41
arXiv:quant-ph/0312059v4 28 Jun 2005 Decoherence, the measurement problem, and interpretations of quantum mechanics Maximilian Schlosshauer Department of Physics, University of Washington, Seattle, Washington 98195, USA Environment-induced decoherence and superselection have been a subject of intensive research over the past two decades, yet their implications for the foundational problems of quantum mechanics, most notably the quantum measurement problem, have remained a matter of great controversy. This paper is intended to clarify key features of the decoherence program, including its more recent results, and to investigate their application and consequences in the context of the main interpretive approaches of quantum mechanics. Contents I. Introduction 1 II. The measurement problem 3 A. Quantum measurement scheme 3 B. The problem of definite outcomes 4 1. Superpositions and ensembles 4 2. Superpositions and outcome attribution 4 3. Objective vs. subjective definiteness 5 C. The preferred-basis problem 6 D. The quantum-to-classical transition and decoherence 6 III. The decoherence program 7 A. Resolution into subsystems 8 B. The concept of reduced density matrices 9 C. A modified von Neumann measurement scheme 10 D. Decoherence and local suppression of interference 10 1. General formalism 10 2. An exactly solvable two-state model for decoherence11 E. Environment-induced superselection 12 1. Stability criterion and pointer basis 12 2. Selection of quasiclassical properties 13 3. Implications for the preferred-basis problem 14 4. Pointer basis vs instantaneous Schmidt states 15 F. Envariance, quantum probabilities, and the Born rule 16 1. Environment-assisted invariance 17 2. Deducing the Born rule 17 3. Summary and outlook 18 IV. The role of decoherence in interpretations of quantum mechanics 18 A. General implications of decoherence for interpretations 19 B. The standard and the Copenhagen interpretation 20 1. The problem of definite outcomes 20 2. Observables, measurements, and environment-induced superselection 21 3. The concept of classicality in the Copenhagen interpretation 22 C. Relative-state interpretations 22 1. Everett branches and the preferred-basis problem 23 2. Probabilities in Everett interpretations 24 3. The “existential interpretation” 25 D. Modal interpretations 26 1. Property assignment based on environment-induced superselection 26 2. Property assignment based on instantaneous Schmidt decompositions 27 * Electronic address: [email protected] 3. Property assignment based on decompositions of the decohered density matrix 27 4. Concluding remarks 28 E. Physical collapse theories 28 1. The preferred-basis problem 29 2. Simultaneous presence of decoherence and spontaneous localization 29 3. The tails problem 30 4. Connecting decoherence and collapse models 30 5. Summary and outlook 31 F. Bohmian mechanics 31 1. Particles as fundamental entities 31 2. Bohmian trajectories and decoherence 32 G. Consistent histories interpretations 33 1. Definition of histories 33 2. Probabilities and consistency 33 3. Selection of histories and classicality 34 4. Consistent histories of open systems 35 5. Schmidt states vs pointer basis as projectors 35 6. Exact vs approximate consistency 36 7. Consistency and environment-induced superselection 36 8. Summary and discussion 37 V. Concluding remarks 37 Acknowledgments 38 References 38 I. INTRODUCTION The implications of the decoherence program for the foundations of quantum mechanics have been the subject of an ongoing debate since the first precise formulation of the program in the early 1980s. The key idea promoted by decoherence is the insight that realistic quantum sys- tems are never isolated, but are immersed in the sur- rounding environment and interact continuously with it. The decoherence program then studies, entirely within the standard quantum formalism (i.e., without adding any new elements in the mathematical theory or its in- terpretation), the resulting formation of quantum corre- lations between the states of the system and its envi- ronment and the often surprising effects of these system- environment interactions. In short, decoherence brings

arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

Embed Size (px)

Citation preview

Page 1: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

arX

iv:q

uant

-ph/

0312

059v

4 2

8 Ju

n 20

05

Decoherence, the measurement problem, and interpretations of quantum

mechanics

Maximilian Schlosshauer∗

Department of Physics, University of Washington, Seattle, Washington 98195, USA

Environment-induced decoherence and superselection have been a subject of intensive research overthe past two decades, yet their implications for the foundational problems of quantum mechanics,most notably the quantum measurement problem, have remained a matter of great controversy.This paper is intended to clarify key features of the decoherence program, including its morerecent results, and to investigate their application and consequences in the context of the maininterpretive approaches of quantum mechanics.

Contents

I. Introduction 1

II. The measurement problem 3A. Quantum measurement scheme 3B. The problem of definite outcomes 4

1. Superpositions and ensembles 42. Superpositions and outcome attribution 43. Objective vs. subjective definiteness 5

C. The preferred-basis problem 6D. The quantum-to-classical transition and decoherence 6

III. The decoherence program 7A. Resolution into subsystems 8B. The concept of reduced density matrices 9C. A modified von Neumann measurement scheme 10D. Decoherence and local suppression of interference 10

1. General formalism 102. An exactly solvable two-state model for decoherence11

E. Environment-induced superselection 121. Stability criterion and pointer basis 122. Selection of quasiclassical properties 133. Implications for the preferred-basis problem 144. Pointer basis vs instantaneous Schmidt states 15

F. Envariance, quantum probabilities, and the Born rule 161. Environment-assisted invariance 172. Deducing the Born rule 173. Summary and outlook 18

IV. The role of decoherence in interpretations of

quantum mechanics 18A. General implications of decoherence for interpretations 19B. The standard and the Copenhagen interpretation 20

1. The problem of definite outcomes 202. Observables, measurements, and

environment-induced superselection 213. The concept of classicality in the Copenhagen

interpretation 22C. Relative-state interpretations 22

1. Everett branches and the preferred-basis problem 232. Probabilities in Everett interpretations 243. The “existential interpretation” 25

D. Modal interpretations 261. Property assignment based on environment-induced

superselection 262. Property assignment based on instantaneous

Schmidt decompositions 27

∗Electronic address: [email protected]

3. Property assignment based on decompositions ofthe decohered density matrix 27

4. Concluding remarks 28

E. Physical collapse theories 28

1. The preferred-basis problem 29

2. Simultaneous presence of decoherence andspontaneous localization 29

3. The tails problem 30

4. Connecting decoherence and collapse models 30

5. Summary and outlook 31

F. Bohmian mechanics 31

1. Particles as fundamental entities 31

2. Bohmian trajectories and decoherence 32

G. Consistent histories interpretations 33

1. Definition of histories 33

2. Probabilities and consistency 33

3. Selection of histories and classicality 34

4. Consistent histories of open systems 35

5. Schmidt states vs pointer basis as projectors 35

6. Exact vs approximate consistency 36

7. Consistency and environment-inducedsuperselection 36

8. Summary and discussion 37

V. Concluding remarks 37

Acknowledgments 38

References 38

I. INTRODUCTION

The implications of the decoherence program for thefoundations of quantum mechanics have been the subjectof an ongoing debate since the first precise formulation ofthe program in the early 1980s. The key idea promotedby decoherence is the insight that realistic quantum sys-tems are never isolated, but are immersed in the sur-rounding environment and interact continuously with it.The decoherence program then studies, entirely withinthe standard quantum formalism (i.e., without addingany new elements in the mathematical theory or its in-terpretation), the resulting formation of quantum corre-lations between the states of the system and its envi-ronment and the often surprising effects of these system-environment interactions. In short, decoherence brings

Page 2: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

2

about a local suppression of interference between pre-ferred states selected by the interaction with the envi-ronment.

Bub (1997) termed decoherence part of the “new or-thodoxy” of understanding quantum mechanics—as theworking physicist’s way of motivating the postulates ofquantum mechanics from physical principles. Proponentsof decoherence called it an “historical accident” (Joos,2000, p. 13) that the implications for quantum mechan-ics and for the associated foundational problems wereoverlooked for so long. Zurek (2003b, p. 717) suggests

The idea that the “openness” of quantum sys-

tems might have anything to do with the transi-

tion from quantum to classical was ignored for

a very long time, probably because in classi-

cal physics problems of fundamental importance

were always settled in isolated systems.

When the concept of decoherence was first introducedto the broader scientific community by Zurek’s (1991)article in Physics Today, it elicited a series of contentiouscomments from the readership (see the April 1993 issue ofPhysics Today). In response to his critics, Zurek (2003b,p. 718) states

In a field where controversy has reigned for so

long this resistance to a new paradigm [namely,

to decoherence] is no surprise.

Omnes (2002, p. 2) had this assessment:

The discovery of decoherence has already much

improved our understanding of quantum mechan-

ics. (. . . ) [B]ut its foundation, the range of its

validity and its full meaning are still rather ob-

scure. This is due most probably to the fact that

it deals with deep aspects of physics, not yet fully

investigated.

In particular, the question whether decoherence provides,or at least suggests, a solution to the measurement prob-lem of quantum mechanics has been discussed for severalyears. For example, Anderson (2001, p. 492) writes in anessay review

The last chapter (. . . ) deals with the quantum

measurement problem (. . . ). My main test, al-

lowing me to bypass the extensive discussion, was

a quick, unsuccessful search in the index for the

word “decoherence” which describes the process

that used to be called “collapse of the wave func-

tion.”

Zurek speaks in various places of the “apparent” or “ef-fective” collapse of the wave function induced by the in-teraction with environment (when embedded into a min-imal additional interpretive framework) and concludes(Zurek, 1998, p. 1793)

A “collapse” in the traditional sense is no longer

necessary. (. . . ) [The] emergence of “objective

existence” [from decoherence] (. . . ) significantly

reduces and perhaps even eliminates the role of

the “collapse” of the state vector.

D’Espagnat, who considers the explanation of our ex-periences (i.e., of “appearances”) as the only “sure” re-quirement of a physical theory, states (d’Espagnat, 2000,p. 136)

For macroscopic systems, the appearances are

those of a classical world (no interferences etc.),

even in circumstances, such as those occurring in

quantum measurements, where quantum effects

take place and quantum probabilities intervene

(. . . ). Decoherence explains the just mentioned

appearances and this is a most important result.

(. . . ) As long as we remain within the realm of

mere predictions concerning what we shall ob-

serve (i.e., what will appear to us)—and refrain

from stating anything concerning “things as they

must be before we observe them”—no break in

the linearity of quantum dynamics is necessary.

In his monumental book on the foundations of quantummechanics (QM), Auletta (2000, p. 791) concludes that

the Measurement theory could be part of the in-

terpretation of QM only to the extent that it

would still be an open problem, and we think

that this is largely no longer the case.

This is mainly so because, according to Auletta (2000,p. 289),

decoherence is able to solve practically all the

problems of Measurement which have been dis-

cussed in the previous chapters.

On the other hand, even leading adherents of decoherencehave expressed caution or even doubt that decoherencehas solved the measurement problem. Joos (2000, p. 14)writes

Does decoherence solve the measurement prob-

lem? Clearly not. What decoherence tells us, is

that certain objects appear classical when they

are observed. But what is an observation? At

some stage, we still have to apply the usual prob-

ability rules of quantum theory.

Along these lines, Kiefer and Joos (1999, p. 5) warn that

One often finds explicit or implicit statements to

the effect that the above processes are equivalent

to the collapse of the wave function (or even solve

the measurement problem). Such statements are

certainly unfounded.

In a response to Anderson’s (2001, p. 492) comment,Adler (2003, p. 136) states

I do not believe that either detailed theoretical

calculations or recent experimental results show

that decoherence has resolved the difficulties as-

sociated with quantum measurement theory.

Page 3: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

3

Similarly, Bacciagaluppi (2003b, p. 3) writes

Claims that simultaneously the measurement

problem is real [and] decoherence solves it are

confused at best.

Zeh asserts (Joos et al., 2003, Ch. 2)

Decoherence by itself does not yet solve the

measurement problem (. . . ). This argument is

nonetheless found wide-spread in the literature.

(. . . ) It does seem that the measurement problem

can only be resolved if the Schrodinger dynamics

(. . . ) is supplemented by a nonunitary collapse

(. . . ).

The key achievements of the decoherence program, apartfrom their implications for conceptual problems, do notseem to be universally understood either. Zurek (1998,p. 1800) remarks

[The] eventual diagonality of the density matrix

(. . . ) is a byproduct (. . . ) but not the essence

of decoherence. I emphasize this because diago-

nality of [the density matrix] in some basis has

been occasionally (mis-)interpreted as a key ac-

complishment of decoherence. This is mislead-

ing. Any density matrix is diagonal in some ba-

sis. This has little bearing on the interpretation.

These remarks show that a balanced discussion of thekey features of decoherence and their implications for thefoundations of quantum mechanics is overdue. The deco-herence program has made great progress over the pastdecade, and it would be inappropriate to ignore its rel-evance in tackling conceptual problems. However, it isequally important to realize the limitations of decoher-ence in providing consistent and noncircular answers tofoundational questions.

An excellent review of the decoherence program hasrecently been given by Zurek (2003b). It deals pri-marily with the technicalities of decoherence, althoughit contains some discussion on how decoherence can beemployed in the context of a relative-state interpreta-tion to motivate basic postulates of quantum mechanics.A helpful first orientation and overview, the entry byBacciagaluppi (2003a) in the Stanford Encyclopedia of

Philosophy features a relatively short (in comparison tothe present paper) introduction to the role of decoher-ence in the foundations of quantum mechanics, includingcomments on the relationship between decoherence andseveral popular interpretations of quantum theory. Inspite of these valuable recent contributions to the litera-ture, a detailed and self-contained discussion of the roleof decoherence in the foundations of quantum mechanicsseems still to be lacking. This review article is intendedto fill the gap.

To set the stage, we shall first, in Sec. II, review themeasurement problem, which illustrates the key difficul-ties that are associated with describing quantum mea-surement within the quantum formalism and that are all

in some form addressed by the decoherence program. InSec. III, we then introduce and discuss the main featuresof the theory of decoherence, with a particular emphasison their foundational implications. Finally, in Sec. IV,we investigate the role of decoherence in various inter-pretive approaches of quantum mechanics, in particularwith respect to the ability to motivate and support (ordisprove) possible solutions to the measurement problem.

II. THE MEASUREMENT PROBLEM

One of the most revolutionary elements introduced intophysical theory by quantum mechanics is the superposi-tion principle, mathematically founded in the linearity ofthe Hilbert state space. If |1〉 and |2〉 are two states, thenquantum mechanics tells us that any linear combinationα|1〉+β|2〉 also corresponds to a possible state. Whereassuch superpositions of states have been experimentallyextensively verified for microscopic systems (for instance,through the observation of interference effects), the appli-cation of the formalism to macroscopic systems appearsto lead immediately to severe clashes with our experienceof the everyday world. A book has never been ever ob-served to be in a state of being both “here” and “there”(i.e., to be in a superposition of macroscopically distin-guishable positions), nor does a Schrodinger cat that isa superposition of being alive and dead bear much re-semblence to reality as we perceive it. The problem is,then, how to reconcile the vastness of the Hilbert space ofpossible states with the observation of a comparativelyfew “classical” macrosopic states, defined by having asmall number of determinate and robust properties suchas position and momentum. Why does the world appearclassical to us, in spite of its supposed underlying quan-tum nature, which would, in principle, allow for arbitrarysuperpositions?

