Classical Physic

In subject area: Mathematics

The deterministic account of classical physics which implies that any disturbance of a system composed of an immense number of parts invariably leads to chaotic disorder, is in quantum physics replaced by a description according to which the result of any interaction between atomic systems is the outcome of a competition between various individual processes by which the states of the new systems, like those of the original systems, in a simple way are defined by the atomic particles they contain.

From: Niels Bohr Collected Works, 1999

Add to Mendeley

Chapters and Articles

You might find these chapters and articles relevant to this topic.

3.1 The hard road to determinism in classical physics

Classical physics is widely assumed to provide a friendly environment for determinism. In fact, determinism must overcome a number of obstacles in order to achieve success in this setting. First, classical spacetime structure may not be sufficiently rich to support Laplacian determinism for particle motions. Second, even if the spacetime structure is rich, uniqueness can fail in the initial value problem for Newtonian equations of motion if the force function does not satisfy suitable continuity conditions. Third, the equations of motion that typically arise for classical particles plus classical fields, or for classical fields alone, do not admit an initial value formulation unless supplementary conditions are imposed. Fourth, even in cases where local (in time) uniqueness holds for the initial value problem, solutions can break down after a finite time.

The following subsection takes up the first of these topics — the connection between determinism and the structure and ontology of classical spacetimes. The others are taken up in due course.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780444515605500178

The second part takes up the basic physics. There is a chapter on the concepts of classical physics, followed by chapters on electromagnetism and on quantum mechanics. These chapters as much of the book contain both descriptive parts with attempts to show the general features and formula sections, where I go in some detail. The principles of these chapters will not be followed up in the rest, and they serve as reference to later developments. This part also contains the basics of thermodynamics and statistical physics, here in a descriptive style. These parts will be developed further in later parts, and shall be regarded as important main threads throughout the book.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780444527981500014
2012, Classical and Quantum InformationDan C. Marinescu, Gabriela M. Marinescu

Measurements, Preparation, and Extraction of Classical Information from Quantum Information

Classical information is independent of the medium used to transport it. Yet, classical information is often carried by the same types of particles as quantum information (e.g., electrons, or photons). Why should we expect quantum information to be different from classical information?

To answer this question, we should establish if it is possible to freely convert one type of information to another and then recover the original information [230]. We can convert classical to quantum information and then convert back the quantum information to the original classical information. Formally, this process consists of two stages: preparation, when the quantum information is generated from the classical one, and measurements, when classical information is obtained from the quantum information (Figure 1.6a).

FIGURE 1.6. Classical and quantum information and conversion from one to another. Classical information is represented by thin arrows and quantum information as thick arrows. (a) Classical information can be regarded as a particular form of quantum information. (b) Classical information can be recovered from the quantum information when the preparation phase is followed by a measurement. The conversion path is: classicalquantumclassical. (c) Quantum information cannot be recovered when the preparation follows the measurement; the measurement is an irreversible process and alters the state of quantum systems. The conversion path in this case is quantumclassicalquantum.

The remaining question is if we can convert quantum to classical information and then convert the classical information to quantum information indistinguishable from the original one; this means that we should first perform a measurement to extract classical information and then use it to prepare quantum information (Figure 1.6b). The only possibility to compare quantum mechanical systems is in terms of statistical experiments, and this is not possible; because a measurement is an irreversible process, it alters the state of a quantum system. Chapter 2, devoted to quantum measurements, covers the arguments supporting this statement.

We conclude that, indeed, quantum information is qualitatively different from classical information. Even though classical and quantum information can be carried out by the same types of particles, the physical processes are different; in the former case, the physical processes are subject to classical physics, and later, to quantum physics. As we shall see in Chapter 2, some states of a quantum system, the pure states, have a classical counterpart, and a measurement allows us to distinguish orthogonal pure states. Therefore, classical information can be regarded as a particular form of quantum information (Figure 1.6c).

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780123838742000011

3 DETERMINISM AND INDETERMINISM IN CLASSICAL PHYSICS

3.1 The hard road to determinism in classical physics

Classical physics is widely assumed to provide a friendly environment for determinism. In fact, determinism must overcome a number of obstacles in order to achieve success in this setting. First, classical spacetime structure may not be sufficiently rich to support Laplacian determinism for particle motions. Second, even if the spacetime structure is rich, uniqueness can fail in the initial value problem for Newtonian equations of motion if the force function does not satisfy suitable continuity conditions. Third, the equations of motion that typically arise for classical particles plus classical fields, or for classical fields alone, do not admit an initial value formulation unless supplementary conditions are imposed. Fourth, even in cases where local (in time) uniqueness holds for the initial value problem, solutions can break down after a finite time.

The following subsection takes up the first of these topics — the connection between determinism and the structure and ontology of classical spacetimes. The others are taken up in due course.

3.2 Determinism, spacetime structure, and spacetime ontology

Here is the (naive) reason for thinking that neither Laplacian determinism nor any of its cousins stands a chance unless supported by enough spacetime structure of the right kind. Assume that the (fixed) classical spacetime background is characterized by a differentiable manifold M and various geometric object fields O1, O2, …, OM on M. And assume that the laws of physics take the form of equations whose variables are the Oi's and additional object fields P1, P2, …, PN describing the physical contents of the spacetime. (For the sake of concreteness, the reader might want to think of the case where the Pj's are vector fields whose integral curves are supposed to be the world lines of particles.) A symmetry of the spacetime is a diffeomorphism d of M onto itself which preserves the background structure given by the Oi's — symbolically, d*Oi = Oi for all values if i, where d* denotes the drag along by d.17 By the assumption on the form of the laws, a spacetime symmetry d must also be a symmetry of the laws of motion in the sense that if M,O1,O2,,OM,P1,P2,,PN satisfies the laws of motion, then so does M,O1,O2,,OM,d*P1,d*P2,,d*PN.18

Now the poorer the structure of the background spacetime, the richer the spacetime symmetries. And if the spacetime symmetry group is sufficiently rich, it will contain elements that are the identity map on the portion of spacetime on or below some time slice t = const but non-identity above. We can call such a map a ‘determinism killing symmetry’ because when applied to any solution of the equations of motion, it produces another solution that is the same as the first for all past times but is different from the first at future times, which is a violation of even the weakest version of future Laplacian determinism.

As an example, take Leibnizian spacetime,19 whose structure consists of all and only the following: a notion of absolute or observer-independent simultaneity; a temporal metric (giving the lapse of time between non-simultaneous events); and a Euclidean spatial metric (giving the spatial distance between events lying on a given plane of absolute simultaneity). In a coordinate system (xα, t), α = 1, 2, 3 adapted to this structure, the spacetime symmetries are

(1)xαxα=xα=Rβα(t)xβ+aα(t)α,β=1,2,3,tt=t+const

where Rβα(t) is an orthogonal time dependent matrix and the aα(t) are arbitrary smooth functions of t. Clearly, the symmetries (1) contain determinism killing symmetries.

It is also worth noting that if the structure of spacetime becomes very minimal, no interesting laws of motion, deterministic or not, seem possible. For example, suppose that the time metric and the space metric are stripped from Leibnizian spacetime, leaving only the planes of absolute simultaneity. And suppose that the laws of physics specify that the world is filled with a plenum of constant mass dust particles and that the world lines of these particles are smooth curves that never cross. Then either every smooth, non-crossing motion of the dust is allowed by the laws of motion or none is, for any two such motions are connected by a symmetry of this minimal spacetime.

Two different strategies for saving determinism in the face of the above construction can be tried. They correspond to radically different attitudes towards the ontology of spacetime. The first strategy is to beef up the structure of the background spacetime. Adding a standard of rotation kills the time dependence in Rβα(t), producing what is called Maxwellian spacetime. But since the aα(t) are still arbitrary functions of t there remain determinism killing symmetries. Adding a standard of inertial or straight line motion linearizes the aα(t) to vαt + cα, where the vα and cα are constants, producing neo-Newtonian spacetime20 whose symmetries are given by the familiar Galilean transformations

(2)xαxα=Rβαxβ+vαt+cαα,β=1,2,3,tt=t+const

The mappings indicated by (2) do not contain determinism killing symmetries since if such a map is the identity map for a finite stretch of time, no matter how short, then it is the identity map period. Note that this way of saving determinism carries with it an allegiance to “absolute” quantities of motion: in neo-Newtonian spacetime it makes good sense to ask whether an isolated particle is accelerating or whether an isolated extended body is rotating. To be sure, this absolute acceleration and rotation can be called ‘relational’ quantities, but the second place in the relation is provided by the structure of the spacetime — in particular, by the inertial structure — and not by other material bodies, as is contemplated by those who champion relational accounts of motion.

The second strategy for saving determinism proceeds not by beefing up the structure of the background spacetime but by attacking a hidden assumption of the above construction — the “container view” of spacetime. Picturesquely, this assumption amounts to thinking of spacetime as a medium in which particles and fields reside. More precisely, in terms of the above apparatus, it amounts to the assumption that M,O1,O2,,OM,P1,P2,,PN and M,O1,O2,,OM,d*P1,d*P2,,d*PN, where d is any diffeomorphism of M such that d*PjPj for some j, describe different physical situations, even when d is a spacetime symmetry, i.e. d*Oi = Oi for all i. Rejecting the container view leads to (one form of) relationism about spacetime. A spacetime relationist will take the above construction to show that, on pain of abandoning the possibility of determinism, those who are relationists about motion should also be relationists about spacetime. Relationists about motion hold that talk of absolute motion is nonsensical and that all meaningful talk about motion must be construed as talk about the relative motions of material bodies. They are, thus, unable to avail themselves of the beef-up strategy for saving determinism; so, if they want determinism, they must grasp the lifeline of relationism about spacetime.

Relationalism about motion is a venerable position, but historically it has been characterized more by promises than performances. Newton produced a stunningly successful theory of the motions of terrestrial and celestial bodies. Newton's opponents promised that they could produce theories just as empirically adequate and as explanatorily powerful as his without resorting to the absolute quantities of motion he postulated. But mainly what they produced was bluster rather than workable theories.21 Only in the twentieth century were such theories constructed (see [Barbour, 1974] and [Barbour and Bertotti, 1977]; and see [Barbour, 1999] for the historical antecedents of these theories), well after Einstein's GTR swept away the notion of a fixed background spacetime and radically altered the terms of the absolute vs. relational debate.

3.3 Determinism and gauge symmetries

When philosophers hear the word “gauge” they think of elementary particle physics, Yang-Mills theories, etc. This is a myopic view. Examples of non-trivial gauge freedom arise even in classical physics — in fact, we just encountered an example in the preceding subsection. The gauge notion arises for a theory where there is “surplus structure” (to use Michael Redhead's phrase) in the sense that the state descriptions provided by the theory correspond many-one to physical states. For such a theory a gauge transformation is, by definition, a transformation that connects those descriptions that correspond to the same physical state.

The history of physics shows that the primary reason for seeing gauge freedom at work is to maintain determinism. This thesis has solid support for the class of cases of most relevance to modern physics, viz. where the equations of motion/field equations are derivable from an action principle and, thus, the equations of motion are in the form of Euler-Lagrange equations.22 When the Lagrangian is non-singular, the appropriate initial data picks out a unique solution of the Euler-Lagrange equations and Laplacian determinism holds.23 If, however, the action admits as variational symmetries a Lie group whose parameters are arbitrary functions of the independent variables, then we have a case of underdetermination because Noether's second theorem tells us that the Euler-Lagrange equations have to satisfy a set of mathematical identities.24 When these independent variables include time, arbitrary functions of time will show up in solutions to the Euler-Lagrange equations, apparently wrecking determinism.

The point can be illustrated with the help of a humble example of particle mechanics constructed within the Maxwellian spacetime introduced in the preceding subsection. An appropriate Lagrangian invariant under the symmetries of this spacetime is given by

(3)L=mjmk2M(X˙jX˙k)2V(|XjSk|),M:=imi.

This Lagrangian is singular in the sense that Hessian matrix 2L/x˙ix˙j does not have an inverse. The Euler-Lagrange equations are

(4)ddt(mi(x˙j1Mkmkx˙k))=Vx˙i.

These equations do not determine the evolution of the particle positions uniquely: if xi(t) is a solution, so is xi(t)=xi(t)+f(t), for arbitrary f (t), confirming the intuitive argument given above for the apparent breakdown of determinism. Determinism can be restored by taking the transformation xi(t) → xi(t) + f (t) as a gauge transformation.

The systematic development of this approach to gauge was carried out by P. A. M. Dirac in the context of the Hamiltonian formalism.25 A singular Lagrangian system corresponds to a constrained Hamiltonian system. The primary constraints appear as a result of the definition of the canonical momenta. (In the simple case of a first-order Lagrangian L(q,q˙,t), where q stands for the configuration variables and q˙:dq/dt, the canonical momentum is p:L/q˙) The secondary constraints arise as a consequence of the demand that the primary constraints be preserved by the motion. The total set of constraints picks out the constraint surface C(q,p) of the Hamiltonian phase space γ(q, p). The first class constraints are those that commute on C(q,p) with all of the constraints. It is these first class constraints that are taken as the generators of the gauge transformations. The gauge invariant quantities (a.k.a. “observables”) are then the phase function F: γ(q, p) ℝ that are constant along the gauge orbits.

Applying the formalism to our toy case of particle mechanics in Maxwellian spacetime, the canonical momenta are:

(5)pi:=Lx˙i=miMkmk(x˙ix˙k)=mix˙imiMkmkx˙k.

