equations under the umbrella of Yang-Mills theories.
-
4.
The dissipation mechanism appearing in the QME (12) is formulated in terms of the creation and annihilation operators associated with the free Hamiltonian and hence consists of an exchange of particles between the system and its environment, where the exchange of high-energy particles is strongly favored. In our view, this irreversible contribution to dynamics suggests a particle ontology, also in the light of QFT phenomenology. Thus, the formal structure of the theory seems to reflect appropriately the experimental evidence available from particle accelerators.
-
5.
Finally, as we have seen in this section, the formulation of the thermodynamic QME of DQFT relies heavily on the Fock space associated with the creation operators of the momentum eigenstates of the free Hamiltonian, which we interpret as particles. The particle-free state vector |0〉 of the Fock space may be interpreted as the ground and vacuum state of the free theory. The density matrices obtained from the QME (12), in which the full Hamiltonian with all interactions is employed in formulating the reversible dynamics, describe the states of the fully interacting theory, including the steady state at a given temperature. In this picture, then, the vacuum states of the free and interacting theories have clearly distinct characters and significance, so that we get new insight into the problems raised by Haag’s theorem.