A. Quantum measurement scheme

This question is usually illustrated in the contextof quantum measurement where microscopic superposi-tions are, via quantum entanglement, amplified into themacroscopic realm and thus lead to very “nonclassical”states that do not seem to correspond to what is actuallyperceived at the end of the measurement. In the idealmeasurement scheme devised by von Neumann (1932),a (typically microscopic) system S, represented by ba-sis vectors |sn〉 in a Hilbert space HS , interacts witha measurement apparatus A, described by basis vectors|an〉 spanning a Hilbert space HA, where the |an〉 areassumed to correspond to macroscopically distinguish-able “pointer” positions that correspond to the outcome

Page 4: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

4

of a measurement if S is in the state |sn〉.1Now, if S is in a (microscopically “unproblematic”)

superposition∑

n cn|sn〉, and A is in the initial “ready”state |ar〉, the linearity of the Schrodinger equation en-tails that the total system SA, assumed to be representedby the Hilbert product space HS⊗HA, evolves accordingto

( ∑

n

cn|sn〉)|ar〉 t−→

n

cn|sn〉|an〉. (2.1)

This dynamical evolution is often referred to as a pre-

measurement in order to emphasize that the process de-scribed by Eq. (2.1) does not suffice to directly concludethat a measurement has actually been completed. This isso for two reasons. First, the right-hand side is a super-

position of system-apparatus states. Thus, without sup-plying an additional physical process (say, some collapsemechanism) or giving a suitable interpretation of such asuperposition, it is not clear how to account, given the fi-nal composite state, for the definite pointer positions thatare perceived as the result of an actual measurement—i.e., why do we seem to perceive the pointer to be inone position |an〉 but not in a superposition of posi-tions? This is the problem of definite outcomes. Sec-ond, the expansion of the final composite state is in gen-eral not unique, and therefore the measured observableis not uniquely defined either. This is the problem of

the preferred basis. In the literature, the first difficultyis typically referred to as the measurement problem, butthe preferred-basis problem is at least equally important,since it does not make sense even to inquire about specificoutcomes if the set of possible outcomes is not clearly de-fined. We shall therefore regard the measurement prob-lem as composed of both the problem of definite outcomesand the problem of the preferred basis, and discuss thesecomponents in more detail in the following.

B. The problem of definite outcomes

1. Superpositions and ensembles

The right-hand side of Eq. (2.1) implies that after thepremeasurement the combined system SA is left in a purestate that represents a linear superposition of system-pointer states. It is a well-known and important prop-erty of quantum mechanics that a superposition of statesis fundamentally different from a classical ensemble ofstates, where the system actually is in only one of thestates but we simply do not know in which (this is often

1 Note that von Neumann’s scheme is in sharp contrast to theCopenhagen interpretation, where measurement is not treated asa system-apparatus interaction described by the usual quantum-mechanical formalism, but instead as an independent componentof the theory, to be represented entirely in fundamentally classi-cal terms.

referred to as an “ignorance-interpretable,” or “proper”ensemble).

This can be shown explicitely, especially on micro-scopic scales, by performing experiments that lead to thedirect observation of interference patterns instead of therealization of one of the terms in the superposed purestate, for example, in a setup where electrons pass in-dividually (one at a time) through a double slit. As iswell known, this experiment clearly shows that, withinthe standard quantum-mechanical formalism, the elec-tron must not be described by either one of the wavefunctions describing the passage through a particular slit(ψ1 or ψ2), but only by the superposition of these wavefunctions (ψ1+ψ2), since the correct density distribution of the pattern on the screen is not given by the sum ofthe squared wave functions describing the addition of in-dividual passages through a single slit ( = |ψ1|2+ |ψ2|2),but only by the square of the sum of the individual wavefunctions ( = |ψ1 + ψ2|2).

Put differently, if an ensemble interpretation could beattached to a superposition, the latter would simply rep-resent an ensemble of more fundamentally determinedstates, and based on the additional knowledge broughtabout by the results of measurements, we could simplychoose a subensemble consisting of the definite pointerstate obtained in the measurement. But then, since thetime evolution has been strictly deterministic accordingto the Schrodinger equation, we could backtrack thissubensemble in time and thus also specify the initialstate more completely (“postselection”), and thereforethis state necessarily could not be physically identicalto the initially prepared state on the left-hand side ofEq. (2.1).

2. Superpositions and outcome attribution

In the standard (“orthodox”) interpretation of quan-tum mechanics, an observable corresponding to a phys-ical quantity has a definite value if and only if the sys-tem is in an eigenstate of the observable; if the systemis, however, in a superposition of such eigenstates, as inEq. (2.1), it is, according to the orthodox interpretation,meaningless to speak of the state of the system as havingany definite value of the observable at all. (This is fre-quently referred to as the so-called eigenvalue-eigenstatelink, or “e-e link” for short.) The e-e link, however, isby no means forced upon us by the structure of quan-tum mechanics or by empirical constraints (Bub, 1997).The concept of (classical) “values” that can be ascribedthrough the e-e link based on observables and the exis-tence of exact eigenstates of these observables has there-fore frequently been either weakened or altogether aban-donded. For instance, outcomes of measurements aretypically registered in position space (pointer positions,etc.), but there exist no exact eigenstates of the positionoperator, and the pointer states are never exactly mutu-ally orthogonal. One might then (explicitely or implic-

Page 5: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

5

itly) promote a “fuzzy” e-e link, or give up the concept ofobservables and values entirely and directly interpret thetime-evolved wave functions (working in the Schrodingerpicture) and the corresponding density matrices. Also,if it is regarded as sufficient to explain our perceptionsrather than describe the “absolute” state of the entireuniverse (see the argument below), one might only re-quire that the (exact or fuzzy) e-e link hold in a “rela-tive” sense, i.e., for the state of the rest of the universerelative to the state of the observer.

Then, to solve the problem of definite outcomes, someinterpretations (for example, modal interpretations andrelative-state interpretations) interpret the final-state su-perposition in such a way as to explain the existence, orat least the subjective perception, of “outcomes” evenif the final composite state has the form of a super-position. Other interpretations attempt to solve themeasurement problem by modifying the strictly unitarySchrodinger dynamics. Most prominently, the ortho-dox interpretation postulates a collapse mechanism thattransforms a pure-state density matrix into an ignorance-interpretable ensemble of individual states (a “propermixture”). Wave-function collapse theories add stochas-tic terms to the Schrodinger equation that induce an ef-fective (albeit only approximate) collapse for states ofmacroscopic systems (Ghirardi et al., 1986; Gisin, 1984;Pearle, 1979, 1999), while other authors suggested thatcollapse occurs at the level of the mind of a conscious ob-server (Stapp, 1993; Wigner, 1963). Bohmian mechanics,on the other hand, upholds a unitary time evolution ofthe wavefunction, but introduces an additional dynam-ical law that explicitely governs the always-determinatepositions of all particles in the system.

3. Objective vs. subjective definiteness

In general, (macroscopic) definiteness—and thus a so-lution to the problem of outcomes in the theory of quan-tum measurement—can be achieved either on an onto-

logical (objective) or an observational (subjective) level.Objective definiteness aims at ensuring “actual” definite-ness in the macroscopic realm, whereas subjective defi-niteness only attempts to explain why the macroscopicworld appears to be definite—and thus does not makeany claims about definiteness of the underlying physi-cal reality (whatever this reality might be). This raisesthe question of the significance of this distinction withrespect to the formation of a satisfactory theory of thephysical world. It might appear that a solution to themeasurement problem based on ensuring subjective, butnot objective, definiteness is merely good “for all prac-tical purposes”—abbreviated, rather disparagingly, as“FAPP” by Bell (1990)—and thus not capable of solv-ing the “fundamental” problem that would seem relevantto the construction of the “precise theory” that Bell de-manded so vehemently.

It seems to the author, however, that this critism is

not justified, and that subjective definiteness should beviewed on a par with objective definitess with respectto a satisfactory solution to the measurement problem.We demand objective definiteness because we experiencedefiniteness on the subjective level of observation, andit should not be viewed as an a priori requirement fora physical theory. If we knew independently of our ex-perience that definiteness existed in nature, subjectivedefiniteness would presumably follow as soon as we hademployed a simple model that connected the “external”physical phenomena with our “internal” perceptual andcognitive apparatus, where the expected simplicity ofsuch a model can be justified by referring to the pre-sumed identity of the physical laws governing externaland internal processes. But since knowledge is based onexperience, that is, on observation, the existence of ob-jective definiteness could only be derived from the obser-vation of definiteness. And, moreover, observation tellsus that definiteness is in fact not a universal propertyof nature, but rather a property of macroscopic objects,where the borderline to the macroscopic realm is diffi-cult to draw precisely; mesoscopic interference experi-ments have demonstrated clearly the blurriness of theboundary. Given the lack of a precise definition of theboundary, any demand for fundamental definiteness onthe objective level should be based on a much deeper andmore general commitment to a definiteness that appliesto every physical entity (or system) across the board, re-gardless of spatial size, physical property, and the like.

Therefore, if we realize that the often deeply felt com-mitment to a general objective definiteness is only basedon our experience of macroscopic systems, and that thisdefiniteness in fact fails in an observable manner for mi-croscopic and even certain mesoscopic systems, the au-thor sees no compelling grounds on which objective def-initeness must be demanded as part of a satisfactoryphysical theory, provided that the theory can account forsubjective, observational definiteness in agreement withour experience. Thus the author suggests that the samelegitimacy be attributed to proposals for a solution ofthe measurement problem that achieve “only” subjectivebut not objective definiteness—after all, the measure-ment problem arises solely from a clash of our experi-ence with certain implications of the quantum formalism.D’Espagnat (2000, pp. 134–135) has advocated a similarviewpoint:

The fact that we perceive such “things” as macro-

scopic objects lying at distinct places is due,

partly at least, to the structure of our sensory and

intellectual equipment. We should not, there-

fore, take it as being part of the body of sure

knowledge that we have to take into account for

defining a quantum state. (. . . ) In fact, scien-

tists most righly claim that the purpose of science

is to describe human experience, not to describe

“what really is”; and as long as we only want to

describe human experience, that is, as long as we

are content with being able to predict what will

Page 6: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

6

be observed in all possible circumstances (. . . )

we need not postulate the existence—in some

absolute sense—of unobserved (i.e., not yet ob-

served) objects lying at definite places in ordinary

3-dimensional space.

C. The preferred-basis problem

The second difficulty associated with quantum mea-surement is known as the preferred-basis problem, whichdemonstrates that the measured observable is in generalnot uniquely defined by Eq. (2.1). For any choice of sys-tem states |sn〉, we can find corresponding apparatusstates |an〉, and vice versa, to equivalently rewrite thefinal state emerging from the premeasurement interac-tion, i.e., the right-hand side of Eq. (2.1). In general,however, for some choice of apparatus states the corre-sponding new system states will not be mutually orthog-onal, so that the observable associated with these stateswill not be Hermitian, which is usually not desired (how-ever, not forbidden—see the discussion by Zurek, 2003b).Conversely, to ensure distinguishable outcomes, we must,in general, require the (at least approximate) orthogonal-ity of the apparatus (pointer) states, and it then followsfrom the biorthogonal decomposition theorem that theexpansion of the final premeasurement system-apparatusstate of Eq. (2.1),

|ψ〉 =∑

n

cn|sn〉|an〉, (2.2)

is unique, but only if all coefficients cn are distinct. Oth-erwise, we can in general rewrite the state in terms ofdifferent state vectors,

|ψ〉 =∑

n

c′n|s′n〉|a′n〉, (2.3)

such that the same postmeasurement state seems to cor-respond to two different measurements, that is, of the

observables A =∑

n λn|sn〉〈sn| and B =∑

n λ′n|s′n〉〈s′n|

of the system, respectively, although in general A and Bdo not commute.

As an example, consider a Hilbert space H = H1 ⊗H2

where H1 and H2 are two-dimensional spin spaces withstates corresponding to spin up or spin down along agiven axis. Suppose we are given an entangled spin stateof the Einstein-Podolsky-Rosen form (Einstein et al.,1935)

|ψ〉 =1√2(|z+〉1|z−〉2 − |z−〉1|z+〉2), (2.4)

where |z±〉1,2 represents the eigenstates of the observableσz corresponding to spin up or spin down along the z axisof the two systems 1 and 2. The state |ψ〉 can howeverequivalently be expressed in the spin basis correspondingto any other orientation in space. For example, when

using the eigenstates |x±〉1,2 of the observable σx (whichrepresents a measurement of the spin orientation alongthe x axis) as basis vectors, we get

|ψ〉 =1√2(|x+〉1|x−〉2 − |x−〉1|x+〉2). (2.5)

Now suppose that system 2 acts as a measuring device forthe spin of system 1. Then Eqs. (2.4) and (2.5) imply thatthe measuring device has established a correlation withboth the z and the x spin of system 1. This means that, ifwe interpret the formation of such a correlation as a mea-surement in the spirit of the von Neumann scheme (with-out assuming a collapse), our apparatus (system 2) couldbe considered as having measured also the x spin once ithas measured the z spin, and vice versa—in spite of thenoncommutativity of the corresponding spin observablesσz and σx. Moreover, since we can rewrite Eq. (2.4) ininfinitely many ways, it appears that once the apparatushas measured the spin of system 1 along one direction, itcan also be regarded as having measured the spin alongany other direction, again in apparent contradiction withquantum mechanics due to the noncommutativity of thespin observables corresponding to different spatial orien-tations.

It thus seems that quantum mechanics has nothing tosay about which observable(s) of the system is (are) beingrecorded, via the formation of quantum correlations, bythe apparatus. This can be stated in a general theorem(Auletta, 2000; Zurek, 1981): When quantum mechanicsis applied to an isolated composite object consisting of asystem S and an apparatus A, it cannot determine whichobservable of the system has been measured—in obviouscontrast to our experience of the workings of measuringdevices that seem to be “designed” to measure certainquantities.

D. The quantum-to-classical transition and decoherence

In essence, as we have seen above, the measurementproblem deals with the transition from a quantum world,described by essentially arbitrary linear superpositions ofstate vectors, to our perception of “classical” states in themacroscopic world, that is, a comparatively small subsetof the states allowed by the quantum-mechanical super-position principle, having only a few, but determinateand robust, properties, such as position, momentum, etc.The question of why and how our experience of a “clas-sical” world emerges from quantum mechanics thus liesat the heart of the foundational problems of quantumtheory.

Decoherence has been claimed to provide an explana-tion for this quantum-to-classical transition by appeal-ing to the ubiquitous immersion of virtually all physicalsystems in their environment (“environmental monitor-ing”). This trend can also be read off nicely from thetitles of some papers and books on decoherence, for ex-ample, “The emergence of classical properties through

Page 7: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

7

interaction with the environment” (Joos and Zeh, 1985),“Decoherence and the transition from quantum to clas-sical” (Zurek, 1991), and “Decoherence and the appear-ance of a classical world in quantum theory” (Joos et al.,2003). We shall critically investigate in this paper towhat extent the appeal to decoherence for an explana-tion of the quantum-to-classical transition is justified.

III. THE DECOHERENCE PROGRAM

As remarked earlier, the theory of decoherence is basedon a study of the effects brought about by the inter-action of physical systems with their environment. Inclassical physics, the environment is usually viewed asa kind of disturbance, or noise, that perturbs the sys-tem under consideration in such a way as to negativelyinfluence the study of its “objective” properties. There-fore science has established the idealization of isolatedsystems, with experimental physics aiming at eliminat-ing any outer sources of disturbance as much as possiblein order to discover the “true” underlying nature of thesystem under study.

The distinctly nonclassical phenomenon of quantumentanglement, however, has demonstrated that the cor-relations between two systems can be of fundamental im-portance and can lead to properties that are not presentin the individual systems.2 The earlier view of phenom-ena arising from quantum entanglement as “paradoxa”has generally been replaced by the recognition of entan-glement as a fundamental property of nature.

The decoherence program3 is based on the idea thatsuch quantum correlations are ubiquitous; that nearlyevery physical system must interact in some way withits environment (for example, with the surrounding pho-tons that then create the visual experience within theobserver), which typically consists of a large numberof degrees of freedom that are hardly ever fully con-trolled. Only in very special cases of typically micro-scopic (atomic) phenomena, so goes the claim of the de-coherence program, is the idealization of isolated systemsapplicable so that the predictions of linear quantum me-chanics (i.e., a large class of superpositions of states)can actually be observationally confirmed. In the ma-jority of the cases accessible to our experience, however,interaction with the environment is so dominant as topreclude the observation of the “pure” quantum world,imposing effective superselection rules (Cisnerosy et al.,1998; Galindo et al., 1962; Giulini, 2000; Wick et al.,1952, 1970; Wightman, 1995) onto the space of observ-able states that lead to states corresponding to the “clas-

2 Broadly speaking, this means that the (quantum-mechanical)whole is different from the sum of its parts.

3 For key ideas and concepts, see Joos and Zeh (1985); Joos et al.

(2003); Kubler and Zeh (1973); Zeh (1970, 1973, 1995, 1997,2000); Zurek (1981, 1982, 1991, 1993, 2003b).

sical” properties of our experience. Interference betweensuch states gets locally suppressed and is thus claimed tobecome inaccessible to the observer.