These momenta are not independent but must satisfy three primary constraints, which require the vanishing of the x, y, and z-components of the total momentum:

(6)Φα=ipiα=0,α=1,2,3.

These primary constraints are the only constraints — there are no secondary constraints — and they are all first class. These constraints generate in each configuration variable xi the same gauge freedom; namely, a Euclidean shift given by the same arbitrary function of time. The gauge invariant variables, such relative particle positions and relative particle momenta, do evolve deterministically.

The technical elaboration of the constraint formalism is complicated, but one should not lose sight of the fact that the desire to save determinism is the motivation driving the enterprise. Here is a relevant passage from [Henneaux and Teitelboim, 1992], one of the standard references on constrained Hamiltonian systems:

The presence of arbitrary functions … in the total Hamiltonian tells us that not all the q's and p's [the configuration variables and their canonical momenta] are observable [i.e. genuine physical magnitudes]. In other words, although the physical state is uniquely defined once a set of q's and p's is given, the converse is not true — i.e., there is more than one set of values of the canonical variables representing a given physical state. To see how this conclusion comes about, we note that if we are given an initial set of canonical variables at the time t1 and thereby completely define the physical state at that time, we expect the equations of motion to fully determine the physical state at other times. Thus, by definition, any ambiguity in the value of the canonical variables at t2t1 should be a physically irrelevant ambiguity. [pp. 16–17]

As suggested by the quotation, the standard reaction to the apparent failure of determinism is to blame the appearance on the redundancy of the descriptive apparatus: the correspondence between the state descriptions in terms of the original variables — the q's and p's — and the physical state is many-to-one; when this descriptive redundancy is removed, the physical state is seen to evolve deterministically. There may be technical difficulties is carrying through this reaction. For example, attempting to produce a reduced phase space — whose state descriptions corresponding one-one to physical states — by quotienting out the gauge orbits can result in singularities. But when such technical obstructions are not met, normal (i.e. unconstrained) Hamiltonian dynamics applies to the reduced phase space, and the reduced phase space variables evolve deterministically.

In addition to this standard reaction to the apparent failure of determinism in the above examples, two others are possible. The first heterodoxy takes the apparent violation of determinism to be genuine. This amounts to (a) treating what the constraint formalism counts as gauge dependent quantities as genuine physical magnitudes, and (b) denying that these magnitudes are governed by laws which, when conjoined with the laws already in play, restore determinism. The second heterodoxy accepts the orthodox conclusion that the apparent failure of determinism is merely apparent; but it departs from orthodoxy by accepting (a), and it departs from the first heterodoxy by denying (b) and, accordingly, postulates the existence of additional laws that restore determinism. Instances that superficially conform to part (a) of the two heterodoxies are easy to construct from examples found in physics texts where the initial value problem is solved by supplementing the equations of motion, stated in terms of gauge-dependent variables, with a gauge condition that fixes a unique solution. For instance, Maxwell's equations written in terms of electromagnetic potentials do not determine a unique solution corresponding to the initial values of the potentials and their time derivatives. Imposing the Lorentz gauge condition converts Maxwell's equations to second order hyperbolic partial differential equations (pdes) that do admit an initial value formulation (see Section 4.2).26 Similar examples can be concocted in general relativity theory where orthodoxy treats the metric potentials as gauge variables (see Section 6.2). In these examples orthodoxy is aiming to get at the values of the gauge independent variables via a choice of gauge. If this aim is not kept clearly in mind, the procedure creates the illusion that gauge-dependent variables have physical significance. It is exactly this illusion that the two heterodoxies take as real. The second heterodoxy amounts to taking the gauge conditions not as matters of calculational convenience but as additional physical laws. I know of no historical examples where this heterodoxy has led to fruitful developments in physics.

Since there is no a priori guarantee that determinism is true, the fact that the orthodox reading of the constraint formalism guarantees that the equations of motion admit an initial value formulation must mean that substantive assumptions that favor determinism are built into the formalism. That is indeed the case, for the Lagrangian/Hamiltonian formalism imposes a structure on the space of solutions: in the geometric language explained in Chapter 1 and 2 of this volume, the space of solutions has a symplectic or pre-symplectic structure. This formalism certainly is not guaranteed to be applicable to all of the equations of motion the Creator might have chosen as laws of motion; indeed, it is not even guaranteed to be applicable to all Newtonian type second order odes. In the 1880s Helmholtz found a set of necessary conditions for equations of this type to be derivable from an action principle; these conditions were later proved to be (locally) sufficient as well as necessary. After more than a century, the problem of finding necessary and sufficient conditions for more general types of equations of motion, whether in the form of odes or pdes, to be derivable from an action principle is still an active research topic.27

3.4 Determinism for fields and fluids in Newtonian physics

Newtonian gravitational theory can be construed as a field theory. The gravitational force is given by Fgrav = −Δ φ, where the gravitational potential φ satisfies the Poisson equation

(7)2φ=ρ

with ρ being the mass density. If φ is a solution to Poisson's equation, then so is φ′ = φ + g (x)f (t) where g (x) is a linear function of the spatial variables and f (t) is an arbitrary function of t. Choose f so that f (t) = 0 for t ≤ 0 but f (t) > 0 for t > 0. The extra gravitational force, proportional to f (t), that a test particle experiences in the primed solution after t = 0 is undetermined by anything in the past.

The determinism wrecking solutions to (7) can be ruled out by demanding that gravitational forces be tied to sources. But to dismiss homogeneous solutions to the Poisson equation is to move in the direction of treating the Newtonian gravitational field as a mere mathematical device that is useful in describing gravitational interactions which, at base, are really direct particle interactions.28 In this way determinism helps to settle the ontology of Newtonian physics: the insistence on determinism in Newtonian physics demotes fields to second-class status. In relativistic physics fields come into their own, and one of the reasons is that the relativistic spacetime structure supports field equations that guarantee deterministic evolution of the fields (see Section 4.2).

In the Newtonian setting the field equations that naturally arise are elliptic (e.g. the Poisson equation) or parabolic, and neither type supports determinism-without-crutches. An example of the latter type of equation is the classical heat equation

(8)2Φ=kΦt

where Φ is the temperature variable and k is the coefficient of heat conductivity.29 Solutions to (8) can cease to exist after a finite time because the temperature “blows up.” Uniqueness also fails since, using the fact that the heat equation propagates heat arbitrarily fast, it is possible to construct surprise solutions Φs with the properties that (i) Φs is infinitely differentiable, and (ii) Φs(x, t) = 0 for all t ≤ 0 but Φs(x, t) ≠ 0 for t > 0 (see [John, 1982, Sec. 7.1]). Because (8) is linear, if Φ is a solution then so is Φ′ = Φ + Φs. And since Φ and Φ′ agree for all t ≠ 0 but differ for t > 0, the existence of the surprise solutions completely wrecks determinism.

Uniqueness of solution to (8) can be restored by adding the requirement that Φ ≥ 0, as befits its intended interpretation of Φ as temperature; for Widder [1975, 157] has shown that if a solution of Φ(x, t) of (8) vanishes at t = 0 and is nonnegative for all x and all t ≥ 0, then it must be identically zero. But one could have wished that, rather than having to use a stipulation of non-negativity to shore up determinism, determinism could be established and then used to show that if the temperature distribution at t = 0 is non-negative for all x, then the uniquely determined evolution keeps the temperature non-negative. Alternatively, both uniqueness and existence of solutions of (8) can be obtained by limiting the growth of |Φ(x, t)| as |x| → ∞. But again one could have wished that such limits on growth could be derived as a consequence of the deterministic evolution rather than having to be stipulated as conditions that enable determinism.

Appearances of begging the question in favor of determinism could be avoided by providing at the outset a clear distinction between kinematics and dynamics, the former being a specification of the space of possible states. For example, a limit on the growth of quantum mechanical wave functions does not beg the question of determinism provided by the Schrödinger equation since the limit follows from the condition that the wave function is an element of a Hilbert space, which is part of the kinematical prescription of QM (see Section 5). Since this prescription is concocted to underwrite the probability interpretation of the wave function, we get the ironic result that the introduction of probabilities, which seems to doom determinism, also serves to support it. The example immediately above, as well as the examples of the preceding subsection and the one at the beginning of this subsection, indicate that in classical physics the kinematical/dynamical distinction can sometimes be relatively fluid and that considerations of determinism are used in deciding where to draw the line. The following example will reinforce this moral.30

The Navier-Stokes equations for an incompressible fluid moving in N,N=2,3, read

(9a)Dudt=p+vΔu 
(9b)div(u)=0

where u (x, t) = (u1, u2, …, uN) is the velocity of the fluid, p (x, t) is the pressure, v = const. ≥ 0 is the coefficient of viscosity, and D/dt:=/t+j=1Nuj/xj is the convective derivative (see Foias at al. 2001 for a comprehensive survey). If the fluid is subject to an external force, an extra term has to be added to the right hand side of (9a). The Euler equations are the special case where u = 0. The initial value problem for (9a-b) is posed by giving the initial data

(9)u(x,0)=u0(x)

where u0 (x) is a smooth (C) divergence-free vector field, and is solved by smooth functions u,pC(Nx[0,)) satisfying (9)-(10). For physically reasonable solutions it is required both that u0(x) should not grow too large as |x| → ∞ and that the energy of the fluid is bounded for all time:

(10)N|u(x,t)|2dx<for all t>0.

When v = 0 the energy is conserved, whereas for v > 0 it dissipates.

For N = 2 it is known that a physically reasonable smooth solution exists for any given u0(x). For N = 3 the problem is open. However, for this case it is known that the problem has a positive solution if the time interval [0, ∞) for which the solution is required to exist is replaced by [0, T) where T is a possibly finite number that depends on u0(x). When T is finite it is known as the “blowup time” since |u (x, t)| must become unbounded as t approaches T. For the Euler equations a finite blowup time implies that the vorticity (i.e. the curl of u (x, t)) becomes unbounded as t approaches T.

Smooth solutions to the Navier-Stokes equations, when they exist, are known to be unique. This claim would seem to be belied by the symmetries of the Navier-Stokes equations since if u (x, t) = f (x, t), p (x, t) = g (x, t) is a solution then so is the transformed u˜(x,t)=f(xɛα(t),t)+ɛαt,p˜(x,t)=g(xɛα(t),t)ɛxαt+12ɛ2αtt, where α (t) is an arbitrary smooth function of t alone (see Olver 1993, pp. 130 and 177 (Exer. 2.15)). Choosing α (t) such that α (0) = αt(0) = αtt(0) = 0 but α (t) ≠ 0 for t > 0 results in different solutions for the same initial data unless f (xɛ α (t), t) + ɛαt = f (x, t). However, the transformed solution violates the finiteness of energy condition (11).

The situation on the existence of solutions can be improved as follows. Multiplying (9a-b) by a smooth test function and integrating by parts over x and t produces integral equations that are well-defined for any u (x, t) and p (x, t) that are respectively L2 (square integrable) and L1 (integrable). Such a pair is called a weak solution if it satisfies the integral equations for all test functions. Moving from smooth to weak solutions permits the proof of the existence of a solution for all time. But the move reopens the issue of uniqueness, for the uniqueness of weak solutions for the Navier-Stokes equations is not settled. A striking non-uniqueness result for weak solutions of the Euler equations comes from the construction by Scheffer [1994] and Shnirelman [1997] of self-exciting/self-destroying weak solutions: u (x, t) ≡ 0 for t < −1 and t > 1, but is non-zero between these times in a compact region of 3.

It is remarkable that basic questions about determinism for classical equations of motion remain unsettled and that these questions turn on issues that mathematicians regard as worthy of attention. Settling the existence question for smooth solutions for the Navier-Stokes equations in the case of N = 3 brings a $1 million award from the Clay Mathematics Institute (see [Fefferman, 2000]).

3.5 Continuity issues

Consider a single particle of mass m moving on the real line in a potential V (x), x ∈ℝ. The standard existence and uniqueness theorems for the initial value problem of odes can be used to show that the Newtonian equation of motion

(11)mx¨=F(x):=dVdx

has a locally (in time) unique solution if the force function F (x) satisfies a Lipschitz condition.31 An example of a potential that violates the Lipschitz condition at the origin is 92|x|4/3. For the initial data x(0)=0=x˙(0) there are multiple solutions of (12): x (t) ≡ 0, x (t) = t3, and x (t) = −t3, where m has been set to unity for convenience. In addition, there are also solutions x (t) where x (t) = 0 for t < k and ±(tk)3 for tk, where k is any positive constant. That such force functions do not turn up in realistic physical situations is an indication that Nature has some respect for determinism. In QM it turns out that Nature can respect determinism while accommodating some of the non-Lipschitz potentials that would wreck Newtonian determinism (see Section 5.2).

3.6 The breakdown of classical solutions

Consider again the case of a single particle of mass m moving on the real line ℝ in a potential V (x), and suppose that V (x) satisfies the Lipschitz condition, guaranteeing a temporally local unique solution for the initial value problem for the Newtonian equations of motion. However, determinism can fail if the potential is such that the particle is accelerated off to −∞ or +∞ in a finite time.32 Past determinism is violated because two such solutions can agree for all future times tt* (say) — no particle is present at these times anywhere in space — but disagree at past times t < t* on the position and/or velocity of the particle when it is present in space. Since the potential is assumed to be time independent, the equations of motion are time reversal invariant, so taking the time reverses of these escape solutions produces solutions in which hitherto empty space is invaded by particles appearing from spatial infinity. These invader solutions provide violations of future determinism. Piecing together escape and invader solutions produces further insults to determinism.

In the 1890's Paul Painlevé conjectured that for N > 3 point mass particles moving in ℝ3 under their mutually attractive Newtonian gravitational forces, there exist solutions to the Newtonian equations of motion exhibiting non-collision singularities, i.e. although the particles do not collide, the solution ceases to exist after a finite time. Hugo von Zeipel [1908] showed that in such a solution the particle positions must become unbounded in a finite time. Finally, near the close of the 20th century Xia [1992] proved Painlevé conjecture by showing that for N = 5 point mass particles, the Newtonian equations of motion admit solutions in which the particles do not collide but nevertheless manage to accelerate themselves off to spatial infinity in a finite time (see [Saari and Xia, 1995] for an accessible survey).

Determinism can recoup its fortunes by means of the device, already mentioned above, of supplementing the usual initial conditions with boundary conditions at infinity. Or consolation can be taken from two remarks. The first remark is that in the natural phase space measure, the set of initial conditions that lead to Xia type escape solutions has measure zero. But it is unknown whether the same is true of all non-collision singularities. The second remark is that the non-collision singularities result from the unrealistic idealization of point mass particles that can achieve unbounded velocities in a finite time by drawing on an infinitely deep potential well. This remark does not suffice to save determinism when an infinity of finite sized particles are considered, as we will see in the next subsection.

It is interesting to note that for point particles moving under mutually attractive Newtonian gravitational forces, QM cures both the collision33 and non-collision singularities that can spell the breakdown of classical solutions (see Section 5.2). This is more than a mere mathematical curiosity since it is an important ingredient in the explanation of the existence and stability of the hydrogen atom.

3.7 Infinite collections

Consider a collection of billiard balls confined to move along a straight line in Euclidean space. Suppose that the balls act only by contact, that only binary collisions occur, and that each such collision obeys the classical laws of elastic impact. Surely, the reader will say, such a system is as deterministic as it gets. This is so, if the collection is finite. But if the collection is infinite and unbounded velocities are permitted, then determinism fails because even with all of the announced restrictions in place the system can seemingly self-excite itself (see [Lanford, 1974]). Pérez Laraudogoitia [2001] shows how to use such infinite collections to create an analogue of the escape solution of the preceding subsection where all of the particles disappear in a finite amount of time. The time reverse of this scenario is one in which space is initially empty, and then without warning an infinite stream of billiard balls pour in from spatial infinity.

Legislating against unbounded velocities or imposing boundary conditions at infinity does not suffice to restore determinism if the billiard balls can be made arbitrarily small [Pérez Laraudogoitia, 2001]. For then a countably infinite collection of them can be Zeno packed into a finite spatial interval, say (0, 1], by placing the center of the first ball at 1, the second at 1/2, the third at 1/4, etc. Assume for ease of illustration that all the balls have equal mass (≡ 1). A unit mass cue ball moving with unit speed from right to left collides with the first ball and sends a ripple through the Zeno string that lasts for unit time, at the end of which all of the balls are at rest. The boring history in which all the balls are at rest for all time is, of course, also a solution of the laws of impact. Comparing this boring history with the previous one shows that past Laplacian determinism is violated.34

This failure of determinism carries with it a violation of the conservation and energy momentum, albeit in a weak sense; namely, in the inertial frame in which the object balls are initially at rest, the total energy and the total momentum each have different values before and after the collisions start, but in every other inertial frame there is no violation simply because the values are infinite both before and after the collisions.35 Pérez Laraudogoitia [2005] has shown how to construct scenarios in which there is a strong violation of conservation of energy and momentum in that the violation occurs in every inertial frame.

3.8 Domains of dependence

With some artificiality one of the threats to classical determinism discussed above can be summarized using a concept that will also prove very helpful in comparing the fortunes of determinism in classical physics and in relativistic physics. By a causal curve let us understand a (piecewise) smooth curve in spacetime that represents the spacetime trajectory for a physically possible transfer of energy/momentum. Define the future domain of dependence, D+(S), of a spacetime region S as the set of all spacetime points p such that any past directed causal curve with future endpoint at p and no past endpoint intersects S. The past domain of dependence D(S) of S is defined analogously. And the total domain of dependence D (S) is the union D+(S) ∪ D(S). If pD (S) then it would seem that the state in region S does not suffice to determine the state at p since there is a possible causal process that passes through p but never registers on S.

Since neither the kinematics nor the dynamics of classical physics place an upper bound on the velocity at which energy/momentum can be transferred, it would seem that in principle any timelike curve — i.e. any (piecewise) smooth curve oblique to the planes of absolute simultaneity — can count as a causal curve, and as a consequence D (S) = ø even when S is taken to be an entire plane of absolute simultaneity. The examples from Sections 3.4, 3.6, and 3.7 show how the “in principle” can be realized by some systems satisfying Newtonian laws of motion.

We have seen that some threats to classical determinism can be met by beefing up the structure of classical spacetime. And so it is with the threat currently under consideration. Full Newtonian spacetime is what results from neo-Newtonian spacetime by adding absolute space in the form of a distinguished inertial frame (‘absolute space’). In this setting the spacetime symmetries are small enough that there are now finite invariant velocities (intuitively, velocities as measured relative to absolute space), and thus laws can be formulated that set a finite upper bound on the absolute velocity of causal propagation. Nor is this move necessarily ad hoc as shown, for example, by the fact that the formulation of Maxwell's laws of electromagnetism in a classical spacetime setting evidently requires the services of a distinguished inertial frame, the velocity of light c being the velocity as measured in this frame.

But, as is well known, such a formulation is embarrassed by the undetectability of motion with respect to absolute space. This embarrassment provides a direct (albeit anachronistic) route from classical to relativistic spacetime. Adopting for classical spacetimes the same geometric language used in the special and general theories of relativity (see [Earman, 1989, Ch. 2]), absolute space is represented by a covariantly constant timelike vector field Aa, the integral curves of which are the world lines of the points of absolute space. The space metric is represented by a degenerate second rank contravariant tensor hab, which together with Aa defines a tensor that is formally a Minkowski metric: ηab:= habAa Ab. The unobservability of absolute motion means that there is no preferred way to split ηab into an hab part and a Aa Ab part, suggesting that ηab is physically as well as formally a Lorentz metric. As we will see in Section 4.1, this puts determinism on much firmer ground in that domains of dependence of local or global time slices are non-empty in the spacetime setting of STR.

3.9 Determinism, predictability, and chaos

Laplace's vision of a deterministic universe makes reference to an “intelligence” (which commentators have dubbed ‘Laplace's Demon’):

We ought to regard the present state of the universe as the effect of its antecedent state and as the cause of the state that is to follow. An intelligence knowing all of the forces acting in nature at a given instant, as well as the momentary positions of all things in the universe, would be able to comprehend in one single formula the motions of the largest bodies as well as the lightest atoms in the world, provided that its intellect were sufficiently powerful to subject all data to analysis; to it nothing would be uncertain, the future as well as the past would be present to its eyes.36

Perhaps by taking Laplace's vision too literally, philosophers and physicists alike conflate determinism and predictability. The conflation leads them to reason as follows: here is a case where predictability fails; thus, here is a case where determinism fails.37 This is a mistake that derives from a failure to distinguish determinism — an ontological doctrine about how the world evolves — from predictability — an epistemic doctrine about what can inferred, by various restricted means, about the future (or past) state of the world from a knowledge of its present state.

There is, however, an interesting connection between determinism and practical predictability for laws of motion that admit an initial value problem that is well-posed in the sense that, in some appropriate topology, the solutions depend continuously on the initial data.38 The standard existence and uniqueness proofs for the initial value problem for the odes used in particle mechanics also furnish a proof of well-posedness, which can be traced to the fact that the existence proof is constructive in that it gives a procedure for constructing a series of approximations that converge to the solution determined by the initial data.

To illustrate the implications of well-posedness for predictability, consider the toy case of a system consisting of a single massive particle obeying Newtonian equations of motion. If a suitable Lipschitz condition is satisfied, then for any given values of the position q (0) and velocity q˙(0) of the particle at t = 0 there exists (for some finite time interval surrounding t = 0) a unique solution: symbolically q(t)=F(q(0),q˙(0),t). And further, since this initial value problem is well-posed, for any fixed t > 0 (within the interval for which the solution is guaranteed to exist), F is a continuous function of q (0) and q˙(0). Suppose then that the practical prediction task is to forecast the actual position q¯(t*) of the particle at some given t* > 0 with an accuracy of ɛ > 0, and suppose that although measurements of position or velocity are not error free, the errors can be made arbitrarily small. By the continuity of F, there exist δ1 > 0 and δ2 > 0 such that if |q(0)q¯(0)|<δ1 and |q˙(0)q˙¯(0)|<δ2, then |q (t*) – q¯(t*)| < ɛ. Thus, measuring at t = 0 the actual particle position and velocity with accuracies ±δ1/2 and ±δ2/2 respectively ensures that when the measured values are plugged into F, the value of the function for t = t* answers to the assigned prediction task. (Note, however, that since the actual initial state is unknown, so are the required accuracies ±δ1/2 and ±δ2/2, which may depend on the unknown state as well as on ɛ and t*. This hitch could be overcome if there were minimum but non-zero values of δ1 and δ2 that answered to the given prediction task whatever the initial state; but there is no a priori guarantee that such minimum values exist. A prior measurement with known accuracy of the position and velocity at some t** < 0 will put bounds, which can be calculated from F, on the position and velocity at t = 0. And then the minimum values can be calculated for accuracies δ1 and δ2 of measurements at t = 0 that suffice for the required prediction task for any values of the position and velocity within the calculated bounds.)

Jacques Hadamard, who made seminal contributions to the Cauchy or initial value problem for pdes, took the terminology of “well-posed” (a.k.a. “properly posed”) quite literally. For he took it as a criterion for the proper mathematical description of a physical system that the equations of motion admit an initial value formulation in which the solution depends continuously on the initial data (see [Hadamard, 1923, 32]). However, the standard Courant-Hilbert reference work, Methods of Mathematical Physics, opines that

“properly posed” problems are by far not the only ones which appropriately reflect real phenomena. So far, unfortunately, little mathematical progress has been made in the important task of solving or even identifying such problems that are not “properly posed” but still are important and motivated by realistic situations. [1962, Vol. 2, 230].

Some progress can be found in [Payne, 1975] and the references cited therein.

Hadamard was of the opinion that if the time development of a system failed to depend continuously on the initial conditions, then “it would appear to us as being governed by pure chance (which, since Poincaré,39 has been known to consist precisely in such a discontinuity in determinism) and not obeying any law whatever” [1923, 38]. Currently the opinion is that the appearance of chance in classical systems is due not to the failure of well-posedness but to the presence of chaos.

The introduction of deterministic chaos does not change any of the above conclusions about determinism and predictability. There is no generally agreed upon definition of chaos, but the target class of cases can be picked out either in terms of cause or effects. The cause is sensitive dependence of solutions on initial conditions, as indicated, for example, by positive Lyapunov exponents. The effects are various higher order ergodic properties, such as being a mixing system, being a K-system, being a Bernoulli system, etc.40 Generally a sensitive dependence on initial conditions plus compactness of the state space is sufficient to secure such properties. The sensitive dependence of initial condition that is the root cause of chaotic behavior does not contradict the continuous dependence of solutions on initial data, and, therefore, does not undermine the task of predicting with any desired finite accuracy the state at a fixed future time, assuming that error in measuring the initial conditions can be made arbitrarily small. If, however, there is a fixed lower bound on the accuracy of measurements — say, because the measuring instruments are macroscopic and cannot make discriminations below some natural macroscopic scale — then the presence of deterministic chaos can make some prediction tasks impossible. In addition, the presence of chaos means that no matter how small the error (if non zero) in ascertaining the initial conditions, the accuracy with which the future state can be forecast degrades rapidly with time. To ensure the ability to predict with some given accuracy ɛ > 0 for all t > 0 by ascertaining the initial conditions at t = 0 with sufficiently small error δ > 0, it would be necessary to require not only well-posedness but stability, which is incompatible with chaos.41

Cases of classical chaos also show that determinism on the microlevel is not only compatible with stochastic behavior at the macro level but also that the deterministic microdynamics can ground the macro-stochasticity. For instance, the lowest order ergodic property — ergodicity — arguably justifies the use of the microcanonical probability distribution and provides for a relative frequency interpretation; for it implies that the microcanonical distribution is the only stationary distribution absolutely continuous with respect to Lebesque measure and that the measure of a phase volume is equal to the limiting relative frequency of the time the phase point spends in the volume. In these cases there does not seem to be a valid contrast between “objective” and “epistemic” probabilities. The probabilities are epistemic in the sense that conditionalizing on a mathematically precise knowledge of the initial state reduces the outcome probability to 0 or 1. But the probabilities are not merely epistemic in the sense of merely expressing our ignorance, for they are supervenient on the underlying microdynamics.