Probably the most surprising aspect of decoherenceis the effectiveness of the system-environment interac-tions. Decoherence typically takes place on extremelyshort time scales and requires the presence of only aminimal environment (Joos and Zeh, 1985). Due tothe large number of degrees of freedom of the envi-ronment, it is usually very difficult to undo system-environment entanglement, which has been claimed as asource of our impression of irreversibility in nature (see,for example, Kiefer and Joos, 1999; Zeh, 2001; Zurek,1982, 2003b; Zurek and Paz, 1994). In general, the ef-fect of decoherence increases with the size of the sys-tem (from microscopic to macroscopic scales), but it isimportant to note that there exist, admittedly some-what exotic, examples for which the decohering influ-ence of the environment can be sufficiently shielded tolead to mesoscopic and even macroscopic superpositions.One such example would be the case of superconduct-ing quantum interference devices (SQUIDs), in which su-perpositions of macroscopic currents become observable(Friedman et al., 2000; van der Wal et al., 2000). Con-versely, some microscopic systems (for instance, certainchiral molecules that exist in different distinct spatialconfigurations) can be subject to remarkably strong de-coherence.

The decoherence program has dealt with the followingtwo main consequences of environmental interaction:

(1) Environment-induced decoherence. The fast localsuppression of interference between different statesof the system. However, since only unitary timeevolution is employed, global phase coherence isnot actually destroyed—it becomes absent fromthe local density matrix that describes the sys-tem alone, but remains fully present in the to-tal system-environment composition.4 We shalldiscuss environment-induced local decoherence inmore detail in Sec. III.D.

(2) Environment-induced superselection. The selectionof preferred sets of states, often referred to as“pointer states,” that are robust (in the sense ofretaining correlations over time) in spite of theirimmersion in the environment. These states aredetermined by the form of the interaction betweenthe system and its environment and are suggestedto correspond to the “classical” states of our ex-perience. We shall consider this mechanism inSec. III.E.

4 Note that the persistence of coherence in the total state is impor-tant to ensure the possibility of describing special cases in whichmesoscopic or macrosopic superpositions have been experimen-tally realized.

Page 8: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

8

Another, more recent aspect of the decoherence program,termed enviroment-assisted invariance or “envariance,”was introduced by Zurek (2003b,c, 2004b) and furtherdeveloped in Zurek (2004a). In particular, Zurek usedenvariance to explain the emergence of probabilities inquantum mechanics and to derive Born’s rule based oncertain assumptions. We shall review envariance andZurek’s derivation of the Born rule in Sec. III.F.

Finally, let us emphasize that decoherence arises from adirect application of the quantum mechanical formalismto a description of the interaction of a physical systemwith its environment. By itself, decoherence is thereforeneither an interpretation nor a modification of quantummechanics. Yet the implications of decoherence need tobe interpreted in the context of the different interpre-tations of quantum mechanics. Also, since decoherenceeffects have been studied extensively in both theoreticalmodels and experiments (for a survey, see, for example,Joos et al., 2003; Zurek, 2003b), their existence can betaken as a well-confirmed fact.

A. Resolution into subsystems

Note that decoherence derives from the presuppositionof the existence and the possibility of a division of theworld into “system(s)” and “environment.” In the deco-herence program, the term “environment” is usually un-derstood as the “remainder” of the system, in the sensethat its degrees of freedom are typically not (cannot be,do not need to be) controlled and are not directly rele-vant to the observation under consideration (for example,the many microsopic degrees of freedom of the system),but that nonetheless the environment includes “all thosedegrees of freedom which contribute significantly to theevolution of the state of the apparatus” (Zurek, 1981,p. 1520).

This system–environment dualism is generally associ-ated with quantum entanglement, which always describesa correlation between parts of the universe. As long asthe universe is not resolved into individual subsystems,there is no measurement problem: the state vector |Ψ〉of the entire universe5 evolves deterministically accord-

ing to the Schrodinger equation i~ ∂∂t |Ψ〉 = H |Ψ〉, which

poses no interpretive difficulty. Only when we decom-pose the total Hilbert-state space H of the universe intoa product of two spaces H1 ⊗H2, and accordingly formthe joint-state vector |Ψ〉 = |Ψ1〉|Ψ2〉, and want to ascribean individual state (besides the joint state that describesa correlation) to one of the two systems (say, the appara-tus), does the measurement problem arise. Zurek (2003b,p. 718) puts it like this:

5 If we dare to postulate this total state—see counterarguments byAuletta (2000).

In the absence of systems, the problem of inter-

pretation seems to disappear. There is simply

no need for “collapse” in a universe with no sys-

tems. Our experience of the classical reality does

not apply to the universe as a whole, seen from

the outside, but to the systems within it.

Moreover, terms like “observation,” “correlation,” and“interaction” will naturally make little sense without adivision into systems. Zeh has suggested that the localityof the observer defines an observation in the sense thatany observation arises from the ignorance of a part of theuniverse; and that this also defines the “facts” that canoccur in a quantum system. Landsman (1995, pp. 45–46)argues similarly:

The essence of a “measurement,” “fact” or

“event” in quantum mechanics lies in the non-

observation, or irrelevance, of a certain part of

the system in question. (. . . ) A world without

parts declared or forced to be irrelevant is a world

without facts.

However, the assumption of a decomposition of the uni-verse into subsystems—as necessary as it appears to befor the emergence of the measurement problem and forthe definition of the decoherence program—is definitelynontrivial. By definition, the universe as a whole is aclosed system, and therefore there are no “unobserveddegrees of freedom” of an external environment whichwould allow for the application of the theory of decoher-ence to determine the space of quasiclassical observablesof the universe in its entirety. Also, there exists no gen-eral criterion for how the total Hilbert space is to bedivided into subsystems, while at the same time muchof what is called a property of the system will dependon its correlation with other systems. This problem be-comes particularly acute if one would like decoherencenot only to motivate explanations for the subjective per-ception of classicality (as in Zurek’s “existential inter-pretation,” see Zurek, 1993, 1998, 2003b, and Sec. IV.Cbelow), but moreover to allow for the definition of quasi-classical “macrofacts.” Zurek (1998, p. 1820) admits thissevere conceptual difficulty:

In particular, one issue which has been often

taken for granted is looming big, as a founda-

tion of the whole decoherence program. It is the

question of what are the “systems” which play

such a crucial role in all the discussions of the

emergent classicality. (. . . ) [A] compelling ex-

planation of what are the systems—how to define

them given, say, the overall Hamiltonian in some

suitably large Hilbert space—would be undoubt-

edly most useful.

A frequently proposed idea is to abandon the notion of an“absolute” resolution and instead postulate the intrinsicrelativity of the distinct state spaces and properties thatemerge through the correlation between these relatively

Page 9: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

9

defined spaces (see, for example, the proposals, unrelatedto decoherence, of Everett, 1957; Mermin, 1998a,b; andRovelli, 1996). This relative view of systems and corre-lations has counterintuitive, in the sense of nonclassical,implications. However, as in the case of quantum entan-glement, these implications need not be taken as para-doxa that demand further resolution. Accepting someproperties of nature as counterintuitive is indeed a satis-factory path to take in order to arrive at a description ofnature that is as complete and objective as is allowed bythe range of our experience (which is based on inherentlylocal observations).

B. The concept of reduced density matrices

Since reduced density matrices are a key tool of deco-herence, it will be worthwile to briefly review their ba-sic properties and interpretation in the following. Theconcept of reduced density matrices emerged in the ear-liest days of quantum mechanics (Furry, 1936; Landau,1927; von Neumann, 1932; for some historical remarks,see Pessoa, 1998). In the context of a system of two en-tangled systems in a pure state of the Einstein-Podolsky-Rosen-type,

|ψ〉 =1√2(|+〉1|−〉2 − |−〉1|+〉2), (3.1)

it had been realized early that for an observable O that

pertains only to system 1, O = O1 ⊗ I2, the pure-statedensity matrix ρ = |ψ〉〈ψ| yields, according to the trace

rule 〈O〉 = Tr(ρO) and given the usual Born rule forcalculating probabilities, exactly the same statistics asthe reduced density matrix ρ1 obtained by tracing overthe degrees of freedom of system 2 (i.e., the states |+〉2and |−〉2),

ρ1 = Tr2|ψ〉〈ψ| = 2〈+|ψ〉〈ψ|+〉2 + 2〈−|ψ〉〈ψ|−〉2, (3.2)

since it is easy to show that, for this observable O,

〈O〉ψ = Tr(ρO) = Tr1(ρ1O1). (3.3)

This result holds in general for any pure state |ψ〉 =∑i αi|φi〉1|φi〉2 · · · |φi〉N of a resolution of a system into

N subsystems, where the |φi〉j are assumed to formorthonormal basis sets in their respective Hilbert spaces

Hj , j = 1 · · ·N . For any observable O that pertains only

to system j, O = I1⊗ I2⊗· · ·⊗ Ij−1⊗Oj⊗ Ij+1⊗· · ·⊗ IN ,

the statistics of O generated by applying the trace rulewill be identical regardless of whether we use the pure-state density matrix ρ = |ψ〉〈ψ| or the reduced density

matrix ρj = Tr1,...,j−1,j+1,...,N |ψ〉〈ψ|, since again 〈O〉 =

Tr(ρO) = Trj(ρjOj).The typical situation in which the reduced density ma-

trix arises is this: Before a premeasurement-type inter-action, the observer knows that each individual system is

in some (unknown) pure state. After the interaction, i.e.,after the correlation between the systems is established,the observer has access to only one of the systems, say,system 1; everything that can be known about the stateof the composite system must therefore be derived frommeasurements on system 1, which will yield the possibleoutcomes of system 1 and their probability distribution.All information that can be extracted by the observeris then, exhaustively and correctly, contained in the re-duced density matrix of system 1, assuming that the Bornrule for quantum probabilities holds.

Let us return to the Einstein-Podolsky-Rosen-type ex-ample, Eqs. (3.1) and (3.2). If we assume that the statesof system 2 are orthogonal, 2〈+|−〉2 = 0, ρ1 becomesdiagonal,

ρ1 = Tr2|ψ〉〈ψ| =1

2(|+〉〈+|)1 +

1

2(|−〉〈−|)1. (3.4)

But this density matrix is formally identical to the den-sity matrix that would be obtained if system 1 were in amixed state, i.e., in either one of the two states |+〉1 and|−〉1 with equal probabilties—as opposed to the super-position |ψ〉, in which both terms are considered present,which could in principle be confirmed by suitable inter-ference experiments. This implies that a measurementof an observable that only pertains to system 1 cannotdiscriminate between the two cases, pure vs mixed state.6

However, note that the formal identification of the re-duced density matrix with a mixed-state density matrixis easily misinterpreted as implying that the state of thesystem can be viewed as mixed too (see also the discus-sion by d’Espagnat, 1988). Density matrices are only acalculational tool for computing the probability distri-bution of a set of possible outcomes of measurements;they do not specify the state of the system.7 Since thetwo systems are entangled and the total composite sys-tem is still described by a superposition, it follows fromthe standard rules of quantum mechanics that no indi-vidual definite state can be attributed to one of the sys-tems. The reduced density matrix looks like a mixed-state density matrix because, if one actually measuredan observable of the system, one would expect to get adefinite outcome with a certain probability; in terms ofmeasurement statistics, this is equivalent to the situationin which the system is in one of the states from the set ofpossible outcomes from the beginning, that is, before themeasurement. As Pessoa (1998, p. 432) puts it, “takinga partial trace amounts to the statistical version of theprojection postulate.”

6 As discussed by Bub (1997, pp. 208–210), this result also holdsfor any observable of the composite system that factorizes intothe form O = O1 ⊗ O2, where O1 and O2 do not commute withthe projection operators (|±〉〈±|)1 and (|±〉〈±|)2, respectively.

7 In this context we note that any nonpure density matrix can bewritten in many different ways, demonstrating that any partitionin a particular ensemble of quantum states is arbitrary.

Page 10: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

10

C. A modified von Neumann measurement scheme

Let us now reconsider the von Neumann model forideal quantum-mechanical measurement, Eq. (3.5), butnow with the environment included. We shall denote theenvironment by E and represent its state before the mea-surement interaction by the initial state vector |e0〉 ina Hilbert space HE . As usual, let us assume that thestate space of the composite object system-apparatus-environment is given by the tensor product of the indi-vidual Hilbert spaces, HS ⊗ HA ⊗ HE . The linearity ofthe Schrodinger equation then yields the following timeevolution of the entire system SAE ,

( ∑

n

cn|sn〉)|ar〉|e0〉

(1)−→( ∑

n

cn|sn〉|an〉)|e0〉

(2)−→∑

n

cn|sn〉|an〉|en〉, (3.5)

where the |en〉 are the states of the environment associ-ated with the different pointer states |an〉 of the measur-ing apparatus. Note that while for two subsystems, say,S and A, there always exists a diagonal (“Schmidt”) de-composition of the final state of the form

∑n cn|sn〉|an〉,

for three subsystems (for example, S, A, and E), a de-composition of the form

∑n cn|sn〉|an〉|en〉 is not always

possible. This implies that the total Hamiltonian thatinduces a time evolution of the above kind, Eq. (3.5),must be of a special form.8

Typically, the |en〉 will be product states of many mi-crosopic subsystem states |εn〉i corresponding to the in-dividual parts that form the environment, i.e., |en〉 =|εn〉1|εn〉2|εn〉3 · · · . We see that a nonseparable and inmost cases, for all practical purposes, irreversible (due tothe enormous number of degrees of freedom of the en-vironment) correlation has been established between thestates of the system–apparatus combination SA and thedifferent states of the environment E . Note that Eq. (3.5)also implies that the environment has recorded the stateof the system—and, equivalently, the state of the system-apparatus composition. The environment, composed ofmany subsystems, thus acts as an amplifying, higher-order measuring device.

D. Decoherence and local suppression of interference

Interaction with the environment typically leads to arapid vanishing of the diagonal terms in the local densitymatrix describing the probability distribution for the out-comes of measurements on the system. This effect has

8 For an example of such a Hamiltonian, see the model of Zurek(1981, 1982) and its outline in Sec. III.D.2 below. For a criti-cal comment regarding limitations on the form of the evolutionoperator and the possibility of a resulting disagreement with ex-perimental evidence, see Pessoa (1998).

become known as environment-induced decoherence, andit has also frequently been claimed to imply at least apartial solution to the measurement problem.

1. General formalism

In Sec. III.B, we have already introduced the conceptof local (or reduced) density matrices and pointed outsome caveats on their interpretation. In the context ofthe decoherence program, reduced density matrices ariseas follows. Any observation will typically be restricted tothe system-apparatus component, SA, while the manydegrees of freedom of the environment E remain unob-served. Of course, typically some degrees of freedom ofthe environment will always be included in our obser-vation (e.g., some of the photons scattered off the ap-paratus) and we shall accordingly include them in the“observed part SA of the universe.” The crucial pointis that there still remains a comparatively large numberof environmental degrees of freedom that will not be ob-served directly.

Suppose then that the operator OSA represents an ob-

servable of SA only. Its expectation value 〈OSA〉 is givenby

〈OSA〉 = Tr(ρSAE [OSA ⊗ IE ]) = TrSA(ρSAOSA), (3.6)

where the density matrix ρSAE of the total SAE combi-nation,

ρSAE =∑

mn

cmc∗n|sm〉|am〉|em〉〈sn|〈an|〈en|, (3.7)

has, for all practical purposes of statistical prediction,been replaced by the local (or reduced) density matrixρSA, obtained by “tracing out the unobserved degrees ofthe environment,” that is,

ρSA = TrE(ρSAE) =∑

mn

cmc∗n|sm〉|am〉〈sn|〈an|〈en|em〉.