Patrick Suppes [1991; 1993] has used such cases to argue that, because we are confined to the macrolevel, determinism becomes for us a “transcendental” issue since we cannot tell whether we are dealing with a case of irreducible stochasticity or a case of deterministic chaos. Although I feel some force to the argument, I am not entirely persuaded. There are two competing hypotheses to explain observed macro-stochasticity: it is due to micro-determinism plus sensitive dependence on initial conditions vs. it is due to irreducible micro-stochasticity. The work in recent decades on deterministic chaos supplies the details on how the first hypothesis can be implemented. The details of the second hypothesis need to be filled in; particular, it has to be explained how the observed macro-stochasticity supervenes on the postulated micro-stochasticity.42 And then it has to be demonstrated that the two hypotheses are underdetermined by all possible observations on the macrolevel. If both of these demands were met, we would be faced with a particular instance of the general challenge to scientific realism posed by underdetermination of theory by observational evidence, and all of the well-rehearsed moves and countermoves in the realism debate would come into play. But it is futile to fight these battles until some concrete version of the second hypothesis is presented.

3.10 Laplacian demons, prediction, and computability

Since we are free to imagine demons with whatever powers we like, let us suppose that Laplace's Demon can ascertain the initial conditions of the system of interest with absolute mathematical precision. As for computational ability, let us suppose that the Demon has at its disposal a universal Turing machine. As impressive as these abilities are, they may not enable the Demon to predict the future state of the system even if it is deterministic. Returning to the example of the Newtonian particle from the preceding subsection, if the values of the position and velocity of the particle at time t = 0 are plugged into the function F (q (0), q˙(0), t) that specifies the solution q (t), the result is a function F(t) of t; and plugging different values of the initial conditions results in different F(t) — indeed, by the assumption of determinism, the F(t)s corresponding to different initial conditions must differ on any finite interval of time no matter how small. Since there is a continuum of distinct initial conditions, there is thus a continuum of distinct F(t)s. But only a countable number of these F(t)s will be Turing computable functions.43 Thus, for most of the initial conditions the Demon encounters, it is unable to predict the corresponding particle position q (t) at t > 0 by using its universal Turing machine to compute the value of F(t) at the relevant value of t — in Pitowsky's [1996] happy turn of phrase, the Demon must consult an Oracle in order to make a sure fire prediction.

However, if q (0) and q˙(0) are both Turing computable real numbers, then an Oracle need not be consulted since the corresponding F(t) is a Turing computable function; and if t is a Turing computable real number, then so is F(t). This follows from the fact that the existence and uniqueness proofs for odes gives an effective procedure for generating a series of approximations that converges effectively to the solution; hence, if computable initial data are fed into the procedure, the result is an effectively computable solution function. Analogous results need not hold when the equations of motion are pdes. Jumping ahead to the relativistic context, the wave equation for a scalar field provides an example where Turing computability of initial conditions is not preserved by deterministic evolution (see Section 4.4).

A more interesting example where our version of Laplace's Demon must consult an Oracle has been discussed by Moore [1990; 1991] and Pitowsky [1996]. Moore constructed an embedding of an abstract universal Turing machine into a concrete classical mechanical system consisting of a particle bouncing between parabolic and flat mirrors arranged so that the motion of the particle is confined to a unit square. Using this embedding Moore was able to show how recursively unsolvable problems can be translated into prediction tasks about the future behavior of the particle that the Demon cannot carry out without help from an Oracle, even if it knows the initial state of the particle with absolute precision! For example, Turing's theorem says that there is no recursive algorithm to decide whether a universal Turing machine halts on a given input. Since the halting state of the universal Turing machine that has been embedded in the particle-mirror system corresponds to the particle's entering a certain region of the unit square to which it is thereafter confined, the Demon cannot predict whether the particle will ever enter this region. The generalization of Turing's theorem by Rice [1953] shows that many questions about the behavior of a universal Turing machine in the unbounded future are recursively unsolvable, and these logical questions will translate into physical questions about the behavior of the particle in the unbounded future that the Demon cannot answer without consulting an Oracle.

The reader might ask why we should fixate on the Turing notion of computability. Why not think of a deterministic mechanical system as an analogue computer, regardless of whether an abstract Turing machine can be embedded in the system? For instance, in the above example of the Newtonian particle with deterministic motion, why not say that the particle is an analogue computer whose motion “computes,” for any given initial conditions q (0), q˙(0), the possibly non-Turing computable function q (t) = F (q (0), q˙(0), t)? I see nothing wrong with removing the scare quotes and developing a notion of analogue computability along these lines. But the practical value of such a notion is dubious. Determining which function of t is being computed and accessing the value computed for various values of t requires ascertaining the particle position with unbounded accuracy.

Connections between non-Turing computability and general relativistic spacetimes that are inhospitable to a global version of Laplacian determinism will be mentioned below in Section 6.6.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780444515605500178

7 WHY CLASSICAL STATES AND OBSERVABLES?

‘We have found a strange footprint on the shores of the unknown. We have devised profound theories, one after another, to account for its origins. At last, we have succeeded in reconstructing the creature that made the footprint. And lo! It is our own.’ [Eddington, 1920, 200-201]

The conclusion of Sections 5 and 6 is that quantum theory may give rise to classical behaviour in certain states and with respect to certain observables. For example, we have seen that in the limit ℏ → 0 coherent states and operators of the form Qℏ (f), respectively, are appropriate, whereas in the limit N → ∞ one should use classical states (nomen est omen!) as defined in Subsection 6.2 and macroscopic observables. If, instead, one uses superpositions of such states, or observables with the wrong limiting behaviour, no classical physics emerges. Thus the question remains why the world at large should happen to be in such states, and why we turn out to study this world with respect to the observables in question. This question found its original incarnation in the measurement problem (cf. Subsection 2.5), but this problem is really a figure-head for a much wider difficulty.

Over the last 25 years,307 two profound and original answers to this question have been proposed.

7.1 Decoherence

The first goes under the name of decoherence. Pioneering papers include [van Kampen, 1954; Zeh, 1970; Zurek, 1981; 1982],308 and [Joos and Zeh, 1985], and some recent reviews are [Bub, 1999; Auletta, 2001; Joos et al., 2003; Zurek, 2003; Blanchard and Olkiewicz, 2003; Bacciagaluppi, 2004] and [Schlosshauer, 2004].309 More references will be given in due course. The existence (and excellence) of these reviews obviates the need for a detailed treatment of decoherence in this article, all the more so since at the time of writing this approach appears to be in a transitional stage, conceptually as well as mathematically (as will be evident from what follows). Thus we depart from the layout of our earlier chapters and restrict ourselves to a few personal comments.

1.

Mathematically, decoherence boils down to the idea of adding one more link to the von Neumann chain (see Subsection 2.5) beyond S + A (i.e. the system and the apparatus). Conceptually, however, there is a major difference between decoherence and older approaches that took such a step: whereas previously (e.g., in the hands of von Neumann, London and Bauer, Wigner, etc.)310 the chain converged towards the observer, in decoherence it diverges away from the observer. Namely, the third and final link is now taken to be the environment (taken in a fairly literal sense in agreement with the intuitive meaning of the word). In particular, in realistic models the environment is treated as an infinite system (necessitating the limit N → ∞), which has the consequence that (in simple models where the pointer has discrete spectrum) the post-measurement state Σn cnΨn ⊗ ϕn ⊗ χn (in which the χn are mutually orthogonal) is only reached in the limit t → ∞. However, as already mentioned in Subsection 6.6, infinite time is only needed mathematically in order to make terms of the type ~ exp −γt (with γ > 0) zero rather than just very small: in many models the inner products (χn, χm) are actually negligible for nm within surprisingly short time scales.311

If only in view of the need for limits of the type N → ∞ (for the environment) and t → ∞, in our opinion decoherence is best linked to stance 1 of the Introduction: its goal is to explain the approximate appearance of the classical world from quantum mechanics seen as a universally valid theory. However, decoherence has been claimed to support almost any opinion on the foundations of quantum mechanics; cf. [Bacciagaluppi, 2004] and [Schlosshauer, 2004] for a critical overview and also see Point 3 below.

2.

Originally, decoherence entered the scene as a proposed solution to the measurement problem (in the precise form stated at the end of Subsection 2.5). For the restriction of the state Σn cnΨn ⊗ ϕn ⊗ χn to S + A (i.e. its trace over the degrees of freedom of the environment) is mixed in the limit t → ∞, which means that the quantum-mechanical interference between the states Ψn⊗ϕn for different values of n has become ‘delocalized’ to the environment, and accordingly is irrelevant if the latter is not observed (i.e. omitted from the description). Unfortunately, the application of the ignorance interpretation of the mixed post-measurement state of S + A is illegal even from the point of view of stance 1 of the Introduction. The ignorance interpretation is only valid if the environment is kept within the description and is classical (in having a commutative C*-algebra of observables). The latter assumption [Primas, 1983], however, makes the decoherence solution to the measurement problem circular.312

In fact, as quite rightly pointed out by Bacciagaluppi [2004], decoherence actually aggravates the measurement problem. Where previously this problem was believed to be man-made and relevant only to rather unusual laboratory situations (important as these might be for the foundations of physics), it has now become clear that “measurement” of a quantum system by the environment (instead of by an experimental physicist) happens everywhere and all the time: hence it remains even more miraculous than before that there is a single outcome after each such measurement. Thus decoherence as such does not provide a solution to the measurement problem [Leggett, 2002];313 Adler, 2003; Joos and Zeh, 2003], but is in actual fact parasitic on such a solution.

3.

There have been various responses to this insight. The dominant one has been to combine decoherence with some interpretation of quantum mechanics: decoherence then finds a home, while conversely the interpretation in question is usually enhanced by decoherence. In this context, the most popular of these has been the many-worlds interpretation, which, after decades of obscurity and derision, suddenly started to be greeted with a flourish of trumpets in the wake of the popularity of decoherence. See, for example, [Saunders, 1993; 1995; Joos et al., 2003] and [Zurek, 2003]. In quantum cosmology circles, the consistent histories approach has been a popular partner to decoherence, often in combination with many worlds; see below. The importance of decoherence in the modal interpretation has been emphasized by Dieks [1989b] and Bene and Dieks [2002], and practically all authors on decoherence find the opportunity to pay some lip-service to Bohr in one way or another. See [Bacciagaluppi, 2004] and [Schlosshauer, 2004] for a critical assessment of all these combinations.

In our opinion, none of the established interpretations of quantum mechanics will do the job, leaving room for genuinely new ideas. One such idea is the return of the environment: instead of “tracing it out”, as in the original setting of decoherence theory, the environment should not be ignored! The essence of measurement has now been recognized to be the redundancy of the outcome (or “record“) of the measurement in the environment. It is this very redundancy of information about the underlying quantum object that “objectifies” it, in that the information becomes accessible to a large number of observers without necessarily disturbing the object314 [Zurek, 2003; Ollivier et al. 2004; Blume-Kohout and Zurek, 2004; 2005]. This insight (called “Quantum Darwinism“) has given rise to the “existential” interpretation of quantum mechanics due to Zurek [2003].

4.

Another response to the failure of decoherence (and indeed all other approaches) to solve the measurement problem (in the sense of failing to win a general consensus) has been of a somewhat more pessimistic (or, some would say, pragmatic) kind: all attempts to explain the quantum world are given up, yielding to the point of view that ‘the appropriate aim of physics at the fundamental level then becomes the representation and manipulation of information’ [Bub, 2004]. Here ‘measuring instruments ultimately remain black boxes at some level’, and one concludes that all efforts to understand measurement (or, for that matter, EPR-correlations) are futile and pointless.315

5.

Night thoughts of a quantum physicist, then?316 Not quite. Turning vice into virtue: rather than solving the measurement problem, the true significance of the decoherence program is that it gives conditions under which there is no measurement problem! Namely, foregoing an explanation of the transition from the state Σn cnΨn ⊗ ϕn ⊗ χn of S + A + ε to a single one of the states Ψn ⊗ ϕn of S + A, at the heart of decoherence is the claim that each of the latter states is robust against coupling to the environment (provided the Hamiltonian is such that Ψn ⊗ ϕn tensored with some initial state Iε of the environment indeed evolves into Ψn ⊗ ϕn ⊗ χn, as assumed so far). This implies that each state Ψn ⊗ ϕn remains pure after coupling to the environment and subsequent restriction to the original system plus apparatus, so that at the end of the day the environment has had no influence on it. In other words, the real point of decoherence is the phenomenon of einselection (for environment-induced superselection), where a state is ‘einselected’ precisely when (given some interaction Hamiltonian) it possesses the stability property just mentioned. The claim, then, is that einselected states are often classical, or at least that classical states (in the sense mentioned at the beginning of this section) are classical precisely because they are robust against coupling to the environment. Provided this scenario indeed gives rise to the classical world (which remains to be shown in detail), it gives a dynamical explanation of it. But even short of having achieved this goal, the importance of the notion of einselection cannot be overstated; in our opinion, it is the most important and powerful idea in quantum theory since entanglement (which einselection, of course, attempts to undo!).

6.

The measurement problem, and the associated distinction between system and apparatus on the one hand and environment on the other, can now be omitted from decoherence theory. Continuing the discussion in Subsection 3.4, the goal of decoherence should simply be to find the robust or einselected states of a object O coupled to an environment ε, as well as the induced dynamics thereof (given the time-evolution of O + ε). This search, however, must include the correct identification of the object O within the total S + ε, namely as a subsystem that actually has such robust states. Thus the Copenhagen idea that the Heisenberg cut between object and apparatus be movable (cf. Subsection 3.2) will not, in general, extend to the “Primas-Zurek” cut between object and environment. In traditional physics terminology, the problem is to find the right “dressing” of a quantum system so as to make at least some of its states robust against coupling to its environment [Amann and Primas, 1997; Brun and Hartle, 1999; Omnès, 2002]. In other words: What is a system? To mark this change in perspective, we now change notation from Q (for “object”) to S (for “system”). Various tools for the solution of this problem within the decoherence program have now been developed — with increasing refinement and also increasing reliance on concepts from information theory [Zurek, 2003] — but the right setting for it seems the formalism of consistent histories, see below.

7.

Various dynamical regimes haven been unearthed, each of which leads to a different class of robust states [Joos et al., 2003; Zurek, 2003; Schlosshauer, 2004]. Here HS is the system Hamiltonian, HI is the interaction Hamiltonian between system and environment, and Hε is the environment Hamiltonian. As stated, no reference to measurement, object or apparatus need be made here.

In the regime HS << HI, for suitable Hamiltonians the robust states are the traditional pointer states of quantum measurement theory. This regime conforms to von Neumann's [1932] idea that quantum measurements be almost instantaneous. If, moreover, Hε << HI as well — with or without a measurement context — then the decoherence mechanism turns out to be universal in being independent of the details of ε and Hε [Strunz et al., 2003).

If HS Ã HI, then (at least in models of quantum Brownian motion) the robust states are coherent states (either of the traditional Schrödinger type, or of a more general nature as defined in Subsection 5.1); see [Zurek et al., 1993] and [Zurek, 2003]. This case is, of course, of supreme importance for the physical relevance of the results quoted in our Section 5 above, and — if only for this reason — decoherence theory would benefit from more interaction with mathematically rigorous results on quantum stochastic analysis.317

Finally, if HS >> HI, then the robust states turn out to be eigenstates of the system Hamiltonian HS [Paz and Zurek, 1999; Ollivier et al., 2004]. In view of our discussion of such states in Subsections 5.5 and 5.6, this shows that robust states are not necessarily classical. It should be mentioned that in this context decoherence theory largely coincides with standard atomic physics, in which the atom is taken to be the system S and the radiation field plays the role of the environment ε; see [Gustafson and Sigal, 2003] for a mathematically minded introductory treatment and [Bach et al., 1998; 1999] for a full (mathematical) meal.

8.

Further to the above clarification of the role of energy eigenstates, decoherence also has had important things to say about quantum chaos [Zurek, 2003; Joos et al., 2003]. Referring to our discussion of wave packet revival in Subsection 2.4, we have seen that in atomic physics wave packets do not behave classically on long time scales. Perhaps surprisingly, this is even true for certain chaotic macroscopic systems: cf. the case of Hyperion mentioned in the Introduction and at the end of Subsection 5.2. Decoherence now replaces the underlying superposition by a classical probability distribution, which reflects the chaotic nature of the limiting classical dynamics. Once again, the transition from the pertinent pure state of system plus environment to a single observed system state remains clouded in mystery. But granted this transition, decoherence sheds new light on classical chaos and circumvents at least the most flagrant clashes with observation.318

9.

Robustness and einselection form the state side or Schrödinger picture of decoherence. Of course, there should also be a corresponding observable side or Heisenberg picture of decoherence. But the transition between the two pictures is more subtle than in the quantum mechanics of closed systems. In the Schrödinger picture, the whole point of einselection is that most pure states simply disappear from the scene. This may be beautifully visualized on the example of a two-level system with Hilbert space HS = ℂ2 [Zurek, 2003]. If ↑ and ↓ (cf. (33)) happen to be the robust vector states of the system after coupling to an appropriate environment, and if we identify the corresponding density matrices with the north-pole (0, 0, 1) ∈ B3 and the south-pole (0, 0, −1) ∈ B3, respectively (cf. (3)), then following decoherence all other states move towards the axis connecting the north- and south poles (i.e. the intersection of the z-axis with B3) as t → ∞. In the Heisenberg picture, this disappearance of all pure states except two corresponds to the reduction of the full algebra of observables M2(ℂ) of the system to its diagonal (and hence commutative) subalgebra ℂ ⊕ ℂ in the same limit. For it is only the latter algebra that contains enough elements to distinguish ↑ and ↓ without containing observables detecting interference terms between these pure states.

10.

To understand this in a more abstract and general way, we recall the mathematical relationship between pure states and observables [Landsman, 1998]. The passage from a C*-algebra A of observables of a given system to its pure states is well known: as a set, the pure state space P(A) is the extreme boundary of the total state space S(A) (cf. footnote 259). In order to reconstruct A from P(A), the latter needs to be equipped with the structure of a transition probability space (see Subsection 6.3) through (27). Each element AA defines a function  on P(A) by Â(ω) = ω(A). Now, in the simple case that A is finite-dimensional (and hence a direct sum of matrix algebras), one can show that each function  is a finite linear combination of the form A = Σi pωi, where ωiP(A) and the elementary functions pρ on P(A) are defined by pρ(σ) = p (ρ, σ). Conversely, each such linear combination defines a function A for some AA. Thus the elements of A (seen as functions on the pure state space P(A)) are just the transition probabilities and linear combinations thereof. The algebraic structure of A may then be reconstructed from the structure of P(A) as a Poisson space with a transition probability (cf. Subsection 6.5). In this sense P(A) uniquely determines the algebra of observables of which it is the pure state space. For example, the space consisting of two points with classical transition probabilities (31) leads to the commutative algebra A = ℂ ⊕ ℂ, whereas the unit two-sphere in ℝ3 with transition probabilities (32) yields A = M2(ℂ).

This reconstruction procedure may be generalized to arbitrary C*-algebras [Landsman, 1998], and defines the precise connection between the Schrödinger picture and the Heisenberg picture that is relevant to decoherence. These pictures are equivalent, but in practice the reconstruction procedure may be difficult to carry through.

11.

For this reason it is of interest to have a direct description of decoherence in the Heisenberg picture. Such a description has been developed by Blanchard and Olkiewicz [2003], partly on the basis of earlier results by Olkiewicz [1999a, b; 2000]. Mathematically, their approach is more powerful than the Schrödinger picture on which most of the literature on decoherence is based. Let AS = B(HS) and Aε = B(Hε), and assume one has a total Hamiltonian H acting on HSHε as well as a fixed state of the environment, represented by a density matrix ρε (often taken to be a thermal equilibrium state). If ρS is a density matrix on HS (so that the total state is ρS ⊗ ρε), the Schrödinger picture approach to decoherence (and more generally to the quantum theory of open systems) is based on the time-evolution

(1)ρS(t)=TrHɛ(e-itHρSρEeitH)

The Heisenberg picture, on the other hand, is based on the associated operator time-evolution for AB(HS) given by

(2)A(t)=TrHɛ(ρEeitHA1e-itH),
since this yields the equivalence of the Schrödinger and Heisenberg pictures expressed by
(3)TrHS(ρS(t)A)=TrHS(ρSA(t))

More generally, let AS and Aε be unital C*-algebras with spatial tensor product ASAε, equipped with a time-evolution αt and a fixed state ωε on Aε. This defines a conditional expectation Pε: ASAεAS by linear and continuous extension of Pε(AB) = Aωε(B), and consequently a reduced time-evolution AA (t) on AS via

(4)A(t)=PE(αt(A1))

See, for example, Alicki and Lendi [1987]; in our context, this generality is crucial for the potential emergence of continuous classical phase spaces; see below.319 Now the key point is that decoherence is described by a decomposition AS = AS(1)A(2)S as a vector space (not as a C*-algebra), where A(1)S is a C*-algebra, with the property that limt→∞ A (t) = 0 (weakly) for all AAS(2), whereas AA (t) is an automorphism on AS(1) for each finite t. Consequently, A(1)S is the effective algebra of observables after decoherence, and it is precisely the pure states on A(1)S that are robust or einselected in the sense discussed before.

12.

For example, if AS = M2(ℂ) and the states ↑ and ↓ are robust under decoherence, then AS = ℂ ⊕ ℂ and A(2)S consists of all 2 × 2 matrices with zeros on the diagonal. In this example A(1)S is commutative hence classical, but this may not be the case in general. But if it is, the automorphic time-evolution on AS(1) induces a classical flow on its structure space, which should be shown to be Hamiltonian using the techniques of Section 6.320

In any case, there will be some sort of classical behaviour of the decohered system whenever AS(1) has a nontrivial center.321 If this center is discrete, then the induced time-evolution on it is necessarily trivial, and one has the typical measurement situation where the center in question is generated by the projections on the eigenstates of a pointer observable with discrete spectrum. This is generic for the case where AS is a type i factor. However, type ii and iii factors may give rise to continuous classical systems with nontrivial time-evolution; see [Lugiewicz and Olkiewicz, 2002; 2003]. We cannot do justice here to the full technical details and complications involved here. But we would like to emphasize that further to quantum field theory and the theory of the thermodynamic limit, the present context of decoherence should provide important motivation for specialists in the foundations of quantum theory to learn the theory of operator algebras.322

7.2 Consistent histories

Whilst doing so, one is well advised to work even harder and simultaneously familiarize oneself with consistent histories. This approach to quantum theory was pioneered by Griffiths [1984] and was subsequently taken up by Omnès [1992] and others. Independently, Gell-Mann and Hartle [1990; 1993] proposed analogous ideas. Like decoherence, the consistent histories method has been the subject of lengthy reviews [Hartle, 1995] and even books [Omnès, 1994; 1999; Griffiths, 2002] by the founders. See also the reviews by Kiefer [2003] and Halliwell [2004], the critiques by Dowker and Kent [1996], Kent [1998], Bub [1999], and Bassi and Ghirardi [2000], as well as the various mathematical reformulations and reinterpretations of the consistent histories program [Isham, 1994; 1997; Isham and Linden, 1994; 1995; Isham et al., 1994; Isham and Butterfield, 2000; Rudolph, 1996a; 1996b; 2000; Rudolph and Wright, 1999].

The relationship between consistent histories and decoherence is somewhat peculiar: on the one hand, decoherence is a natural mechanism through which appropriate sets of histories become (approximately) consistent, but on the other hand these approaches appear to have quite different points of departure. Namely, where decoherence starts from the idea that (quantum) systems are naturally coupled to their environments and therefore have to be treated as open systems, the aim of consistent histories is to deal with closed quantum systems such as the Universe, without a priori talking about measurements or observers. However, this distinction is merely historical: as we have seen in item 6 in the previous subsection, the dividing line between a system and its environment should be seen as a dynamical entity to be drawn according to certain stability criteria, so that even in decoherence theory one should really study the system plus its environment as a whole from the outset.323 And this is precisely what consistent historians do.

As in the preceding subsection, and for exactly the same reasons, we format our treatment of consistent histories as a list of items open to discussion.

1.

The starting point of the consistent histories formulation of quantum theory is conventional: one has a Hilbert space H, a state ρ, taken to be the initial state of the total system under consideration (realized as a density matrix on H) and a Hamiltonian H (defined as a self-adjoint operator on H). What is unconventional is that this total system may well be the entire Universe. Each property α of the total system is mathematically represented by a projection Pα on H; for example, if α is the property that the energy takes some value ε, then the operator Pα is the projection onto the associated eigenspace (assuming ε belongs to the discrete spectrum of H). In the Heisenberg picture, Pα evolves in time as Pα(t) according to (12); note that Pα(t) is once again a projection.

A historyA is a chain of properties (or propositions) (α1(t1),…, αn(tn)) indexed by n different times t1 < … < tn; here A is a multi-label incorporating both the properties (α1,…, αn) and the times (t1,…, tn). Such a history indicates that each property αi holds at time ti, i = 1,…, n. Such a history may be taken to be a collection {α(t)}t∈ defined for all times, but for simplicity one usually assumes that α(t) ≠ 1 (where 1 is the trivial property that always holds) only for a finite set of times t; this set is precisely {t1,…, tn}. An example suggested by Heisenberg (1927) is to take αi to be the property that a particle moving through a Wilson cloud chamber may be found in a cell Δi ⊂ ℝ6 of its phase space; the history (α1(t1),…, αn(tn)) then denotes the state of affairs in which the particle is in cell Δ1 at time t1, subsequently is in cell δ2 at time t2, etcetera. Nothing is stated about the particle's behaviour at intermediate times. Another example of a history is provided by the double slit experiment, where α1 is the particle's launch at the source at t1 (which is usually omitted from the description), α2 is the particle passing through (e.g.) the upper slit at t2, and α3 is the detection of the particle at some location L at the screen at t3. As we all know, there is a potential problem with this history, which will be clarified below in the present framework.

The fundamental claim of the consistent historians seems to be that quantum theory should do no more (or less) than making predictions about the probabilities that histories occur. What these probabilities actually mean remains obscure (except perhaps when they are close to zero or one, or when reference is made to some measurement context; see [Hartle, 2005]), but let us first see when and how one can define them. The only potentially meaningful mathematical expression (within quantum mechanics) for the probability of a history ℍA with respect to a state ρ is [Groenewold, 1952; Wigner, 1963]

(5)p(A)=Tr(CAρCA*),
where
(6)CA=Pαn(tn)Pα1(t1)

Note that CA is generally not a projection (and hence a property) itself (unless all Pαi mutually commute). In particular, when ρ = [Ψ] is a pure state (defined by some unit vector Ψ ∈ H), one simply has

(7)p(A)=CAΨ2=Pαn(tn)Pα1(t1)Ψ2

When n = 1 this just yields the Born rule. Conversely, see Isham (1994) for a derivation of (5) from the Born rule.324

2.