(3.8)So far, ρSA contains characteristic interference terms|sm〉|am〉〈sn|〈an|, m 6= n, since we cannot assume fromthe outset that the basis vectors |em〉 of the environmentare necessarily mutually orthogonal, i.e., that 〈en|em〉 =0 if m 6= n. Many explicit physical models for the inter-action of a system with the environment (see Sec. III.D.2below for a simple example), however, have shown thatdue to the large number of subsystems that compose theenvironment, the pointer states |en〉 of the environmentrapidly approach orthogonality, 〈en|em〉(t) → δn,m, suchthat the reduced density matrix ρSA becomes approxi-mately orthogonal in the “pointer basis” |an〉; that is,

ρSAt−→ ρ dSA ≈

n

|cn|2|sn〉|an〉〈sn|〈an|

=∑

n

|cn|2P (S)n ⊗ P (A)

n . (3.9)

Page 11: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

11

Here, P(S)n and P

(A)n are the projection operators onto

the eigenstates of S and A, respectively. Therefore theinterference terms have vanished in this local represen-tation, i.e., phase coherence has been locally lost. Thisis precisely the effect referred to as environment-induceddecoherence. The decohered local density matrices de-scribing the probability distribution of the outcomes ofa measurement on the system-apparatus combination isformally (approximately) identical to the correspondingmixed-state density matrix. But as we pointed out inSec. III.B, we must be careful in interpreting this stateof affairs, since full coherence is retained in the total den-sity matrix ρSAE .

2. An exactly solvable two-state model for decoherence

To see how the approximate mutual orthogonality ofthe environmental state vectors arises, let us discuss asimple model first introduced by Zurek (1982). Considera system S with two spin states |⇑〉, |⇓〉 that inter-acts with an environment E described by a collectionof N other two-state spins represented by |↑k〉, |↓k〉,k = 1 · · ·N . The self-Hamiltonians HS and HE and theself-interaction Hamiltonian HEE of the environment aretaken to be equal to zero. Only the interaction Hamilto-

nian HSE that describes the coupling of the spin of thesystem to the spins of the environment is assumed to benonzero and of the form

HSE = (|⇑〉〈⇑|−|⇓〉〈⇓|)⊗∑

k

gk(|↑k〉〈↑k|−|↓k〉〈↓k|)⊗

k′ 6=k

Ik′ ,

(3.10)

where the gk are coupling constants and Ik = (|↑k〉〈↑k|+|↓k〉〈↓k|) is the identity operator for the kth environmen-tal spin. Applied to the initial state before the interactionis turned on,

|ψ(0)〉 = (a|⇑〉 + b|⇓〉)N⊗

k=1

(αk|↑k〉 + βk|↓k〉), (3.11)

this Hamiltonian yields a time evolution of the state givenby

|ψ(t)〉 = a|⇑〉|E⇑(t)〉 + b|⇓〉|E⇓(t)〉, (3.12)

where the two environmental states |E⇑(t)〉 and |E⇓(t)〉are

|E⇑(t)〉 = |E⇓(−t)〉 =

N⊗

k=1

(αkeigkt|↑k〉+βke−igkt|↓k〉).

(3.13)The reduced density matrix ρS(t) = TrE(|ψ(t)〉〈ψ(t)|) isthen

ρS(t) = |a|2|⇑〉〈⇑| + |b|2|⇓〉〈⇓|+z(t)ab∗|⇑〉〈⇓| + z∗(t)a∗b|⇓〉〈⇑|, (3.14)

where the interference coefficient z(t) which determinesthe weight of the off-diagonal elements in the reduceddensity matrix is given by

z(t) = 〈E⇑(t)|E⇓(t)〉 =

N∏

k=1

(|αk|2eigkt + |βk|2e−igkt),

(3.15)and thus

|z(t)|2 =

N∏

k=1

1+ [(|αk|2 −|βk|2)2−1] sin2 2gkt. (3.16)

At t = 0, z(t) = 1, i.e., the interference terms are fullypresent, as expected. If |αk|2 = 0 or 1 for each k, i.e.,if the environment is in an eigenstate of the interaction

Hamiltonian HSE of the type |↑1〉|↑2〉|↓3〉 · · · |↑N〉, and/orif 2gkt = mπ (m = 0, 1, . . .), then z(t)2 ≡ 1 so coherenceis retained over time. However, under realistic circum-stances, we can typically assume a random distributionof the initial states of the environment (i.e., of coefficientsαk, βk) and of the coupling coefficients gk. Then, in thelong-time average,

〈|z(t)|2〉t→∞ ≃ 2−NN∏

k=1

[1 + (|αk|2 − |βk|2)2] N→∞−→ 0,

(3.17)so the off-diagonal terms in the reduced density matrixbecome strongly damped for large N .

It can also be shown directly that, given very generalassumptions about the distribution of the couplings gk(namely, requiring their initial distribution to have finitevariance), z(t) exhibits a Gaussian time dependence of

the form z(t) ∝ eiAte−B2t2/2, where A and B are real

constants (Zurek et al., 2003). For the special case inwhich αk = α and gk = g for all k, this behavior of z(t)can be immediately seen by first rewriting z(t) as thebinomial expansion

z(t) = (|α|2eigt + |β|2e−igt)N

=

N∑

l=0

(N

l

)|α|2l|β|2(N−l)eig(2l−N)t. (3.18)

For large N , the binomial distribution can then be ap-proximated by a Gaussian,

(N

l

)|α|2l|β|2(N−l) ≈ e−(l−N |α|2)2/(2N |α|2|β|2)

√2πN |α|2|β|2

, (3.19)

in which case z(t) becomes

z(t) =

N∑

l=0

e−(l−N |α|2)2/(2N |α|2|β|2)

√2πN |α|2|β|2

eig(2l−N)t, (3.20)

that is, z(t) is the Fourier transform of an (approxi-mately) Gaussian distribution and is therefore itself (ap-proximately) Gaussian.

Page 12: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

12

Detailed model calculations, in which the environ-ment is typically represented by a more sophisticatedmodel consisting of a collection of harmonic oscillators(Caldeira and Leggett, 1983; Hu et al., 1992; Joos et al.,2003; Unruh and Zurek, 1989; Zurek, 2003b; Zurek et al.,1993), have shown that the damping occurs on extremelyshort decoherence time scales τD, which are typicallymany orders of magnitude shorter than the thermal relax-ation. Even microscopic systems such as large moleculesare rapidly decohered by the interaction with thermalradiation on a time scale that is much shorter than anypractical observation could resolve; for mesoscopic sys-tems such as dust particles, the 3K cosmic microwavebackground radiation is sufficient to yield strong and im-mediate decoherence (Joos and Zeh, 1985; Zurek, 1991).

Within τD, |z(t)| approaches zero and remains close tozero, fluctuating with an average standard deviation ofthe random-walk type σ ∼

√N (Zurek, 1982). However,

the multiple periodicity of z(t) implies that coherence,and thus the purity of the reduced density matrix, willreappear after a certain time τr, which can be shown tobe very long and of the Poincare type with τr ∼ N !. Formacroscopic environments of realistic but finite sizes, τrcan exceed the lifetime of the universe (Zurek, 1982), butnevertheless always remains finite.

From a conceptual point of view, recurrence of coher-ence is of little relevance. The recurrence time couldonly be infinitely long in the hypothetical case of an in-finitely large environment. In this situation, off-diagonalterms in the reduced density matrix would be irreversiblydamped and lost in the limit t → ∞, which has some-times been regarded as describing a physical collapse ofthe state vector (Hepp, 1972). But the assumption ofinfinite sizes and times is never realized in nature (Bell,1975), nor can information ever be truly lost (as achievedby a “true” state vector collapse) through unitary timeevolution—full coherence is always retained at all timesin the total density matrix ρSAE(t) = |ψ(t)〉〈ψ(t)|.

We can therefore state the general conclusion that,except for exceptionally well-isolated and carefully pre-pared microsopic and mesoscopic systems, the interac-tion of the system with the environment causes the off-diagonal terms of the local density matrix, expressed inthe pointer basis and describing the probability distribu-tion of the possible outcomes of a measurement on thesystem, to become extremely small in a very short pe-riod of time, and that this process is irreversible for allpractical purposes.

E. Environment-induced superselection

Let us now turn to the second main consequenceof the interaction with the environment, namely, theenvironment-induced selection of stable preferred-basisstates. We discussed in Sec. II.C the fact thatthe quantum-mechanical measurement scheme as repre-sented by Eq. (2.1) does not uniquely define the expan-

sion of the postmeasurement state and thereby leavesopen the question of which observable can be consideredas having been measured by the apparatus. This situa-tion is changed by the inclusion of the environment statesin Eq. (3.5), for the following two reasons:

(1) Environment-induced superselection of a preferred

basis. The interaction between the apparatus andthe environment singles out a set of mutually com-muting observables.

(2) The existence of a tridecompositional uniqueness

theorem (Bub, 1997; Clifton, 1994; Elby and Bub,1994). If a state |ψ〉 in a Hilbert space H1⊗H2⊗H3

can be decomposed into the diagonal (“Schmidt”)form |ψ〉 =

∑i αi|φi〉1|φi〉2|φi〉3, the expansion is

unique provided that the |φi〉1 and |φi〉2 aresets of linearly independent, normalized vectors inH1 and H2, respectively, and that |φi〉3 is a setof mutually noncollinear normalized vectors in H3.This can be generalized to an N -decompositionaluniqueness theorem, in which N ≥ 3. Note thatit is not always possible to decompose an arbitrarypure state of more than two systems (N ≥ 3) intothe Schmidt form |ψ〉 =

∑i αi|φi〉1|φi〉2 · · · |φi〉N ,

but if the decomposition exists, its uniqueness isguaranteed.

The tridecompositional uniqueness theorem ensuresthat the expansion of the final state in Eq. (3.5) is unique,which fixes the ambiguity in the choice of the set of pos-sible outcomes. It demonstrates that the inclusion of (atleast) a third “system” (here referred to as the environ-ment) is necessary to remove the basis ambiguity.

Of course, given any pure state in the compositeHilbert space H1 ⊗ H2 ⊗ H3, the tridecompositionaluniqueness theorem neither tells us whether a Schmidtdecomposition exists nor specifies the unique expansionitself (provided the decomposition is possible), and sincethe precise states of the environment are generally notknown, an additional criterion is needed that determineswhat the preferred states will be.

1. Stability criterion and pointer basis

The decoherence program has attempted to define sucha criterion based on the interaction with the environ-ment and the idea of robustness and preservation of cor-relations. The environment thus plays a double role insuggesting a solution to the preferred-basis problem: itselects a preferred pointer basis, and it guarantees itsuniqueness via the tridecompositional uniqueness theo-rem.

In order to motivate the basis superselection approachproposed by the decoherence program, we note that instep (2) of Eq. (3.5) we tacitly assumed that interactionwith the environment does not disturb the establishedcorrelation between the state of the system, |sn〉, and

Page 13: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

13

the corresponding pointer state |an〉. This assumptioncan be viewed as a generalization of the concept of “faith-ful measurement” to the realistic case in which the envi-ronment is included. Faithful measurement in the usualsense concerns step (1), namely, the requirement that themeasuring apparatus A act as a reliable “mirror” of thestates of the system S by forming only correlations ofthe form |sn〉|an〉 but not |sm〉|an〉 with m 6= n. Butsince any realistic measurement process must include theinevitable coupling of the apparatus to its environment,the measurement could hardly be considered faithful as awhole if the interaction with the environment disturbedthe correlations between the system and the apparatus.9

It was therefore first suggested by Zurek (1981) thatthe preferred pointer basis be taken as the basis which“contains a reliable record of the state of the system S”(Zurek, 1981, p. 1519), i.e., the basis in which the system-apparatus correlations |sn〉|an〉 are left undisturbed bythe subsequent formation of correlations with the envi-ronment (the stability criterion). One can then find asufficient criterion for dynamically stable pointer statesthat preserve the system–apparatus correlations in spiteof the interaction of the apparatus with the environ-ment by requiring all pointer state projection opera-

tors P(A)n = |an〉〈an| to commute with the apparatus-

environment interaction Hamiltonian HAE ,10

[P (A)n , HAE ] = 0 for all n. (3.21)

This implies that any correlation of the measured system(or any other system, for instance an observer) with theeigenstates of a preferred apparatus observable,

OA =∑

n

λnP(A)n , (3.22)

is preserved, and that the states of the environment re-

liably mirror the pointer states P(A)n . In this case, the

environment can be regarded as carrying out a nonde-molition measurement on the apparatus. The commu-tativity requirement, Eq. (3.21), is obviously fulfilled if

HAE is a function of OA, HAE = HAE(OA). Conversely,system-apparatus correlations in which the states of theapparatus are not eigenstates of an observable that com-

mutes with HAE will in general be rapidly destroyed bythe interaction.

Put the other way around, this implies that the envi-ronment determines through the form of the interaction

Hamiltonian HAE , a preferred apparatus observable OA,Eq. (3.22), and thereby also the states of the system that

9 For fundamental limitations on the precision of von Neumannmeasurements of operators that do not commute with a glob-ally conserved quantity, see the Wigner-Araki-Yanase theorem(Araki and Yanase, 1960; Wigner, 1952).

10 For simplicity, we assume here that the environment E interactsdirectly only with the apparatus A, but not with the system S.

are measured by the apparatus, that is, reliably recordedthrough the formation of dynamically stable quantumcorrelations. The tridecompositional uniqueness theoremthen guarantees the uniqueness of the expansion of thefinal state |ψ〉 =

∑n cn|sn〉|an〉|en〉 (where no constraints

on the cn have to be imposed) and thereby the uniquenessof the preferred pointer basis.

Other criteria similar to the commutativity require-ment, Eq. (3.21), have been suggested for the selectionof the preferred pointer basis because it turns out that inrealistic cases the simple relation of Eq. (3.21) can usuallyonly be fulfilled approximately (Zurek, 1993; Zurek et al.,1993). More general criteria, for example, have beenbased on the von Neumann entropy −Trρ2

Ψ(t) ln ρ2Ψ(t),

or the purity Trρ2Ψ(t), with the goal of finding the most

robust states or the states which become least entan-gled with the environment in the course of the evolution(Zurek, 1993, 1998, 2003b; Zurek et al., 1993). Pointerstates are obtained by extremizing the measure (i.e., min-imizing entropy, or maximizing purity, etc.) over the ini-tial state |Ψ〉 and requiring the resulting states to be ro-bust when varying the time t. Application of this methodleads to a ranking of the possible pointer states with re-spect to their “classicality,” i.e., their robustness withrespect to interaction with the environment, and thus al-lows for the selection of preferred pointer basis in termsof the “most classical” pointer states (the “predictabilitysieve”; see Zurek, 1993; Zurek et al., 1993). Althoughthe proposed criteria differ somewhat and other mean-ingful criteria are likely to be suggested in the future,it is hoped that in the macrosopic limit the resultingstable pointer states obtained from different criteria willturn out to be very similar (Zurek, 2003b). For sometoy models (in particular, for harmonic-oscillator mod-els that lead to coherent states as pointer states), thishas already been verified explicitly (see, for example,Diosi and Kiefer, 2000; Eisert, 2004; Joos et al., 2003;Kubler and Zeh, 1973; Zurek, 1993).

2. Selection of quasiclassical properties

System-environment interaction Hamiltonians fre-quently describe a scattering process of surrounding par-ticles (photons, air molecules, etc.) interacting with thesystem under study. Since the force laws describing suchprocesses typically depend on some power of distance(such as ∝ r−2 in Newton’s or Coulomb’s force law),the interaction Hamiltonian will usually commute withthe position basis, such that, according the commutativ-ity requirement of Eq. (3.21), the preferred basis will bein position space. The fact that position is frequentlythe determinate property of our experience can then beexplained by referring to the dependence of most inter-actions on distance (Zurek, 1981, 1982, 1991).