Whatever one might think about the metaphysics of quantum mechanics, a probability makes no sense whatsoever when it is only attributed to a single history (except when it is exactly zero or one). The least one should have is something like a sample space (or event space) of histories, each (measurable) subset of which is assigned some probability such that the usual (Kolmogorov) rules are satisfied. This is a (well-known) problem even for a single time t and a single projection Pα (i.e. n = 1). In that case, the problem is solved by finding a self-adjoint operator A of which Pα is a spectral projection, so that the sample space is taken to be the spectrum σ(A) of A, with α ⊂ σ(A). Given Pα, the choice of A is by no means unique, of course; different choices may lead to different and incompatible sample spaces. In practice, one usually starts from A and derives the Pα as its spectral projections Pα = ∫α dP (λ), given that the spectral resolution of A is A = ∫ dP (λ) λ. Subsequently, one may then either coarse-grain or fine-grain this sample space. The former is done by finding a partition σ(A) = ∐i αi (disjoint union), and only admitting elements of the σ-algebra generated by the αi as events (along with the associated spectral projection Pαi), instead of all (measurable) subsets of σ(A). To perform fine-graining, one supplements A by operators that commute with A as well as with each other, so that the new sample space is the joint spectrum of the ensuing family of mutually commuting operators.

In any case, in what follows it turns out to be convenient to work with the projections Pα instead of the subsets α of the sample space; the above discussion then amounts to extending the given projection on H to some Boolean sublattice of the lattice P(H) of all projections on H.325 Any state ρ then defines a probability measure on this sublattice in the usual way [Beltrametti and Cassinelli, 1984].

3.

Generalizing this to the multi-time case is not a trivial task, somewhat facilitated by the following device (Isham, 1994). Put HN = ⊗NH, where N is the cardinality of the set of all times ti relevant to the histories in the given collection,326 and, for a given history ℍA, define

(8)A=Pαn(tn)Pα1(t1)

Here Pαi (ti) acts on the copy of H in the tensor product HN labeled by ti, so to speak. Note that ℂA is a projection on HN (whereas CA in (6) is generally not a projection on H). Furthermore, given a density matrix ρ on H as above, define the decoherence functional d as a map from pairs of histories into ℂ by

(9)d(A,B)=Tr(CAρCB*)

The main point of the consistent histories approach may now be summarized as follows: a collection {ℍA}A∈A of histories can be regarded as a sample space on which a state ρ defines a probability measure via (5), which of course amounts to

(10)p(A)=d(A,A),
provided that:
(a)

The operators {ℂA}A∈A form a Boolean sublattice of the lattice P(HN) of all projections on HN;

(b)

The real part of d (ℍA, ℍB) vanishes whenever ℍA is disjoint from ℍB.327

In that case, the set {ℍ}A∈A is called consistent. It is important to realize that the possible consistency of a given set of histories depends (trivially) not only on this set, but in addition on the dynamics and on the initial state.

Consistent sets of histories generalize families of commuting projections at a single time. There is no great loss in replacing the second condition by the vanishing of d (ℍA, ℍB) itself, in which case the histories ℍA and ℍB are said to decohere.328 For example, in the double slit experiment the pair of histories {ℍA, ℍB} where α1 = β1 is the particle's launch at the source at t1, α22) is the particle passing through the upper (lower) slit at t2, and α3 = β3 is the detection of the particle at some location L at the screen, is not consistent. It becomes consistent, however, when the particle's passage through either one of the slits is recorded (or measured) without the recording device being included in the histories (if it is, nothing would be gained). This is reminiscent of the von Neumann chain in quantum measurement theory, which indeed provides an abstract setting for decoherence (cf. item 1 in the preceding subsection). Alternatively, the set can be made consistent by omitting α2 and β2. See [Griffiths, 2002] for a more extensive discussion of the double slit experiment in the language of consistent histories.

More generally, coarse-graining by simply leaving out certain properties is often a promising attempt to make a given inconsistent set consistent; if the original history was already consistent, it can never become inconsistent by doing so. Fine-graining (by embedding into a larger set), on the other hand, is a dangerous act in that it may render a consistent set inconsistent.

4.

What does it all mean? Each choice of a consistent set defines a “universe of discourse” within which one can apply classical probability theory and classical logic [Omnès, 1992]. In this sense the consistent historians are quite faithful to the Copenhagen spirit (as most of them acknowledge): in order to understand it, the quantum world has to be looked at through classical glasses. In our opinion, no convincing case has ever been made for the absolute necessity of this Bohrian stance (cf. Subsection 3.1), but accepting it, the consistent histories approach is superior to Copenhagen in not relying on measurement as an a priori ingredient in the interpretation of quantum mechanics.329 It is also more powerful than the decoherence approach in turning the notion of a system into a dynamical variable: different consistent sets describe different systems (and hence different environments, defined as the rest of the Universe); cf. item 6 in the previous subsection.330 In other words, the choice of a consistent set boils down to a choice of “relevant variables” against “irrelevant” ones omitted from the description. As indeed stressed in the literature, the act of identification of a certain consistent set as a universe of discourse is itself nothing but a coarse-graining of the Universe as a whole.

5.

But these conceptual successes come with a price tag. Firstly, consistent sets turn out not to exist in realistic models (at least if the histories in the set carry more than one time variable). This has been recognized from the beginning of the program, the response being that one has to deal with approximately consistent sets for which (the real part of) d (ℍA, ℍB) is merely very small. Furthermore, even the definition of a history often cannot be given in terms of projections. For example, in Heisenberg's cloud chamber example (see item 1 above), because of his very own uncertainty principle it is impossible to write down the corresponding projections Pαi. A natural candidate would be Pα=QB(χΔ), cf. (19) and (28), but in view of (21) this operator fails to satisfy P2α = Pα, so that it is not a projection (although it does satisfy the second defining property of a projection P*α = Pα). This merely reflects the usual property Q(f)2Q(f2) of any quantization method, and necessitates the use of approximate projections [Omnès, 1997]. Indeed, this point calls for a reformulation of the entire consistent histories approach in terms of positive operators instead of projections [Rudolph, 1996a, b].

These are probably not serious problems; indeed, the recognition that classicality emerges from quantum theory only in an approximate sense (conceptually as well as mathematically) is a profound one (see the Introduction), and it rather should be counted among its blessings that the consistent histories program has so far confirmed it.

6.

What is potentially more troubling is that consistency by no means implies classicality beyond the ability (within a given consistent set) to assign classical probabilities and to use classical logic. Quite to the contrary, neither Schrödinger cat states nor histories that look classical at each time but follow utterly unclassical trajectories in time are forbidden by the consistency conditions alone [Dowker and Kent, 1996]. But is this a genuine problem, except to those who still believe that the earth is at the centre of the Universe and/or that humans are privileged observers? It just seems to be the case that — at least according to the consistent historians — the ontological landscape laid out by quantum theory is far more “inhuman” (or some would say “obscure”) than the one we inherited from Bohr, in the sense that most consistent sets bear no obvious relationship to the world that we observe. In attempting to make sense of these, no appeal to “complementarity” will do now: for one, the complementary pictures of the quantum world called for by Bohr were classical in a much stronger sense than generic consistent sets are, and on top of that Bohr asked us to only think about two such pictures, as opposed to the innumerable consistent sets offered to us. Our conclusion is that, much as decoherence does not solve the measurement problem but rather aggravates it (see item 2 in the preceding subsection), also consistent histories actually make the problem of interpreting quantum mechanics more difficult than it was thought to be before. In any case, it is beyond doubt that the consistent historians have significantly deepened our understanding of quantum theory — at the very least by providing a good bookkeeping device!

7.

Considerable progress has been made in the task of identifying at least some (approximately) consistent sets that display (approximate) classical behaviour in the full sense of the word [Gell-Mann and Hartle, 1993; Omnès, 1992; 1997; Halliwell, 1998; 2000; 2004; Brun and Hartle, 1999; Bosse and Hartle, 2005]. Indeed, in our opinion studies of this type form the main concrete outcome of the consistent histories program. The idea is to find a consistent set {ℍ}A∈A with three decisive properties:

(a)

Its elements (i.e. histories) are strings of propositions with a classical interpretation;

(b)

Any history in the set that delineates a classical trajectory (i.e. a solution of appropriate classical equations of motion) has probability (10) close to unity, and any history following a classically impossible trajectory has probability close to zero;

(c)

The description is sufficiently coarse-grained to achieve consistency, but is sufficiently fine-grained to turn the deterministic equations of motion following from (b) into a closed system.

When these goals are met, it is in this sense (no more, no less) that the consistent histories program can claim with some justification that it has indicated (or even explained) ‘How the quantum Universe becomes classical’ [Halliwell, 2005].

Examples of propositions with a classical interpretation are quantized classical observables with a recognizable interpretation (such as the operators Bħ (χδ) mentioned in item 5), macroscopic observables of the kind studied in Subsection 6.1, and hydrodynamic variables (i.e. spatial integrals over conserved currents). These represent three different levels of classicality, which in principle are connected through mutual fine- or coarse-grainings.331 The first are sufficiently coarse-grained to achieve consistency only in the limit ℏ → 0 (cf. Section 5), whereas the latter two are already coarse-grained by their very nature. Even so, also the initial state will have to be “classical” in some sense in order te achieve the three targets (a) – (c).

All this is quite impressive, but we would like to state our opinion that neither decoherence nor consistent histories can stand on their own in explaining the appearance of the classical world. Promising as these approaches are, they have to be combined at least with limiting techniques of the type described in Sections 5 and 6 — not to speak of the need for a new metaphysics! For even if it is granted that decoherence yields the disappearance of superpositions of Schrödinger cat type, or that consistent historians give us consistent sets none of whose elements contain such superpositions among their properties, this by no means suffices to explain the emergence of classical phase spaces and flows thereon determined by classical equations of motion. Since so far the approaches cited in Sections 5 and 6 have hardly been combined with the decoherence and/or the consistent histories program, a full explanation of the classical world from quantum theory is still in its infancy. This is not merely true at the technical level, but also conceptually; what has been done so far only represents a modest beginning. On the positive side, here lies an attractive challenge for mathematically minded researchers in the foundations of physics!

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780444515605500087

I

It would carry us too far to recall in detail how, with the elimination of mythical cosmological ideas and arguments referring to purposes of our own actions, a consistent scheme of mechanics was built up on the basis of Galileo’s pioneering work, and how it reached completion through Newton’s mastership. Above all, the principles of Newtonian mechanics meant a far-reaching clarification of the problem of cause and effect in permitting, through observation of the state of a physical system at a given instant, defined by measurable quantities, prediction of its state at any subsequent time. It is well known how a deterministic or causal account of this kind led to the mechanical conception of nature and came to stand as an ideal of scientific explanation in all domains of knowledge, irrespective of the way experience is obtained. It is therefore important to recall that the study of wider fields of physical experience has revealed the necessity of a closer consideration of the observational problem.

Within its large field of application, classical mechanics presents an objective description in the sense that it is based on a well-defined use of pictures and ideas referring to events in daily life. Still, however rational the idealizations used in Newtonian mechanics might appear, they actually went far beyond the range of experience to which our elementary concepts are adapted. Thus, the adequate use of the very notions of absolute space and time is inherently connected with the practically instantaneous propagation of light, which allows us to locate the bodies around us independently of their velocities and to arrange events in a unique time sequence. The attempt at developing a consistent account of electromagnetic and optical phenomena revealed, however, that observers moving relative to each other with great velocities will coordinate events differently. Such observers may, in fact, not only take a different view of shapes and positions of rigid bodies, but events at separate points of space, which to one observer appear as simultaneous, may to another be judged as occurring at different times.

Far from giving rise to confusion and complication, the exploration of the extent to which the account of physical phenomena depends on the standpoint of the observer proved an invaluable guide in tracing general physical laws common to all observers. Retaining the idea of determinism, but relying only on relations between unambiguous measurements referring ultimately to coincidences of events, Albert Einstein succeeded in remolding and generalizing the whole edifice of classical physics, and in lending to our world picture a unity surpassing all previous expectations. In the general theory of relativity, the description is based on a curved four-dimensional space-time metric which automatically accounts for gravitational effects and the singular role of the speed of light signals representing an upper limit for any consistent use of the physical concept of velocity. The introduction of such unfamiliar but well-defined mathematical abstractions in no way implies any ambiguity. It offers, rather, an instructive illustration of how a widening of the conceptual framework affords the appropriate means of eliminating subjective elements and enlarging the scope of objective description.

New unsuspected aspects of the observational problem should be disclosed by the exploration of the atomic constitution of matter. The ideas of a limited divisibility of substances, introduced to explain the persistence of their characteristic properties in spite of the variety of natural phenomena, go back to antiquity. Still, almost to our day, such views were regarded as essentially hypothetical in the sense that they seemed inaccessible to direct confirmation by observation, because of the coarseness of our sense organs and of the tools we use, which themselves are composed of innumerable atoms. Nevertheless, with the great progress in chemistry and physics in the last centuries, atomic ideas proved increasingly fruitful, and it was found possible to obtain a general understanding of the principles of thermodynamics through a direct application of classical mechanics to the interaction of atoms and molecules in incessant motion.