This holds, in particular, for mesoscopic and macro-scopic systems, as demonstrated, for instance, by thepioneering study of Joos and Zeh (1985), in which sur-

Page 14: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

14

rounding photons and air molecules are shown to contin-uously “measure” the spatial structure of dust particles,leading to rapid decoherence into an apparent (improper)mixture of wave packets that are sharply peaked in po-sition space. Similar results sometimes even hold for mi-croscopic systems (usually found in energy eigenstates;see below) when they occur in distinct spatial struc-tures that couple strongly to the surrounding medium.For instance, chiral molecules such as sugar are alwaysobserved to be in chirality eigenstates (left-handed andright-handed) which are superpositions of different en-ergy eigenstates (Harris and Stodolsky, 1981; Zeh, 2000).This is explained by the fact that the spatial structureof these molecules is continuously “monitored” by theenvironment, for example, through the scattering of airmolecules, which gives rise to a much stronger couplingthan could typically be achieved by a measuring devicethat was intended to measure, say, parity or energy; fur-thermore, any attempt to prepare such molecules in en-ergy eigenstates would lead to immediate decoherenceinto environmentally stable (“dynamically robust”) chi-rality eigenstates, thus selecting position as the preferredbasis.

On the other hand, it is well known that many systems,especially in the microsopic domain, are typically foundin energy eigenstates, even if the interaction Hamiltoniandepends on a different observable than energy, e.g., posi-tion. Paz and Zurek (1999) have shown that this situa-tion arises when the predominant frequencies present inthe environment are significantly lower than the intrinsicfrequencies of the system, that is, when the separationbetween the energy states of the system is greater thanthe largest energies available in the environment. Then,the environment will only be able to monitor quantitiesthat are constants of motion. In the case of nondegener-acy, this will be energy, thus leading to the environment-induced superselection of energy eigenstates for the sys-tem.

Another example of environment-induced superselec-tion that has been studied is related to the fact that onlyeigenstates of the charge operator are observed, but neversuperpositions of different charges. The existence of thecorresponding superselection rules was first only postu-lated (Wick et al., 1952, 1970), but could subsequentlybe explained in the framework of decoherence by refer-ring to the interaction of the charge with its own Coulomb(far) field, which takes the role of an “environment,” lead-ing to immediate decoherence of charge superpositionsinto an apparent mixture of charge eigenstates (Giulini,2000; Giulini et al., 1995).

In general, three different cases have typically beendistinguished (for example, in Paz and Zurek, 1999) forthe kind of pointer observable emerging from an inter-action with the environment, depending on the relative

strengths of the system’s self-Hamiltonian HS and of the

system-environment interaction Hamiltonian HSE :

(1) When the dynamics of the system are dominated

by HSE , i.e., the interaction with the environment,

the pointer states will be eigenstates of HSE (andthus typically eigenstates of position). This casecorresponds to the typical quantum measurementsetting; see, for example, the model of Zurek (1981,1982), which is outlined in Sec. III.D.2 above.

(2) When the interaction with the environment is weak

and HS dominates the evolution of the system (thatis, when the environment is “slow” in the abovesense), a case that frequently occurs in the micro-scopic domain, pointer states will arise that are en-

ergy eigenstates of HS (Paz and Zurek, 1999).

(3) In the intermediate case, when the evolution of

the system is governed by HSE and HS in roughlyequal strength, the resulting preferred states willrepresent a “compromise” between the first twocases; for instance, the frequently studied modelof quantum Brownian motion has shown the emer-gence of pointer states localized in phase space,i.e., in both position and momentum (Eisert, 2004;Joos et al., 2003; Unruh and Zurek, 1989; Zurek,2003b; Zurek et al., 1993).

3. Implications for the preferred-basis problem

The decoherence program proposes that the preferredbasis be selected by the requirement that correlationsbe preserved in spite of the interaction with the en-vironment, and thus be chosen through the form ofthe system-environment interaction Hamiltonian. Thisseems certainly reasonable, since only such “robust”states will in general be observable—and, after all, wesolely seek an explanation for our experience (see thediscussion in Sec. II.B.3). Although only particular ex-amples have been studied (for a survey and references,see, for example, Blanchard et al., 2000; Joos et al., 2003;Zurek, 2003b), the results thus far suggest that the se-lected properties are in agreement with our observation:for mesoscopic and macroscopic objects the distance-dependent scattering interaction with surrounding airmolecules, photons, etc., will in general give rise to im-mediate decoherence into spatially localized wave pack-ets and thus select position as the preferred basis. Onthe other hand, when the environment is comparably“slow,” as is frequently the case for microsopic systems,environment-induced superselection will typically yieldenergy eigenstates as the preferred states.

The clear merit of the approach of environment-induced superselection lies in the fact that the preferredbasis is not chosen in an ad hoc manner simply tomake our measurement records determinate or to matchour experience of which physical quantities are usuallyperceived as determinate (for example, position). In-stead the selection is motivated on physical, observer-freegrounds, that is, through the system-environment inter-action Hamiltonian. The vast space of possible quantum-

Page 15: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

15

mechanical superpositions is reduced so much becausethe laws governing physical interactions depend only ona few physical quantities (position, momentum, charge,and the like), and the fact that precisely these are theproperties that appear determinate to us is explained bythe dependence of the preferred basis on the form of theinteraction. The appearance of “classicality” is thereforegrounded in the structure of the physical laws—certainlya highly satisfying and reasonable approach.

The above argument in favor of the approach ofenvironment-induced superselection could, of course, beconsidered as inadequate on a fundamental level: Allphysical laws are discovered and formulated by us, sothey can contain only the determinate quantities of ourexperience. These are the only quantities we can perceiveand thus include in a physical law. Thus the derivationof determinacy from the structure of our physical lawsmight seem circular. However, we argue again that itsuffices to demand a subjective solution to the preferred-basis problem—that is, to provide an answer to the ques-tion of why we perceive only such a small subset of prop-erties as determinate, not whether there really are de-terminate properties (on an ontological level) and whatthey are (cf. the remarks in Sec. II.B.3).

We might also worry about the generality of thisapproach. One would need to show that any suchenvironment-induced superselection leads,in fact, to pre-cisely those properties that appear determinate to us.But this would require precise knowledge of the systemand the interaction Hamiltonian. For simple toy models,the relevant Hamiltonians can be written down explicitly.In more complicated and realistic cases, this will in gen-eral be very difficult, if not impossible, since the form ofthe Hamiltonian will depend on the particular system orapparatus and the monitoring environment under consid-eration, where, in addition, the environment is not onlydifficult to define precisely, but also constantly changing,uncontrollable, and, in essence, infinitely large.

But the situation is not as hopeless as it might sound,since we know that the interaction Hamiltonian will, ingeneral, be based on the set of known physical lawswhich, in turn, employ only a relatively small numberof physical quantities. So as long as we assume the sta-bility criterion and consider the set of known physicalquantities as complete, we can automatically anticipatethat the preferred basis will be a member of this set. Theremaining, yet very relevant, question is then which sub-set of these properties will be chosen in a specific physi-cal situation (for example, will the system preferably befound in an eigenstate of energy or position?), and towhat extent this will match the experimental evidence.To give an answer, one will usually need a more detailedknowledge of the interaction Hamiltonian and of its rela-tive strength with respect to the self-Hamiltonian of thesystem in order to verify the approach. Besides, as men-tioned in Sec. III.E, there exist other criteria than thecommutativity requirement, and whether they all leadto the same determinate properties is a question that

has not yet been fully explored.Finally, a fundamental conceptual difficulty of the

decoherence-based approach to the preferred-basis prob-lem is the lack of a general criterion for what defines thesystems and the “unobserved” degrees of freedom of theenvironment (see the discussion in Sec. III.A). While inmany laboratory-type situations, the division into sys-tem and environment might seem straightforward, it isnot clear a priori how quasiclassical observables can bedefined through environment-induced superselection ona larger and more general scale, when larger parts ofthe universe are considered where the split into subsys-tems is not suggested by some specific system-apparatus-surroundings setup.

To summarize, environment-induced superselection ofa preferred basis (i) proposes an explanation for why aparticular pointer basis gets chosen at all—by arguingthat it is only the pointer basis that leads to stable, andthus perceivable, records when the interaction of the ap-paratus with the environment is taken into account; and(ii) argues that the preferred basis will correspond to asubset of the set of the determinate properties of ourexperience, since the governing interaction Hamiltonianwill depend solely on these quantities. But it does not tellus precisely what the pointer basis will be in any givenphysical situation, since it will usually be hardly possibleto write down explicitly the relevant interaction Hamil-tonian in realistic cases. This also means that it willbe difficult to argue that any proposed criterion basedon the interaction with the environment will always andin all generality lead to exactly those properties that weperceive as determinate.

More work remains to be done, therefore, to fully ex-plore the general validity and applicability of the ap-proach of environment-induced superselection. But sincethe results obtained thus far from toy models have beenin promising agreement with empirical data, there is littlereason to doubt that the decoherence program has pro-posed a very valuable criterion for explaining the emer-gence of preferred states and their robustness. The factthat the approach is derived from physical principlesshould be counted additionally in its favor.

4. Pointer basis vs instantaneous Schmidt states

The so-called Schmidt basis, obtained by diagonaliz-ing the (reduced) density matrix of the system at eachinstant of time, has been frequently studied with respectto its ability to yield a preferred basis (see, for exam-ple, Albrecht, 1992, 1993; Zeh, 1973), having led some toconsider the Schmidt basis states as describing “instan-taneous pointer states” (Albrecht, 1992). However, as ithas been emphasized (for example, by Zurek, 1993), anydensity matrix is diagonal in some basis, and this ba-sis will in general not play any special interpretive role.Pointer states that are supposed to correspond to qua-siclassical stable observables must be derived from an

Page 16: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

16

explicit criterion for classicality (typically, the stabilitycriterion); the simple mathematical diagonalization pro-cedure of the instantaneous density matrix will gener-ally not suffice to determine a quasiclassical pointer basis(see the studies by Barvinsky and Kamenshchik, 1995;Kent and McElwaine, 1997).

In a more refined method, one refrains from comput-ing instantaneous Schmidt states and instead allows for acharacteristic decoherence time τD to pass during whichthe reduced density matrix decoheres (a process that canbe described by an appropriate master equation) andbecomes approximately diagonal in the stable pointerbasis, the basis that is selected by the stability crite-rion. Schmidt states are then calculated by diagonalizingthe decohered density matrix. Since decoherence usuallyleads to rapid diagonality of the reduced density matrixin the stability-selected pointer basis to a very good ap-proximation, the resulting Schmidt states are typicallyvery similar to the pointer basis except when the pointerstates are very nearly degenerate. The latter situationis readily illustrated by considering the approximatelydiagonalized decohered density matrix

ρ =

(1/2 + δ ω∗

ω 1/2 − δ

), (3.23)

where |ω| ≪ 1 (strong decoherence) and δ ≪ 1 (near-degeneracy; Albrecht, 1993). If decoherence led to exactdiagonality, ω = 0, the eigenstates would be, for any fixedvalue of δ, proportional to (0, 1) and (1, 0) (correspond-ing to the “ideal” pointer states). However, for fixedω > 0 (approximate diagonality) and δ → 0 (degener-acy), the eigenstates become proportional to (±|ω|/ω, 1),which implies that, in the case of degeneracy, the Schmidtdecomposition of the reduced density matrix can yieldpreferred states that are very different from the stablepointer states, even if the decohered, rather than the in-stantaneous, reduced density matrix is diagonalized.

In summary, it is important to emphasize that stability(or a similar criterion) is the relevant requirement forthe emergence of a preferred quasiclassical basis, whichcannot, in general, be achieved by simply diagonalizingthe instantaneous reduced density matrix. However, theeigenstates of the decohered reduced density matrix will,in many situations, approximate the quasiclassical stablepointer states well, especially when these pointer statesare sufficiently nondegenerate.

F. Envariance, quantum probabilities, and the Born rule

In the following, we shall review an interestingand promising approach introduced recently by Zurek(2003b,c, 2004a,b) that aims to explain the emergence ofquantum probabilities and to deduce the Born rule basedon a mechanism termed “environment-assisted invari-ance,” or “envariance” for short, a particular symmetryproperty of entangled quantum states. The original expo-sition of Zurek (2003b) was followed up by several articles

by other authors, who assessed the approach, pointed outmore clearly the assumptions entering into the deriva-tion, and presented variants of the proof (Barnum, 2003;Mohrhoff, 2004; Schlosshauer and Fine, 2005). An ex-panded treatment of envariance and quantum probabili-ties that addresses some of the issues discussed in thesepapers and that offers an interesting outlook on fur-ther implications of envariance can be found in Zurek(2004a). In our outline of the theory of envariance, weshall follow this most recent treatment, as it spells outthe derivation and the required assumptions more ex-plicitly and in greater detail and clarity than in Zurek’searlier (2003b; 2003c; 2004b) papers (cf. also the remarksof Schlosshauer and Fine, 2005).

We include a discussion of Zurek’s proposal here fortwo reasons. First, the derivation is based on the in-clusion of an environment E , entangled with the systemS of interest to which probabilities of measurement out-comes are to be assigned, and thus it matches well thespirit of the decoherence program. Second, and moreimportantly, despite the contributions of decoherence toexplaining the emergence of subjective classicality fromquantum mechanics, a consistent derivation of classical-ity (including a motivation for some of the axioms ofquantum mechanics, as suggested by Zurek, 2003b) re-quires the separate derivation of the Born rule. The deco-herence program relies heavily on the concept of reduceddensity matrices and the related formalism and interpre-tation of the trace operation, see Eq. (3.6), which presup-

pose Born’s rule. Therefore decoherence itself cannot beused to derive the Born rule (as was tried, for example,by Deutsch, 1999 and Zurek, 1998) since otherwise theargument would be rendered circular (Zeh, 1997; Zurek,2003b).

There have been various attempts in the past to re-place the postulate of the Born rule by a derivation.Gleason’s (1957) theorem has shown that if one im-poses the condition that for any orthonormal basis ofa given Hilbert space the sum of the probabilities asso-ciated with each basis vector must add up to one, theBorn rule is the only possibility for the calculation ofprobabilities. However, Gleason’s proof provides little in-sight into the physical meaning of the Born probabilities,and therefore various other attempts, typically based ona relative frequencies approach (i.e., on a counting ar-gument), have been made towards a derivation of theBorn rule in a no-collapse (and usually relative-state)setting (see, for example, Deutsch, 1999; DeWitt, 1971;DeWitt and Graham, 1973; Everett, 1957; Farhi et al.,1989; Geroch, 1984; Graham, 1973; Hartle, 1968). How-ever, it was pointed out that these approaches fail dueto the use of circular arguments (Barnum et al., 2000;Kent, 1990; Squires, 1990; Stein, 1984); cf. also Wallace(2003b) and Saunders (2002).

Zurek’s recently developed theory of envariance pro-vides a promising new strategy for deriving, given certainassumptions, the Born rule in a manner that avoids thecircularities of the earlier approaches. We shall outline

Page 17: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

17

the concept of envariance in the following and show howit can lead to Born’s rule.

1. Environment-assisted invariance

Zurek introduces his definition of envariance as follows.Consider a composite state |ψSE〉 (where, as usual, Srefers to the “system” and E to some “environment”) ina Hilbert space given by the tensor product HS ⊗ HE ,

and a pair of unitary transformations US = uS ⊗ IE and

UE = IS ⊗ uE acting on S and E , respectively. If |ψSE〉 is

invariant under the combined application of US and UE ,

UE(US |ψSE〉) = |ψSE〉, (3.24)

|ψSE〉 is called envariant under uS . In other words, the

change in |ψSE〉 induced by acting on S via US can be

undone by acting on E via UE . Note that envariance isa distinctly quantum feature, absent from pure classicalstates, and a consequence of quantum entanglement.