In this century, the study of newly discovered properties of matter, like natural radioactivity, has convincingly confirmed the fundaments of atomic theory. In particular, through the development of amplification devices, it has been possible to study phenomena essentially dependent on single atoms, and even to obtain extensive knowledge of the structure of atomic systems. The first step was the recognition of the electron as a common constituent of all substances, and an essential completion of our ideas of atomic constitution was obtained by Ernest Rutherford’s discovery of the atomic nucleus containing almost the whole mass of the atom within an extremely small volume. As a result of this, the invariability of the elements in ordinary physical and chemical processes is directly explained by the circumstance that in such processes, although the electron binding may be largely influenced, the nucleus remains unaltered. With his demonstration of the transmutability of atomic nuclei by more powerful agencies, Rutherford, moreover, opened a new field of research, often referred to as modern alchemy, which eventually should lead to the possibility of releasing immense amounts of energy stored in atomic nuclei.

Although many fundamental properties of matter were explained by this simple picture of the atom, it was from the beginning evident that classical ideas of mechanics and electromagnetism did not suffice to account for the essential stability of atomic structures exhibited by the specific properties of the elements. A clue to the elucidation of this problem was afforded, however, by the discovery of the universal quantum of action to which Max Planck was led in the first year of our century by his penetrating analysis of the laws of thermal radiation. This discovery revealed a feature of wholeness in atomic processes quite foreign to the mechanical conception of nature, and made it evident that the classical physical theories are idealizations valid only in the description of phenomena, in the analysis of which all actions are sufficiently large to permit the neglect of the quantum. While this condition is amply fulfilled in phenomena on the ordinary scale, we meet in atomic phenomena regularities of quite a new kind, defying deterministic pictorial description.

A rational generalization of classical physics, allowing for the existence of the quantum but retaining the unambiguous interpretation of experimental evidence defining inertial mass and electric charge of the electron and the nucleus, presented a very difficult task. By concerted efforts of a whole generation of theoretical physicists, a consistent and, within a wide scope, exhaustive description of atomic phenomena, however, was gradually developed. This description makes use of a mathematical formalism in which the variables in the classical physical theories are replaced by symbols subject to a non-commutable algorism involving Planck’s constant. Due to the very character of such mathematical abstractions, the formalism does not allow for pictorial interpretation on accustomed lines, but aims directly at establishing relations between observations obtained under well-defined conditions. Corresponding to the circumstance that different individual quantum processes may take place in a given experimental arrangement, these relations are of an inherently statistical character.

By means of the quantum mechanical formalism, a detailed account of an immense amount of experimental evidence regarding the physical and chemical properties of matter has been achieved. Moreover, by adapting the formalism to the exigencies of relativistic invariance, it has been possible to a large extent to order the rapidly growing new experience concerning the properties of elementary particles and the constitution of atomic nuclei. Notwithstanding, the astounding power of quantum mechanics, the radical departure from accustomed physical explanation, and especially the renunciation of the very idea of determinism, have given rise to doubts in the minds of many physicists and philosophers as to whether we are here dealing with a temporary procedure of expediency or whether we are confronted with an irrevocable step as regards objective description. The clarification of this problem has actually demanded a radical revision of the foundation for the description and comprehension of physical experience.

Therefore we must above all recognize that even when the phenomena transcend the scope of classical physical theories, the account of the experimental arrangement and the recording of the observations must be given in plain language, suitably supplemented by technical physical terminology. This is a clear logical demand since the very word experiment refers to a situation where we can tell others what we have done and what we have learned. The fundamental difference between the analysis of the phenomena in classical and in quantum physics is that in the former the interaction between the objects and the measuring instruments may be neglected or compensated for, while in the latter this interaction forms an integral part of the phenomena. The essential wholeness of a proper quantum phenomenon finds logical expression in the circumstance that any attempt at its well-defined subdivision would require a change in the experimental arrangement incompatible with the appearance of the phenomenon itself.

The impossibility of a separate control of the interaction between the atomic objects and the instruments indispensable for the definition of the experimental conditions prevents in particular the unrestricted combination of space-time coordination and dynamical conservation laws on which the deterministic description in classical physics rests. In fact, any unambiguous use of the concepts of space and time refers to an experimental arrangement involving a transfer of momentum and energy, uncontrollable in principle, to the instruments—like fixed scales and synchronized clocks—required for the definition of the reference frame. Conversely, the account of phenomena governed by conservation of momentum and energy involves, in principle, a renunciation of detailed space-time coordination. These circumstances find quantitative expression in Heisenberg’s indeterminacy relations which specify the reciprocal latitude for the fixation of kinematical and dynamical variables in the definition of the state of a physical system, In accordance with the character of the quantum mechanical formalism such relations cannot be interpreted, however, in terms of attributes of objects referring to classical pictures. We are here dealing with the mutually exclusive conditions for the unambiguous use of the very concepts of space and time, on the one hand, and of dynamical conservation laws, on the other.

In this context, one sometimes speaks of the “disturbance of phenomena by observation” or the “creation of physical attributes to atomic objects by measurements.” Such phrases, however, are apt to cause confusion, since words like phenomena and observation, and attributes and measurements, are here used in a way incompatible with common language and practical definition. In objective description, it is indeed more appropriate to use the word phenomenon only to refer to observations obtained under specified circumstances, including an account of the whole experimental arrangement. In this terminology, the observational problem in quantum physics is deprived of any special intricacy and we are directly reminded that every well-defined atomic phenomenon is closed in itself, since its observation implies a permanent mark on a photographic plate left by the impact of an electron or similar recordings obtained by suitable amplification devices of essentially irreversible functioning. Moreover, it is important to realize that the quantum mechanical formalism allows of well-defined applications only when referring to phenomena closed in such sense, and also that in this respect it represents a rational generalization of classical physics in which every stage of the course of events is described by measurable quantities.

The freedom of experimentation presupposed in classical physics of course is retained and corresponds to the free choice of experimental arrangements for which the mathematical structure of the quantum mechanical formalism offers the appropriate latitude. The circumstance that, in general, one and the same experimental arrangement may yield different recordings, is sometimes picturesquely described as a “choice of nature” between such possibilities. Needless to say, one is here not in any way alluding to a personification of nature, but rather pointing to the logical impossibility of ascertaining directives on accustomed lines for the course of a closed indivisible phenomenon. As regards such directives, logical approach cannot go beyond the deduction of the relative probabilities for the appearance of the individual phenomena under given experimental conditions, and in this respect quantum mechanics fulfills all requirements as a consistent generalization of deterministic mechanical description which it embraces as an asymptotic limit in the case of physical phenomena on a scale sufficiently large to neglect the quantum of action.

A most conspicuous characteristic of atomic physics is the novel relationship between phenomena observed under experimental conditions demanding different elementary concepts for their description. Indeed, however contrasting such phenomena might appear when attempting to picture a course of atomic processes on classical lines, they have to be considered as complementary in the sense that they represent equally essential aspects of well-defined knowledge about atomic systems and together exhaust this knowledge. Far from indicating a departure from our position as detached observed, the notion of complementarity represents the logical expression for our situation as regards objective description in this field of experience, which has demanded a renewed revision of the foundation for the unambiguous use of our elementary concepts. The recognition that the interaction between the measuring tools and the physical systems under investigation constitute an integral part of quantum phenomena has not only revealed an unsuspected limitation of the mechanical conception of nature characterized by attribution of separate properties to physical systems, but has forced us in the ordering of experience to pay proper attention to the conditions of observation.

Returning to the much debated question of what has to be demanded of a physical explanation, it must be borne in mind that classical mechanics already implied the renunciation of cause of uniform motion and especially that relativity theory has taught us how arguments of invariance and equivalence must be counted as categories of rational explanation. Similarly, in the complementary description of quantum physics, we have to do with a further self-consistent generalization which permits us to include regularities decisive for the account of fundamental properties of matter but transcending the scope of deterministic description. The history of physical science thus demonstrates how the exploration of ever wider fields of experience, in revealing unsuspected limitations of accustomed ideas, indicates new ways of restoring logical order. The epistemological lesson contained in the development of atomic physics reminds us of similar situations in description and comprehension of experience far beyond the borders of physical science, and allows us to trace common features promoting the search for unity of knowledge.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/S1876050308702007

Quantum world

Classical physics and theory of relativity developing simultaneously with an understanding of the world of atoms and molecules, elementary particles, and fundamental interactions marked the beginning of a new era — the era of the quantum world.

Why are some substances dense and strong while others are loose and brittle? What is the cause of the spontaneous fission of uranium nuclei? Why is copper a conductor and why is glass an insulator? What is the nature of magnetite and magnetism? How does the sun work and why can solar radiation energy be converted into electricity? What are methane clathrates? What is the essence of the problem of global warming? Why are the spectra of various substances distinguishable and, moreover, discrete? For these and many other questions, classical physics fails to yield quantitative answers. To reply to the queries about the world of atoms and molecules is the mission of quantum mechanics.

Quantum mechanics is a complicated science. Many and various substances surround us; for us they are complicated and unique, but at the same time there is not great difficulty classifying them in terms of chemical elements. Water is H2O, methane is CH4, carbon dioxide is CO2, and so on. Each atom represents an element of Mendeleev’s table, and the table is the fruit of quantum mechanics. Elementary particles, in a sense forming the tangled “invisible” world of reactions and mutual transformations, are hiding inside the atoms. For instance, in a collision of two protons one can see the birth of the meson of π-type or K-mesons may be created, but protons disappear at the same time. Moreover, there exists a deep relation between electromagnetic radiation and particles, so the reaction of a collision of an electron and a positron can lead to the creation of γ-quanta and annihilation of the electron–positron couple. Naturally, the reverse process is also possible when a photon having sufficient energy initiates the appearance of particles. A series of these and many other phenomena from the microworld is described in the specific language of quantum mechanics.

Paying tribute to the history of the development of quantum theory, let us briefly consider a few problems in which the argument of classical physics experiences the greatest difficulties.

We begin with the problem of black body radiation. Let there be a cavity, and its hole represents a perfect black surface. A light, penetrating inside, is repeatedly scattered and absorbed. If thermal energy, which is emitted by the walls of the cavity inside, equals the energy that is absorbed by these same walls, then thermal equilibrium is maintained. How does the energy density u in the cavity depend on wavelength λ and temperature T? Planck replied to this question. Recall, according to Wien’s displacement law,

λmaxckBT=C=const,

in which kB is the Boltzmann constant, c is the speed of light, λmax corresponds to the maximum of function u, and constant C has a specific dimension,

time×energy.

Planck established the general law for the energy per unit volume and per unit of wavelength interval, and he calculated constant C; as a result,

u(λ,T)=8πhcλ5·1ehc/λkBT1

and C=0.2h, and

h=6.63×1027erg·s

is the fundamental Planck constant. With the birth of h, quantum physics was born.

Another controversial problem is the photoelectric effect. Quite surprisingly, under the action of light, according to experiments by Millikan, one might eject the electrons from the metal surface. The emission of electrons is understandable, but the fact that the electron energy depends linearly on frequency ν of incident radiation is much more complicated to understand, especially from the classical point of view. Einstein resolved this conundrum and offered an elegant explanation of the photoelectric effect,

Ee=hνW,

in which Ee is the electron energy and W is the work function. The intrigue of the wave–particle duality of light was thus revived.

Finally, we discuss a question related to the stability of the lifetime of atoms. According to Rutherford’s experiments, most α-particles pass practically without hindrance through a foil of gold, deflected by nuclei at definite angles, which one might easily calculate by applying classical considerations. These experiments prove, firstly, that an atom is a nucleus plus electrons and, secondly, that Coulomb’s law is valid at atomic distances. From Rutherford’s experiments, it is impossible to evaluate the size of atom. Moreover, the planetary model is inconsistent with classical electrodynamics; otherwise, an electron moving with acceleration around the nucleus would be forced to radiate at each revolution, losing its energy, and eventually would fall into the nucleus after only one-hundredth of a nanosecond. A lifetime would not be long.

The explanation of the “paradox” of stability of atoms came from spectroscopy. Analyzing the atomic spectra, Ritz stated that the distribution of frequencies ν occurs in accordance with the combination principle

νsn=νsνn=Ry(n2s2),

in which Ry is the Rydberg constant and s and n are integers other than zero. According to the classical description, the spectral lines must be equidistant from each other. In fact, the lines converge with increasing s, which is confirmed, for instance, by the Balmer series for hydrogen atoms. Bohr, understanding that hν represents the energy itself, proposed a quantum model of an atom, according to which the atom, being in a stationary state, does not emit or absorb radiation, and transitions possible from one state to another occur in compliance with the law of conservation of energy according to a rule

hνsn=EsEn.

If an atom, for example, emits a photon, then an electron jumps from some stationary state with energy Es to the state that is characterized by energy En; in this case, the frequency of the photon equals νsn=(EsEn)/h. The case of absorption is interpreted analogously.

Probability waves

To a particle of the microworld, de Broglie ascribed wavefunction φ. Realizing this determination in a literal sense, de Broglie introduced the relations

E=ωandp=k,

which connect energy E and momentum p of a freely moving particle with frequency ω and wave vector k of a plane wave

φ~ei(ωtk·q),

in which ω=2πν=c|k|and=h/2π. In other words, apart from the material, there is “something” that at each point q of space provides some information about the particle at the moment of time t. Born proposed the probability interpretation for function φ. These waves are probability waves, for instance, the waves of probability where the particle is located at a concrete point of space or where the particle has a concrete value of energy. Despite his phenomenological explanation of the photoelectric effect, Einstein contested vigorously against the probability interpretation of a wavefunction: “God does not play dice.” Despite all its successes, Einstein, as is well known, treated quantum mechanics cautiously.