The main argument of Zurek’s derivation is based on astudy of a composite pure state in the diagonal Schmidtdecomposition

|ψSE〉 =1√2

(eiϕ1 |s1〉|e1〉 + eiϕ2 |s1〉|e1〉

), (3.25)

where the |sk〉 and |ek〉 are sets of orthonormal basisvectors that span the Hilbert spaces HS and HE , respec-tively. The case of higher-dimensional state spaces canbe treated similarly, and a generalization to expansioncoefficients of different magnitudes can be made by ap-plication of a standard counting argument (Zurek, 2003c,2004a). The Schmidt states |sk〉 are identified with theoutcomes, or “events” (Zurek, 2004b, p. 12), to whichprobabilities are to be assigned.

Zurek now states three simple assumptions, called“facts” (Zurek, 2004a, p. 4; see also the discussion inSchlosshauer and Fine, 2005):

(A1) A unitary transformation of the form · · ·⊗ IS ⊗· · ·does not alter the state of S.

(A2) All measurable properties of S, including probabil-ities of outcomes of measurements on S, are fullydetermined by the state of S.

(A3) The state of S is completely specified by the globalcomposite state vector |ψSE〉.

Given these assumptions, one can show that the stateof S and any measurable properties of S cannot be af-fected by envariant transformations. The proof goes as

follows. The effect of an envariant transformation uS⊗IEacting on |ψSE〉 can be undone by a corresponding “coun-

tertransformation” IS⊗uE that restores the original statevector |ψSE〉. Since it follows from (A1) that the lattertransformation has left the state of S unchanged, but

(A3) implies that the final state of S (after the trans-formation and countertransformation) is identical to the

initial state of S, the first transformation uS ⊗ IE cannothave altered the state of S either. Thus, using assump-tion (A2), it follows that an envariant transformation

uS⊗ IE acting on |ψSE〉 leaves any measurable propertiesof S unchanged, in particular the probabilities associatedwith outcomes of measurements performed on S.

Let us now consider two different envariant transfor-mations: A phase transformation of the form

uS(ξ1, ξ2) = eiξ1 |s1〉〈s1| + eiξ2 |s2〉〈s2| (3.26)

that changes the phases associated with the Schmidtproduct states |sk〉|ek〉 in Eq. (3.25), and a swap trans-

formation

uS(1 ↔ 2) = eiξ12 |s1〉〈s2| + eiξ21 |s2〉〈s1| (3.27)

that exchanges the pairing of the |sk〉 with the |el〉. Basedon the assumptions (A1)–(A3) mentioned above, envari-ance of |ψSE〉 under these transformations means thatmeasurable properties of S cannot depend on the phasesϕk in the Schmidt expansion of |ψSE〉, Eq. (3.25). Simi-larily, it follows that a swap uS(1 ↔ 2) leaves the stateof S unchanged, and that the consequences of the swapcannot be detected by any measurement that pertains toS alone.

2. Deducing the Born rule

Together with an additional assumption, this resultcan then be used to show that the probabilities of the“outcomes” |sk〉 appearing in the Schmidt decomposi-tion of |ψSE〉 must be equal, thus arriving at Born’s rulefor the special case of a state-vector expansion with co-efficients of equal magnitude. Zurek (2004a) offers threepossibilities for such an assumption. Here we shall limitour discussion to one of these possible assumptions (seealso the comments in Schlosshauer and Fine, 2005):

(A4) The Schmidt product states |sk〉|ek〉 appearing inthe state-vector expansion of |ψSE〉 imply a directand perfect correlation of the measurement out-comes associated with the |sk〉 and |ek〉. That is,

if an observable OS =∑skl|sk〉〈sl| is measured on

S and |sk〉 is obtained, a subsequent measurement

of OE =∑ekl|ek〉〈el| on E will yield |ek〉 with cer-

tainty (i.e., with probability equal to one).

This assumption explicitly introduces a probabilityconcept into the derivation. (Similarly, the two other pos-sible assumptions suggested by Zurek establish a connec-tion between the state of S and probabilities of outcomesof measurements on S.)

Then, denoting the probability for the outcome |sk〉 byp(|sk〉, |ψSE〉) when the composite system SE is describedby the state vector |ψSE〉, this assumption implies that

p(|sk〉; |ψSE〉) = p(|ek〉; |ψSE〉). (3.28)

Page 18: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

18

After acting on |ψSE〉 with the envariant swap transfor-

mation US = uS(1 ↔ 2) ⊗ IE [see Eq. (3.27)] and usingassumption (A4) again, we get

p(|s1〉; US |ψSE〉) = p(|e2〉; US |ψSE〉),p(|s2〉; US |ψSE〉) = p(|e1〉; US |ψSE〉).

(3.29)

Now, when a “counterswap” UE = IS ⊗ uE(1 ↔ 2) isapplied to |ψSE〉, the original state vector |ψSE〉 is re-

stored, i.e., UE(US |ψSE〉) = |ψSE〉. It then follows fromassumptions (A2) and (A3) listed above that

p(|sk〉; UEUS |ψSE〉) = p(|sk〉; |ψSE〉). (3.30)

Furthermore, assumptions (A1) and (A2) imply that thefirst and second swap cannot have affected the measur-able properties of E and S, respectively, particularly notthe probabilities for outcomes of measurements on E (S),

p(|sk〉; UE US |ψSE〉) = p(|sk〉; US |ψSE〉),p(|ek〉; US |ψSE〉) = p(|ek〉; |ψSE〉).

(3.31)

Combining Eqs. (3.28)–(3.31) yields

p(|s1〉; |ψSE〉)(3.30)= p(|s1〉; UEUS |ψSE〉)

(3.31)= p(|s1〉; US |ψSE〉)

(3.29)= p(|e2〉; US |ψSE〉)

(3.31)= p(|e2〉; |ψSE〉)

(3.28)= p(|s2〉; |ψSE〉), (3.32)

which establishes the desired result p(|s1〉; |ψSE〉) =p(|s2〉; |ψSE〉). The general case of unequal coefficients inthe Schmidt decomposition of |ψSE〉 can then be treatedby means of a simple counting method (Zurek, 2003c,2004a), leading to Born’s rule for probabilities that arerational numbers. Using a continuity argument, this re-sult can be further generalized to include probabilitiesthat cannot be expressed as rational numbers (Zurek,2004a).

3. Summary and outlook

If one grants the stated assumptions, Zurek’s devel-opment of the theory of envariance offers a novel andpromising way of deducing Born’s rule in a noncircularmanner. Compared to the relatively well-studied field ofdecoherence, envariance and its consequences have onlybegun to be explored. In this review, we have focusedon envariance in the context of a derivation of the Bornrule, but other far-reaching implications of envariancehave recently been suggested by Zurek (2004a). Forexample, envariance could also account for the emer-gence of an environment-selected preferred basis (that

is, for environment-induced superselection) without anappeal to the trace operation or to reduced density ma-trices. This could open up the possibility of a redevelop-ment of the decoherence program based on fundamentalquantum-mechanical principles that do not require one topresuppose the Born rule; this also might shed new light,for example, on the interpretation of reduced density ma-trices that has led to much controversy in discussions ofdecoherence (see Sec. III.B). As of now, the developmentof such ideas is at a very early stage, but we can expectfurther interesting results derived from envariance in thenear future.

IV. THE ROLE OF DECOHERENCE IN

INTERPRETATIONS OF QUANTUM MECHANICS

It was not until the early 1970s that the importanceof the interaction of physical systems with their envi-ronments for a realistic quantum-mechanical descriptionof these systems was realized and a proper viewpointon such interactions was established (Zeh, 1970, 1973).It took another decade for the first concise formulationof the theory of decoherence (Zurek, 1981, 1982) to beworked out and for numerical studies to be made thatshowed the ubiquity and effectiveness of decoherence ef-fects (Joos and Zeh, 1985). Of course, by that time,several interpretive approaches to quantum mechanicshad already been established, for example, Everett-stylerelative-state interpretations (Everett, 1957), the con-cept of modal interpretations introduced by van Fraassen(1973, 1991), and the pilot-wave theory of de Broglie andBohm (Bohm, 1952).

When the relevance of decoherence effects was recog-nized by (parts of) the scientific community, decoherenceprovided a motivation for a fresh look at the existing in-terpretations and for the introduction of changes and ex-tensions to these interpretations, as well as for new inter-pretations. Some of the central questions in this contextwere, and still are, the following:

1. Can decoherence by itself solve certain foundationalissues at least for all practical purposes, such asto make certain interpretive additives superfluous?What, then, are the crucial remaining foundationalproblems?

2. Can decoherence protect an interpretation fromempirical disproof?

3. Conversely, can decoherence provide a mechanismto exclude an interpretive strategy as incompati-ble with quantum mechanics and/or as empiricallyinadequate?

4. Can decoherence physically motivate some of theassumptions on which an interpretation is basedand give them a more precise meaning?

Page 19: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

19

5. Can decoherence serve as an amalgam that wouldunify and simplify a spectrum of different interpre-tations?

These and other questions have been widely discussed,both in the context of particular interpretations andwith respect to the general implications of decoherencefor any interpretation of quantum mechanics. In par-ticular, interpretations that uphold the universal va-lidity of the unitary Schrodinger time evolution, mostnotably relative-state and modal interpretations, havefrequently incorporated environment-induced superselec-tion of a preferred basis and decoherence into their frame-work. It is the purpose of this section to critically inves-tigate the implications of decoherence for the existing in-terpretations of quantum mechanics, with particular at-tention to the questions outlined above.

A. General implications of decoherence for interpretations

When measurements are understood as ubiquitous in-teractions that lead to the formation of quantum correla-tions, the selection of a preferred basis becomes in mostcases a fundamental requirement. This also corresponds,in general, to the question of what properties are beingascribed to systems (or worlds, minds, etc.). Thus thepreferred-basis problem is at the heart of any interpreta-tion of quantum mechanics. Some of the difficulties thatmust be faced in solving the preferred-basis problem are

(i) to decide whether the selection of any preferred ba-sis (or quantity or property) is justified at all oronly an artefact of our subjective experience;

(ii) if we decide on (i) in the positive, to select thosedeterminate quantity or quantities (what appearsdeterminate to us does not need to be appear de-terminate to other kinds of observers, nor does itneed to be the “true” determinate property);

(iii) to avoid any ad hoc character of the choice and anypossible empirical inadequacy or inconsistency withthe confirmed predictions of quantum mechanics;

(iv) if a multitude of quantities is selected that applydifferently among different systems, to be able toformulate specific rules that specify the determi-nate quantity or quantities under every circum-stance;

(v) to ensure that the basis is chosen such that if thesystem is embedded in a larger (composite) system,the principle of property composition holds, i.e., theproperty selected by the basis of the original systemshould also persist when the system is consideredas part of a larger composite system.11

11 This is a problem encountered in some modal interpretations (see

The hope is then that environment-induced super-selection of a preferred basis can provide a universalmechanism that fulfills the above criteria and solves thepreferred-basis problem on strictly physical grounds.

A popular reading of the decoherence program typi-cally goes as follows. First, the interaction of the systemwith the environment selects a preferred basis, i.e., a par-ticular set of quasiclassical robust states that commute,at least approximately, with the Hamiltonian governingthe system–environment interaction. Since the form ofthe interaction Hamiltonians usually depends on familiar“classical” quantities, the preferred states will typicallyalso correspond to the small set of “classical” properties.Decoherence then quickly damps superpositions betweenthe localized preferred states when only the system is con-sidered. This is taken as an explanation of the appear-ance to a local observer of a “classical” world of determi-nate, “objective” (in the sense of being robust) proper-ties. The tempting interpretation of these achievementsis then to conclude that this accounts for the observationof unique (via environment-induced superselection) anddefinite (via decoherence) pointer states at the end of themeasurement, and the measurement problem appears tobe solved, at least for all practical purposes.

However, the crucial difficulty in the above reasoningis justifying the second step: How is one to interpret thelocal suppression of interference in spite of the fact thatfull coherence is retained in the total state that describesthe system-environment combination? While the localdestruction of interference allows one to infer the emer-gence of an (improper) ensemble of individually localizedcomponents of the wave function, one still needs to im-pose an interpretive framework that explains why onlyone of the localized states is realized and/or perceived.This has been done in various interpretations of quan-tum mechanics, typically on the basis of the decoheredreduced density matrix to ensure consistency with thepredictions of the Schrodinger dynamics and thus empir-ical adequacy.

In this context, one might raise the question whetherretention of full coherence in the composite state ofthe system-environment combination could ever lead toempirical conflicts with the ascription of definite val-ues to (mesoscopic and macroscopic) systems in somedecoherence-based interpretive approach. After all, onecould think of enlarging the system so as to include theenvironment in such a way that measurements could nowactually reveal the persisting quantum coherence even ona macroscopic level. However, Zurek (1982) asserted thatsuch measurements would be impossible to carry out inpractice, a statement that was supported by a simplemodel calculation by Omnes (1992, p. 356) for a bodywith a macrosopic number (1024) of degrees of freedom.

Clifton, 1996).

Page 20: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

20

B. The standard and the Copenhagen interpretation

As is well known, the standard interpretation (“ortho-dox” quantum mechanics) postulates that every measure-ment induces a discontinuous break in the unitary timeevolution of the state through the collapse of the totalwave function onto one of its terms in the state-vectorexpansion (uniquely determined by the eigenbasis of themeasured observable), which selects a single term in thesuperposition as representing the outcome. The natureof the collapse is not at all explained, and thus the defi-nition of measurement remains unclear. Macroscopic su-perpositions are not a priori forbidden, but are neverobserved since any observation would entail a measure-mentlike interaction. In the following, we shall also con-sider a “Copenhagen” variant of the standard interpreta-tion, which adds an additional key element, postulatingthe necessity of classical concepts in order to describequantum phenomena, including measurements.

1. The problem of definite outcomes

The interpretive rule of orthodox quantum mechanicsthat tells us when we can speak of outcomes is given bythe e-e link.12 This is an “objective” criterion since itallows us to infer the existence of a definite state of thesystem to which a value of a physical quantity can be as-cribed. Within this interpretive framework (and withoutpresuming the collapse postulate) decoherence cannotsolve the problem of outcomes: Phase coherence betweenmacroscopically different pointer states is preserved inthe state that includes the environment, and we can al-ways enlarge the system so as to include (at least parts of)the environment. In other words, the superposition of dif-ferent pointer positions still exists, coherence is only “de-localized into the larger system” (Kiefer and Joos, 1999,p. 5), that is, into the environment—or, as Joos and Zeh(1985, p. 224) put it, “the interference terms still exist,but they are not there”—and the process of decoherencecould in principle always be reversed. Therefore, if weassume the orthodox e-e link to establish the existenceof determinate values of physical quantities, decoherencecannot ensure that the measuring device actually ever isin a definite pointer state (unless, of course, the systemis initially in an eigenstate of the observable), or thatmeasurements have outcomes at all. Much of the generalcriticism directed against decoherence with respect to itsability to solve the measurement problem (at least in thecontext of the standard interpretation) has been centeredon this argument.

12 It is not particularly relevant for the subsequent discussionwhether the e-e link is assumed in its “exact” form, i.e., requir-ing the exact eigenstates of an observable, or a “fuzzy” form thatallows the ascription of definiteness based on only approximateeigenstates or on wave functions with (tiny) “tails.”

Note that, with respect to the global postmeasurementstate vector, given by the final step in Eq. (3.5), the in-teraction with the environment has only led to additionalentanglement. it has not transformed the state vector inany way, since the rapidly increasing orthogonality of thestates of the environment associated with the differentpointer positions has not influenced the state descriptionat all. In brief, the entanglement brought about by inter-action with the environment could even be considered asmaking the measurement problem worse. Bacciagaluppi(2003a, Sec. 3.2) puts it like this:

Intuitively, if the environment is carrying out,

without our intervention, lots of approximate

position measurements, then the measurement

problem ought to apply more widely, also to these

spontaneously occurring measurements. (. . . )

The state of the object and the environment

could be a superposition of zillions of very well

localised terms, each with slightly different po-

sitions, and which are collectively spread over a

macroscopic distance, even in the case of every-

day objects. (. . . ) If everything is in interaction

with everything else, everything is entangled with

everything else, and that is a worse problem than

the entanglement of measuring apparatuses with

the measured probes.

Only once we have formed the reduced pure-state den-sity matrix ρSA, Eq. (3.8), can the orthogonality of theenvironmental states have an effect; then, ρSA dynami-cally evolves into the improper ensemble ρ dSA [Eq. (3.9)].However, as pointed out in our general discussion of re-duced density matrices in Sec. III.B, the orthodox rule ofinterpreting superpositions prohibits regarding the com-ponents in the sum of Eq. (3.9) as corresponding to in-dividual well-defined quantum states.

Rather than considering the postdecoherence state ofthe system (or, more precisely, of the system-apparatuscombination SA), we can instead analyze the influenceof decoherence on the expectation values of observablespertaining to SA; after all, such expectation values arewhat local observers would measure in order to arriveat conclusions about SA. The diagonalized reduceddensity matrix, Eq. (3.9), together with the trace rela-tion, Eq. (3.6), implies that, for all practical purposes,the statistics of the system SA will be indistinguish-able from that of a proper mixture (ensemble) by anylocal observation on SA. That is, given (i) the trace

rule 〈O〉 = Tr(ρO) and (ii) the interpretation of 〈O〉 as

the expectation value of an observable O, the expecta-

tion value of any observable OSA restricted to the localsystem SA will be for all practical purposes identical tothe expectation value of this observable if SA had beenin one of the states |sn〉|an〉 (as if SA were describedby an ensemble of states). In other words, decoherencehas effectively removed any interference terms (such as|sm〉|am〉〈an|〈sn| where m 6= n) from the calculation of

the trace Tr(ρSAOSA) and thereby from the calculation

Page 21: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

21

of the expectation value 〈OSA〉. It has therefore beenclaimed that formal equivalence—i.e., the fact that de-coherence transforms the reduced density matrix into aform identical to that of a density matrix representing anensemble of pure states—yields observational equivalence

in the sense above, namely, the (local) indistiguishabilityof the expectation values derived from these two types ofdensity matrices via the trace rule.

But we must be careful in interpreting the correspon-dence between the mathematical formalism (such as thetrace rule) and the common terms employed in describ-ing “the world.” In quantum mechanics, the identifica-tion of the expression “Tr(ρA)” as the expectation valueof a quantity relies on the mathematical fact that, whenwriting out this trace, it is found to be equal to a sumover the possible outcomes of the measurement, weightedby the Born probabilities for the system to be “thrown”into a particular state corresponding to each of these out-comes in the course of a measurement. This certainlyrepresents our common-sense intuition about the mean-ing of expectation values as the sum over possible valuesthat can appear in a given measurement, multiplied bythe relative frequency of actual occurrence of these val-ues in a series of such measurements. This interpreta-tion, however, presumes (i) that measurements have out-comes, (ii) that measurements lead to definite “values,”(iii) that measurable physical quantities are identified asoperators (observables) in a Hilbert space, and (iv) thatthe modulus square of the expansion coefficients of thestate in terms of the eigenbasis of the observable can beinterpreted as representing probabilities of actual mea-surement outcomes (Born rule).

Thus decoherence brings about an apparent (and ap-proximate) mixture of states that seem, based on themodels studied, to correspond well to those states that weperceive as determinate. Moreover, our observation tellsus that this apparent mixture indeed looks like a properensemble in a measurement situation, as we observe thatmeasurements lead to the “realization” of precisely onestate in the “ensemble.” But within the framework of theorthodox interpretation, decoherence cannot explain thiscrucial step from an apparent mixture to the existenceand/or perception of single outcomes.

2. Observables, measurements, and environment-induced

superselection

In the standard and Copenhagen interpretationS,property ascription is determined by an observable thatrepresents the measurement of a physical quantity andthat in turn defines the preferred basis. However, anyHermitian operator can play the role of an observable,and thus any given state has the potential for an in-finite number of different properties whose attributionis usually mutually exclusive unless the correspondingobservables commute (in which case they share a com-mon eigenbasis which preserves the uniqueness of the pre-

ferred basis). What then determines the observable thatis being measured? As our discussion in Sec. II.C hasdemonstrated, the derivation of the measured observablefrom the particular form of a given state-vector expan-sion can lead to paradoxial results since this expansionis in general nonunique, so the observable must be cho-sen by other means. In the standard and Copenhageninterpretations, it is essentially the “user” who simply“chooses” the particular observable to be measured andthus determines which properties the system possesses.

This positivist point of view has, of course, led to alot of controversy, since it runs counter to the attempt toestablish an observer-independent reality that has beenthe central pursuit of natural science since its beginning.Moreover, in practice, one certainly does not have thefreedom to choose any arbitrary observable and mea-sure it; instead, one has “instruments” (including one’ssenses) that are designed to measure a particular observ-able. For most (and maybe all) practical purposes, thiswill ultimately boil down to a single relevant observable,namely, position. But what, then, makes the instrumentsdesigned for such a particular observable?

Answering this crucial question essentially meansabandoning the orthodox view of treating measurementsas a “black box” process that has little, if any, relationto the workings of actual physical measurements (wheremeasurements can here be understood in the broadestsense of a “monitoring” of the state of the system). Thefirst key point, the formalization of measurements as aformation of quantum correlations between system andapparatus, goes back to the early years of quantum me-chanics and is reflected in the measurement scheme ofvon von Neumann (1932), but it does not resolve theissue of how the choice of observables is made. Thesecond key point, the explicit inclusion of the environ-ment in a description of the measurement process, wasbrought into quantum theory by the studies of deco-herence. Zurek’s (1981) stability criterion discussed inSec. III.E has shown that measurements must be of sucha nature as to establish stable records, where stabilityis to be understood as preserving the system-apparatuscorrelations in spite of the inevitable interaction with thesurrounding environment. The “user” cannot choose theobservables arbitrarily, but must design a measuring de-vice whose interaction with the environment is such as toensure stable records in the sense above (which, in turn,defines a measuring device for this observable). In thereading of orthodox quantum mechanics, this can be in-terpreted as the environment determining the propertiesof the system.

In this sense, the decoherence program has embeddedthe rather formal concept of measurement as proposedby the standard and Copenhagen interpretations—withits vague notion of observables that are seemingly freelychosen by the observer—in a more realistic and physicalframework. This is accomplished via the specification ofobserver-free criteria for the selection of the measured ob-servable through the physical structure of the measuring

Page 22: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

22

device and its interaction with the environment, whichis, in most cases, needed to amplify the measurementrecord and thereby to make it accessible to the externalobserver.

3. The concept of classicality in the Copenhagen interpretation

The Copenhagen interpretation additionally postu-lates that classicality is not to be derived from quan-tum mechanics, for example, as the macroscopic limitof an underlying quantum structure (as is in some senseassumed, but not explicitely derived, in the standard in-terpretation), but instead that it be viewed as an indis-pensable and irreducible element of a complete quantumtheory—and, in fact, be considered as a concept prior toquantum theory. In particular, the Copenhagen inter-pretation assumes the existence of macroscopic measure-ment apparatuses that obey classical physics and thatare not supposed to be described in quantum mechani-cal terms (in sharp contrast to the von Neumann mea-surement scheme, which rather belongs to the standardinterpretation); such a classical apparatus is considerednecessary in order to make quantum-mechanical phenom-ena accessible to us in terms of the “classical” world ofour experience. This strict dualism between the systemS, to be described by quantum mechanics, and the appa-ratus A, obeying classical physics, also entails the exis-tence of an essentially fixed boundary between S and A,which separates the microworld from the macroworld (the“Heisenberg cut”). This boundary cannot be moved sig-nificantly without destroying the observed phenomenon(i.e., the full interacting compound SA).

Especially in the light of the insights gained from de-coherence it seems impossible to uphold the notion of afixed quantum–classical boundary on a fundamental levelof the theory. Environment-induced superselection andsuppression of interference have demonstrated how qua-siclassical robust states can emerge, or remain absent,using the quantum formalism alone and over a broadrange of microscopic to macroscopic scales, and have es-tablished the notion that the boundary between S and Ais to a large extent movable towards A. Similar resultshave been obtained from the general study of quantumnondemolition measurements (see, for example, Chap. 19of Auletta, 2000) which include the monitoring of a sys-tem by its environment. Also note that since the appa-ratus is described in classical terms, it is macroscopic bydefinition; but not every apparatus must be macrosopic:the actual “instrument” could well be microscopic. Onlythe “amplifier” must be macrosopic. As an example, con-sider Zurek’s (1981) toy model of decoherence, outlinedin Sec. III.D.2, in which the instrument can be repre-sented by a bistable atom while the environment playsthe role of the amplifier; a more realistic example is amacrosopic detector of gravitational waves that is treatedas a quantum-mechanical harmonic oscillator.

Based on the progress already achieved by the decoher-

ence program, it is reasonable to anticipate that decoher-ence embedded in some additional interpretive structurecould lead to a complete and consistent derivation of theclassical world from quantum mechanical principles. Thiswould make the assumption of an intrinsically classicalapparatus (which has to be treated outside of the realmof quantum mechanics) appear as neither a necessary nora viable postulate. Bacciagaluppi (2003b, p. 22) refers tothis strategy as “having Bohr’s cake and eating it”: ac-knowledging the correctness of Bohr’s notion of the ne-cessity of a classical world (“having Bohr’s cake”), butbeing able to view the classical world as part of and asemerging from a purely quantum-mechanical world.

C. Relative-state interpretations

Everett’s original (1957) proposal of a relative-stateinterpretation of quantum mechanics has motivated sev-eral strands of interpretation, presumably owing to thefact that Everett himself never clearly spelled out howhis theory was supposed to work. The system-observerduality of orthodox quantum mechanics introduces intothe theory external “observers” who are not described bythe deterministic laws of quantum systems but insteadfollow a stochastic indeterminism. This approach obvi-ously runs into problems when the universe as a wholeis considered: by definition, there cannot be any exter-nal observers. The central idea of Everett’s proposal isthen to abandon duality and instead (i) to assume theexistence of a total state |Ψ〉 representing the state ofthe entire universe and (ii) to uphold the universal valid-ity of the Schrodinger evolution, while (iii) postulatingthat all terms in the superposition of the total state atthe completion of the measurement actually correspondto physical states. Each such physical state can be un-derstood as relative (a) to the state of the other part inthe composite system (as in Everett’s original proposal;also see Mermin, 1998a; Rovelli, 1996), (b) to a particular“branch” of a constantly “splitting” universe (the many-

worlds interpretations, popularized by DeWitt, 1970 andDeutsch, 1985), or (c) to a particular “mind” in the setof minds of the conscious observer (the many-minds in-

terpretation; see, for example, Lockwood, 1996). In otherwords, every term in the final-state superposition can beviewed as representing an equally “real” physical state ofaffairs that is realized in a different “branch of reality.”

Decoherence adherents have typically been inclinedtowards relative-state interpretations (for instance Zeh,1970, 1973, 1993; Zurek, 1998), presumably because theEverett approach takes unitary quantum mechanics es-sentially “as is,” with a minimum of added interpretiveelements. This matches well the spirit of the decoher-ence program, which attempts to explain the emergenceof classicality purely from the formalism of basic quan-tum mechanics. It may also seem natural to identify thedecohering components of the wave function with differ-ent Everett branches. Conversely, proponents of relative-

Page 23: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

23

state interpretations have frequently employed the mech-anism of decoherence in solving the difficulties associ-ated with this class of interpretations (see, for example,Deutsch, 1985, 1996, 2002; Saunders, 1995, 1997, 1998;Vaidman, 1998; Wallace, 2002, 2003a).

There are many different readings and versions ofrelative-state interpretations, especially with respect towhat defines the “branches,” “worlds,” and “minds”;whether we deal with one, a multitude, or an infinityof such worlds and minds; and whether there is an ac-tual (physical) or only perspectival splitting of the worldsand minds into different branches corresponding to theterms in the superposition. Does the world or mind splitinto two separate copies (thus somehow doubling all thematter contained in the orginal system), or is there justa “reassignment” of states to a multitude of worlds orminds of constant (typically infinite) number, or is thereonly one physically existing world or mind in which eachbranch corresponds to different “aspects” (whatever theyare). Regardless, in the following discussion of the keyimplications of decoherence, the precise details and dif-ferences of these various strands of interpretation will,for the most part, be largely irrelevant.

Relative-state interpretations face two core difficulties.First, the preferred-basis problem: If states are only rel-ative, the question arises, relative to what? What deter-mines the particular basis terms that are used to definethe branches, which in turn define the worlds or mindsin the next instant of time? When precisely does the“splitting” occur? Which properties are made determi-nate in each branch, and how are they connected to thedeterminate properties of our experience? Second, whatis the meaning of probabilities, since every outcome ac-tually occurs in some world or mind, and how can Born’srule be motivated in such an interpretive framework?

1. Everett branches and the preferred-basis problem

Stapp (2002, p. 1043) stated the requirement that“a many-worlds interpretation of quantum theory existsonly to the extent that the associated basis problem issolved.” In the context of relative-state interpretations,the preferred-basis problem is not only much more severethan in the orthodox interpretation, but also more fun-damental for several reasons: (i) The branching occurscontinuously and essentially everywhere; in the generalsense of measurements understood as the formation ofquantum correlations, every newly formed correlation,whether it pertains to microscopic or macroscopic sys-tems, corresponds to a branching. (ii) The ontologicalimplications are much more drastic, at least in thoserelative-state interpretations, which assume an actual“splitting” of worlds or minds, since the choice of thebasis determines the resulting “world” or “mind” as awhole.

The environment-based basis superselection criteriaof the decoherence program have frequently been em-

ployed to solve the preferred-basis problem of relative-state interpretations (see, for example, Butterfield, 2001;Wallace, 2002, 2003a; Zurek, 1998). There are severaladvantages in a decoherence-related approach to select-ing the preferred Everett bases: First, no a priori ex-istence of a preferred basis needs to be postulated, butinstead the preferred basis arises naturally from the phys-ical criterion of robustness. Second, the selection willbe likely to yield empirical adequacy, since the decoher-ence program is derived solely from the well-confirmedSchrodinger dynamics (modulo the possibility that ro-bustness may not be the universally valid criterion).Lastly, the decohered components of the wave functionevolve in such a way that they can be reidentified overtime (forming “trajectories” in the preferred state spaces)and thus can be used to define stable, temporally ex-tended Everett branches. Similarly, such trajectories canbe used to ensure robust observer record states and/orenvironmental states that make information about thestate of the system of interest widely accessible to ob-servers (see, for example, Zurek’s “existential interpreta-tion,” outlined in Sec. IV.C.3 below).