Depending on a representation, a wavefunction can be specified as the function of corresponding variables; for instance,

φ(q),χ(p),andθ(E)

are images of a wavefunction in coordinate, momentum, and energy representations, respectively. To not indicate explicitly the chosen representation, it is convenient to express the states of quantum-mechanical systems through abstract vectors in a separable Hilbert space. According to Dirac’s notation, to a vector that characterizes some state n we ascribe symbol |n; furthermore,

n|

is the vector that is the complex conjugate of |n,

c·|n

is the multiplication of vector by a complex number c,

n|m

is the scalar product of two vectors |n and |m, and

m|G|n

is the matrix element of physical quantity G between states n and m. Vectors |n and n| are called ket and bra, respectively.

Any physical vector |Ψ is expressible as a series expansion in terms of orthonormal states |1,|2,,|,, which in a set form a complete basis, and in a linear manner

|Ψ=c1|1+c2|2++c|+,

in which c1,c2,,c, are the pertinent coefficients; basis vectors | fail to be related to each other through any relation of linear type — they are linearly independent. The orthonormality condition means that

k|=δk,andk=1,2,,

that is, the vectors are mutually orthogonal and each of them is normalized to unity. From a physical point of view, the state of a quantum-mechanical system is a superposition of all possible for this system state, and each of which gives its own contribution with a weight that is defined by corresponding coefficient c. Amplitude c thus has a purely probability character, and quantity

|c|2

can be interpreted as the probability of state . Assuming that ket |Ψ is normalized to unity, one might readily obtain this equality,

|c|2=1,

which proves that the total probability, as one should expect, equals unity; the event that the system occupies one available state, obviously, represents a certain event.

In common use, the coordinate representation is applied to the problems of quantum mechanics. In this case, the scalar product of wavefunctions is given by this expression

Mφn*(q)φm(q)dq,

in which M is the manifold, in which functions φn and φm, corresponding to states n and m, respectively, are defined and dq is an element of volume. Quantity

|φn(q)|2dq,

according to Born’s probability interpretation, represents the probability that the values of coordinates of a system lie in an interval between q and q+dq. Thus, diagonal matrix element n|G|n, which is equal to

Mφn*(q)(Gφn(q))dq,

should be understood as the average value of physical quantity G in state n; if G is simply the coordinate function, then

n|G|n=MG(q)|φn(q)|2dq.

The general formula for an arbitrary matrix element in the coordinate representation, by definition, has the form

n|G|m=Mφn*(q)(Gφm(q))dq.

Analogous considerations are applicable for the case of the momentum representation. For instance,

|χn(p)|2dp

is the probability to detect the momentum of a system in the vicinity of point p in the element of volume dp of the momentum space, and

Mχn*(p)(Gχn(p))dp

represents the average value of quantity G in the state with wavefunction χn that is defined in manifold M.

Physical operators

Suppose that q and p are the canonically conjugate coordinate and momentum, φ(q) is the wavefunction in the coordinate representation, and χ(p) is the same function, but in the p-representation; quantities φ(q) and χ(p) are related between each other through the Fourier integral

χ(p)=12π+φ(q)ei(pq/)dq.

The average value of momentum p, on the one side, is expressible as

χ|p|χ=+χ*(p)pχ(p)dp,

and on the other side,

φ|p|φ=+φ*(q)(?)φ(q)dq,

through which symbol “?” designates the operator of momentum in the coordinate representation. Our task,1 taking into account that

χ|p|χ=φ|p|φ

is to determine the explicit form of “?.”

So, using the expansion of χ(p) in the Fourier integral, we have

χ|p|χ=+dp2π+φ*(q)ei(pq/)dq+φ(q)pei(pq/)dq.

Putting

pei(pq/)=iqei(pq/)

and taking into account the boundary conditions φ(±)=0, we execute the integration by parts:

+φ(q)(i/q)ei(pq/)dq=+ei(pq/)(i/q)φ(q)dq;

consequently,

χ|p|χ=+φ*(q)dq+(i/q)φ(q)dq+ei(p(qq)/)dp2π.

As

+ei(p(qq)/)dp2π=δ(qq),

then

χ|p|χ=+(i/q)φ(q)dq+φ*(q)δ(qq)dq=+φ*(q)(i/q)φ(q)dq.

Hence,

χ|p|χ=+χ*(p)pχ(p)dp=+φ*(q)(i/q)φ(q)dq=φ|p|φ

and, in the coordinate representation, the momentum is expressible as a linear differential operator:

p=iq.

With respect to coordinate, the analogous considerations are valid. In this case, we apply the inverse Fourier transformation

φ(q)=12π+χ(p)ei(pq/)dp,

p and q are simply interchanged and i is converted into i. As a result,

q=ip

represents the expression for the operator of coordinate in the momentum representation.

Having determined a commutator for two quantities G and F as follows,

[G,F]=GFFG,

we see that

[q,i/q]φ(q)=iφ(q)and[i/p,p]χ(p)=iχ(p);

hence,

qppq=i,

and the conjugate coordinate and momentum in quantum mechanics fail to conform to the law of commutative multiplication. This condition constitutes properly the main distinction between quantum and classical theories. Classical physics operates with commuting to each other c-numbers, which are generally measurable in the experiment. The natural variables of quantum physics are opposite, unobservable q-numbers. The quantities of q-type, failing to commute generally with each other, act in a Hilbert space of abstract state vectors. The vectors are supposed to be transformed to each other through the action of q-numbers, which are in a sense the pertinent operators. Not all mathematical operators find their application in practice in physics. The principal requirements are linearity and hermitivity. Let

|nand|m

be arbitrary vectors, and let

|Ψ=b|n+c|m,

in which b and c are numerical coefficients. By definition, G is the linear operator if

G|Ψ=bG|n+cG|m;

G is the Hermitian operator if

G=G+,

in which G+ satisfies the relation

n|G|m=m|G+|n*

and represents an operator Hermitian conjugate to G. For instance, as is easily seen in the coordinate representation, the differential operator for momentum possesses the properties of both hermitivity and linearity.

However, there exist such states in which the dynamic variables of quantum origin have determinate values belonging to a class of observable c-numbers. We imply here the eigenvalues of physical operators. So, if q-number G acts on state vector |n without altering its “direction,” an equation

G|n=gn|n

is valid and gn is a number of c-type, then |n and gn are, respectively, eigenvector and eigenvalue of operator G. In a general case, G has a set of eigenvectors, each of which corresponds to a determinate eigenvalue. We show that the eigenvalues of physical operators are strictly real numbers. So, on the one side,

n|G|n=gn;

on the other side, taking into account that G=G+, we have

n|G|n=n|G+|n*=n|G|n*;

consequently,

gn=gn*

completely proves our assertion.

The eigenvectors that belong to various eigenvalues are mutually orthogonal. Let |n and |m be eigenvectors of physical operator G; they correspond to the equations for eigenvalues, respectively, gn and gm:

G|n=gn|nandG|m=gm|m.

As m|G+=gm*m|, G+=G, and gm*=gm, then

m|G|n=gnm|nandm|G|n=gmm|n.

On comparing, we find

(gngm)m|n=0,

hence

m|n=0,

because gngm. The property of mutual orthogonality of eigenvectors plays an important role in the problems of quantum mechanics. Arbitrary vector |Ψ, for instance, can be represented in a form of expansion

|Ψ=ncn|n,

in which cn are the amplitudes of states in terms of the complete system of eigenvectors |n of some operator G. If gn are the eigenvalues corresponding to states n, then the average value of G in state Ψ is given by the expression

Ψ|G|Ψ=mncm*cnm|G|n=n|cn|2gn

that, in probability theory, represents a formula for an expectation value of quantity G.

Finally, we discuss a useful result regarding the possibility of simultaneous measurement of two dynamical quantities. Suppose that

G|n=gn|n,

in which |n and gn represent, respectively, the eigenvectors and eigenvalues of physical operator G. On acting on this equation with some variable F on the left, we have

FG|n=gnF|n.

If [G,F]=0, then

G(F|n)=gn(F|n);

consequently, quantity F|n is the eigenvector of operator G. Hence,

F|n=fn|n,

in which fn are c-numbers. Commutative variables G and F thus have a common complete system of eigenstates. In this case, the dynamical quantities are simultaneously measurable.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780128015889000011

R. J. BOŠKOVIĆ

Niels Bohr, Kobenhavn

Ruder Bošković, whose life-work is receiving greater and greater attention in the scientific world of today, was one of the most prominent figures among the 18th century natural philosophers who enthusiastically elaborated the fundamental conceptions of Newtonian mechanics. Indeed, he did not only make important contributions to mathematics and astronomy, but strove with remarkable imagination and logical power to develop a systematic account of the properties of matter on the basis of interactions of mass points through central forces. In this respect, Bošković’s ideas exerted a deep influence on the work of the next following generation of physicists, resulting in the general mechanistic views which inspired Laplace and, perhaps less directly, even Faraday and Maxwell.

It is true that in our days the approach to such problems has undergone essential changes. Above all, it has been recognized that the consistent description of atomic processes demands a feature of indivisibility, symbolized by the quantum of action and which goes far beyond the old, much debated doctrine of a limited divisibility of matter. This development has revealed an unsuspected limitation of the scope of mechanical pictures and even of the deterministic description of physical phenomena. However, it has been possible, through a most efficient collaboration between physicists from many countries, gradually to develop a rational generalization of the classical theories of mechanics and electrodynamics, which has proved capable of accounting for an ever increasing wealth of experimental data concerning the properties of matter.

When, against this background, one reflects on the development of natural philosophy through the ages, one appreciates the wisdom of the cautious attitude towards atomic problems, which reigned until the last century. I think not only of the belief that, owing to the coarseness of our tools and sense organs, it would never be possible to obtain direct evidence of phenomena on the atomic scale, but also of the often expressed skepticism as to the adequacy of pictorial models in a domain so far removed from ordinary experience. Although the marvellous development of experimental technique has permitted us to record effects of single atomic objects, we are here in a novel situation which has necessitated a radical revision of the fundaments for the unambiguous use of the elementary conceptions, like space and time, and cause and effect, embodied in the language adapted to our orientation in practical life.

The elucidation of the situation with which we are confronted in atomic physics has been obtained by raising anew the old problem of what answers we can receive to questions put to nature in the form of experiments. Of course, no physicist from earliest times has ever thought that he could augment physical knowledge in any other way than by accounting for recordings obtained under well-defined experimental conditions. While, in this respect, there is no change of attitude since Bošković’s time, we have in our days, as is well known, received a new lesson regarding our position as to analysis and synthesis of such knowledge.

Our esteem for the purposefulness of Bošković’s great scientific work, and the inspiration behind it, increases the more as we realize the extent to which it served to pave the way for later developments. In friendly and fruitful international cooperation physicists are working today, in Yugoslavia as in all other countries, for the progress of our knowledge of the atomic constitution of matter and for the application of this knowledge, which holds out promises surpassing even those of the technology based on classical physics. In the pursuit of such novel developments, it is essential that we not only keep an open mind for unforeseen discoveries, but that we are conscious of standing on the foundations laid by the pioneers of our science.

The 200th anniversary of the publication of Bošković’s famous Theoria Philosophiae Naturalis could hardly be commemorated in a more fitting manner than by an international congress in the country of his birth, convened on the occasion of the opening of the museum in Dubrovnik with its historical treasures. In pointing to the future, it is also a most fortunate omen that the great occasion could be combined with the inauguration of the modern research institute in Zagreb, which bears Ruder Bošković’s name and where Meštrović’s impressive statue will daily remind students of the traditions on which they are building and inspire them to fruitful contributions to common human knowledge.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/S1876050308702652

This chapter, being essentially an introduction, slightly opens a door into a tangled labyrinth of quantum mechanics. In a brief review of physical and mathematical principles of the abstract science of amazing phenomena of a microworld, we begin with a discussion of some problems that experience the greatest difficulties when being solved within a frame of classical physics. We explain further the properties of vectors, which describe the states of a system, and the properties of linear operators, which determine the dynamical variables. The main objective of the physical theory is to trace the variation of states and variables in time. For this purpose, we involve Schrödinger and Heisenberg equations that represent two pictures of describing the phenomena in quantum mechanics. As the basic recipes to solve the general equations, we provide the method of factorization, which allows one to solve the problems in a purely algebraic manner, and the perturbation method, which enables one to obtain an approximate solution. We test the method of factorization in an instance of a simple harmonic oscillator. In turn, the power of perturbation theory is demonstrated when considering anharmonic systems. For a description of anharmonic effects, we introduce phenomenologically the formalism of functions of quantum numbers; the meaning of these functions becomes revealed in Chapter 3.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780128015889000011

Abstract:

Two jewels of classical physics are James Clerk Maxwell’s field equations (1861–1862) and the Lorentz force law (Henrick Lorentz, 1892) for electrodynamics. Although this field theory is not exact in some extreme situations, it is bedrock physics that has been used successfully to solve many problems in applied science. A brief introduction to electromagnetism from a mathematical point of view is presented, and the electromagnetic boundary value problem is proved to have unique solutions. Time-harmonic fields are introduced. The limitations of Maxwell’s theory are discussed.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780128041536500208