While the idea of directly associating the environment-selected basis states with Everett worlds seems naturaland straightforward, it has also been subject to criti-cism. Stapp (2002) has argued that an Everett-typeinterpretation must aim at determining a denumerableset of distinct branches that correspond to the appar-ently discrete events of our experience. Among thesebranches one must be able to assign determinate valuesand finite probabilities according to the usual rules andtherefore one would need to be able to specify a denu-merable set of mutually orthogonal projection operators.It is well known, however (Zurek, 1998), that the pre-ferred states chosen through the interaction with the en-vironment via the stability criterion frequently form anovercomplete set of states—often a continuum of narrowGaussian-type wave packets, for example, the coherentstates of harmonic-oscillator models that are not neces-sarily orthogonal (i.e., the Gaussians may overlap; seeKubler and Zeh, 1973; Zurek et al., 1993). Stapp there-fore considers this approach to the preferred-basis prob-lem in relative-state interpretations to be unsatisfactory.Zurek (2003a) has rebutted this criticism by pointing outthat a collection of harmonic oscillators that would leadto such overcomplete sets of Gaussians cannot serve as anadequate model of the human brain, and it is ultimatelyonly in the brain where the perception of denumerabilityand mutual exclusiveness of events must be accountedfor (see Sec. II.B.3); when neurons are more appropri-ately modeled as two-state systems, the issue raised byStapp disappears (for a discussion of decoherence in asimple two-state model, see Sec. III.D.2).13

13 For interesting quantitative results on the role of decoherence inneuronal processes, see Tegmark (2000).

Page 24: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

24

The approach of using environment-induced superse-lection and decoherence to define the Everett brancheshas also been critized on grounds of being “conceptuallyapproximate,” since the stability criterion generally leadsonly to an approximate specification of a preferred basisand therefore cannot give an “exact” definition of the Ev-erett branches (see, for example, the comments of Kent,1990; Zeh, 1973, and also the well-known “anti-FAPP”position of Bell, 1982). Wallace (2003a, pp. 90–91) hasargued against such an objection as

(. . . ) arising from a view implicit in much dis-

cussion of Everett-style interpretations: that cer-

tain concepts and objects in quantum mechan-

ics must either enter the theory formally in its

axiomatic structure, or be regarded as illusion.

(. . . ) [Instead] the emergence of a classical world

from quantum mechanics is to be understood in

terms of the emergence from the theory of cer-

tain sorts of structures and patterns, and . . . this

means that we have no need (as well as no hope!)

of the precision which Kent [in his (1990) critique]

and others (. . . ) demand.

Accordingly, in view of our argument in Sec. II.B.3for considering subjective solutions to the measurementproblem as sufficient, there is no a priori reason to doubtthat an “approximate” criterion for the selection of thepreferred basis can give a meaningful definition of theEverett branches—one that is empirically adequate andthat accounts for our experiences. The environment-superselected basis emerges naturally from the physicallyvery reasonable criterion of robustness together with thepurely quantum mechanical effect of decoherence. Itwould be rather difficult to imagine rgar ab axiomaticallyintroduced “exact” rule could be able to select preferredbases in a manner that is similarly physically motivatedand capable of ensuring empirical adequacy.

Besides using the environment-superselected pointerstates to describe the Everett branches, various authorshave directly used the instantaneous Schmidt decompo-sition of the composite state (or, equivalently, the set oforthogonal eigenstates of the reduced density matrix) todefine the preferred basis (see also Sec. III.E.4). Thisapproach is easier to implement than the explicit searchfor dynamically stable pointer states since the preferredbasis follows directly from a simple mathematical diag-onalization procedure at each instant of time. Further-more, it has been favored by some (e.g., Zeh, 1973) sinceit gives an “exact” rule for basis selection in relative-state interpretations; the consistently quantum originof the Schmidt decomposition, which matches well the“pure quantum-mechanics” spirit of Everett’s proposal(where the formalism of quantum mechanics supplies itsown interpretation), has also been counted as an advan-tage (Barvinsky and Kamenshchik, 1995). In an earlierwork, Deutsch (1985) attributed a fundamental role tothe Schmidt decomposition in relative-state interpreta-tions as defining an “interpretation basis” that imposes

the precise structure that is needed to give meaning toEverett’s basic concept.

However, as pointed out in Sec. III.E.4, emerging ba-sis states based on the instantaneous Schmidt stateswill frequently have properties that are very differentfrom those selected by the stability criterion and thatare undesirably nonclassical. For example, they maylack the spatial localization of the robustness-selectedGaussians (Stapp, 2002). The question to what ex-tent the Schmidt basis states correspond to classicalproperties in Everett-style interpretations was investi-gated in detail by Barvinsky and Kamenshchik (1995).The authors study the similarity of the states selectedby the Schmidt decomposition to coherent states (i.e.,minimum-uncertainty Gaussians) that are chosen asthe “yardstick states” representing classicality (see alsoEisert, 2004). For the investigated toy models it is foundthat only subsets of the Everett worlds correspondingto the Schmidt decomposition exhibit classicality in thissense; furthermore, the degree of robustness of classical-ity in these branches is very sensitive to the choice ofthe initial state and the interaction Hamiltonian, suchthat classicality emerges typically only temporarily, andthe Schmidt basis generally lacks robustness under timeevolution. Similar difficulties with the Schmidt basisapproach have been described by Kent and McElwaine(1997).

These findings indicate that a selection criterion basedon robustness provides a much more meaningful, phys-ically transparent, and general rule for the selection ofquasiclassical branches in relative-state interpretations,especially with respect to its ability to lead to wave-function components representing quasiclassical proper-ties that can be reidentified over time (which a simplediagonalization of the reduced density matrix at each in-stant of time does not, in general, allow for).

2. Probabilities in Everett interpretations

Various attempts unrelated to decoherence have beenmade to find a consistent derivation of the Born probabil-ities (for instance, Deutsch, 1999; DeWitt, 1971; Everett,1957; Geroch, 1984; Graham, 1973; Hartle, 1968) in theexplicit or implicit context of a relative-state interpre-tation, but several arguments have been presented thatshow that these approaches fail.14 When the effects ofdecoherence and environment-induced superselection areincluded, it seems natural to identify the diagonal ele-ments of the decohered reduced density matrix (in theenvironment-superselected basis) with the set of possible

14 See, for example, the critiques of Barnum et al. (2000); Kent(1990); Squires (1990); Stein (1984); however, also note the ar-guments of Wallace (2003b) and Gill (2003), defending the ap-proach of Deutsch (1999); see also Saunders (2002).

Page 25: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

25

elementary events and to interpret the corresponding co-efficients as relative frequencies of worlds (or minds, etc.)in the Everett theory, assuming a typically infinite mul-titude of worlds, minds, etc. Since decoherence enablesone to reidentify the individual localized components ofthe wave function over time (describing, for example, ob-servers and their measurement outcomes attached to in-dividual well-defined “worlds”), this leads to a naturalinterpretation of the Born probabilities as empirical fre-quencies.

However, decoherence cannot yield an actual deriva-tion of the Born rule (for attempts in this direction, seeDeutsch, 1999; Zurek, 1998). As mentioned before, thisis so because the key elements of the decoherence pro-gram, the formalism and the interpretation of reduceddensity matrices and the trace rule, presume the Bornrule. Attempts to consistently derive probabilities fromreduced density matrices and from the trace rule aretherefore subject to the charge of circularity (Zeh, 1997;Zurek, 2003b). In Sec. III.F, we outlined a recent pro-posal by Zurek (2003c) that evades this circularity anddeduces the Born rule from envariance, a symmetry prop-erty of entangled systems, and from certain assumptionsabout the connection between the state of the system Sof interest, the state vector of the composite system SEthat includes an environment E entangled with S, andprobabilities of outcomes of measurements performed onS. Decoherence combined with this approach provides aframework in which quantum probabilities and the Bornrule can be given a rather natural motivation, definition,and interpretation in the context of relative-state inter-pretations.

3. The “existential interpretation”

A well-known Everett-type interpretation that reliesheavily on decoherence has been proposed by Zurek(1993, 1998; see also the recent reevaluation in Zurek,2004a). This approach, termed the “existential interpre-tation,” defines the reality, or objective existence, of astate as the possibility of finding out what the state isand simultaneously leaving it unperturbed, similar to aclassical state.15 Zurek assigns a “relative objective ex-istence” to the robust states (identified with elementary“events”) selected by the environmental stability crite-rion. By measuring properties of the system-environmentinteraction Hamiltonian and employing the robustnesscriterion, the observer can, at least in principle, deter-mine the set of observables that can be measured onthe system without perturbing it and thus find out its

15 This intrinsically requires the notion of open systems, since inisolated systems, the observer would need to know in advancewhat observables commute with the state of the system, in orderto perform a nondemolition measurement that avoids repreparingthe state of the system.

“objective” state. Alternatively, the observer can takeadvantage of the redundant records of the state of thesystem as monitored by the environment. By intercept-ing parts of this environment, for example, a fractionof the surrounding photons, he can determine the stateof the system essentially without perturbing it (cf. alsothe related recent ideas of “quantum Darwinism” and therole of the environment as a “witness,” see Ollivier et al.,2003; Zurek, 2000, 2003b, 2004b).16

Zurek emphasizes the importance of stable recordsfor observers, i.e., robust correlations between theenvironment-selected states and the memory states of theobserver. Information must be represented physically,and thus the “objective” state of the observer who hasdetected one of the potential outcomes of a measurementmust be physically distinct and objectively different fromthe state of an observer who has recorded an alternativeoutcome (since the record states can be determined fromthe outside without perturbing them—see the previousparagraph). The different objective states of the ob-server are, via quantum correlations, attached to differ-ent branches defined by the environment-selected robuststates; they thus ultimately label the different branches ofthe universal state vector. This is claimed to lead to theperception of classicality; the impossibility of perceivingarbitrary superpositions is explained via the quick sup-pression of interference between different memory statesinduced by decoherence, where each (physically distinct)memory state represents an individual observer identity.

A similar argument has been given by Zeh (1993) whoemploys decoherence together with an (implicit) branch-ing process to explain the perception of definite out-comes:

[A]fter an observation one need not necessarily

conclude that only one component now exists

but only that only one component is observed.

(. . . ) Superposed world components describing

the registration of different macroscopic proper-

ties by the “same” observer are dynamically en-

tirely independent of one another: they describe

different observers. (. . . ) He who considers this

conclusion of an indeterminism or splitting of the

observer’s identity, derived from the Schrodinger

equation in the form of dynamically decoupling

(“branching”) wave packets on a fundamental

global configuration space, as unacceptable or

“extravagant” may instead dynamically formal-

ize the superfluous hypothesis of a disappearance

of the “other” components by whatever method

he prefers, but he should be aware that he may

thereby also create his own problems: Any devia-

tion from the global Schrodinger equation must in

16 The partial ignorance is necessary to avoid redefinition of thestate of the system.

Page 26: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

26

principle lead to observable effects, and it should

be recalled that none have ever been discovered.

The existential interpretation has recently been con-nected to the theory of envariance (see Zurek, 2004a,and Sec. III.F). In particular, the derivation of Born’srule based on envariance as outlined in Sec. III.F canbe recast in the framework of the existential interpreta-tion such that probabilities refer explicitly to the futurerecord state of an observer. Such a concept of proba-bility bears similarities with classical probability theory(for more details on these ideas, see Zurek, 2004a).

The existential interpretation continues Everett’s goalof interpreting quantum mechanics using the quantum-mechanical formalism itself. Zurek takes the standardno-collapse quantum theory “as is” and explores to whatextent the incorporation of environment-induced supers-election and decoherence (and recently also envariance)could form a viable interpretation that would, with aminimal additional interpretive framework, be capableof accounting for the perception of definite outcomes andof explaining the origin and nature of probabilities.

D. Modal interpretations

The first type of modal interpretation was suggestedby van Fraassen (1973, 1991), based on his program of“constructive empiricism,” which proposes to take onlyempirical adequacy, but not necessarily “truth,” as thegoal of science. Since then, a large number of interpre-tations of quantum mechanics have been suggested thatcan be considered as modal (for a review and discussionof some of the basic properties and problems of such in-terpretations, see Clifton, 1996).

In general, the approach of modal interpretations con-sists in weakening the orthodox e-e link by allowing forthe assignment of definite measurement outcomes evenif the system is not in an eigenstate of the observablerepresenting the measurement. In this way, one can pre-serve a purely unitary time evolution without the needfor an additional collapse postulate to account for definitemeasurement results. Of course, this immediately raisesthe question of how physical properties that are per-ceived through measurements and measurement resultsare connected to the state, since the bidirectional link isbroken between the eigenstate of the observable (whichcorresponds to the physical property) and the eigenvalue(which represents the manifestation of the value of thisphysical property in a measurement). The general goalof modal interpretations is then to specify rules that de-termine a catalog of possible properties of a system de-scribed by the density matrix ρ at time t. Two differentviews are typically distinguished: a semantic approachthat only changes the way of talking about the connec-tion between properties and state; and a realistic viewthat provides a different specification of what the pos-sible properties of a system really are, given the statevector (or the density matrix).

Such an attribution of possible properties must ful-fill certain requirements. For instance, probabilities foroutcomes of measurements should be consistent with theusual Born probabilities of standard quantum mechan-ics; it should be possible to recover our experience ofclassicality in the perception of macroscopic objects; andan explicit time evolution of properties and their prob-abilities should be definable that is consistent with theresults of the Schrodinger equation. As we shall see inthe following, decoherence has frequently been employedin modal interpretations to motivate and define rules forproperty ascription. Dieks (1994a,b) has argued that oneof the central goals of modal approaches is to provide aninterpretation for decoherence.

1. Property assignment based on environment-induced

superselection

The intrinsic difficulty of modal interpretations is toavoid any ad hoc character of the property assignment,yet to find generally applicable rules that lead to a selec-tion of possible properties that include the determinateproperties of our experience. To solve this problem, var-ious modal interpretations have embraced the results ofthe decoherence program. A natural approach would beto employ the environment-induced superselection of apreferred basis—since it is based on an entirely physicaland very general criterion (namely, the stability require-ment) and has, for the cases studied, been shown to giveresults that agree well with our experience, thus match-ing van Fraassen’s goal of empirical adequacy—to yieldsets of possible quasiclassical properties associated withthe correct probabilities.

Furthermore, since the decoherence program is basedsolely on Schrodinger dynamics, the task of defining atime evolution of the “property states” and their asso-ciated probabilities that is in agreement with the re-sults of unitary quantum mechanics would presumablybe easier than in a model of property assignment inwhich the set of possibilities does not arise dynami-cally via the Schrodinger equation alone (for a detailedproposal for modal dynamics of the latter type, seeBacciagaluppi and Dickson, 1999). The need for explicitdynamics of property states in modal interpretations iscontroversial. One can argue that it suffices to showthat at each instant of time, the set of possibly possessedproperties that can be ascribed to the system is empiri-cally adequate, in the sense of containing the propertiesof our experience, especially with respect to the proper-ties of macroscopic objects (this is essentially the viewof, for example, van Fraassen, 1973, 1991). On the otherhand, this cannot ensure that these properties behaveover time in agreement with our experience (for instance,that macroscopic objects that are left undisturbed donot change their position in space spontaneously in anobservable manner). In other words, the emergence ofclassicality is to be tied not only to determinate prop-

Page 27: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

27

erties at each instant of time, but also to the existenceof quasiclassical “trajectories” in property space. Sincedecoherence allows one to reidentify components of thedecohered density matrix over time, this could be usedto derive property states with continuous, quasiclassicaltrajectorylike time evolution based on Schrodinger dy-namics alone. For some discussions of this approach, seeHemmo (1996) and Bacciagaluppi and Dickson (1999).

The fact that the states emerging from decoherenceand the stability criterion are sometimes nonorthogonalor form a continuum will presumably be of even less rel-evance in modal interpretations than in Everett-style in-terpretations (see Sec. IV.C) since the goal here is solelyto specify sets of possible properties, of which only oneset actually gets assigned to the system. Hence an “over-lap” of the sets is not necessarily a problem (modulothe potential difficulty of a straightforward assignmentof probabilities in such a situation).

2. Property assignment based on instantaneous Schmidt

decompositions

Since it is usually rather difficult to determine explic-itly the robust “pointer states” through the stability (ora similar) criterion, it would not be easy to specify a gen-eral rule for property assignment based on environment-induced superselection. To simplify this situation, sev-eral modal interpretations have restricted themselves tothe orthogonal decomposition of the density matrix todefine the set of properties that can be assigned (see, forinstance, Bub, 1997; Dieks, 1989; Healey, 1989; Kochen,1985; Vermaas and Dieks, 1995). For example, the ap-proach of Dieks (1989) recognizes, by referring to thedecoherence program, the relevance of the environmentby considering a composite system-environment statevector and its diagonal Schmidt decomposition, |ψ〉 =∑

k

√pk |φSk 〉|φEk 〉, which always exists. Possible proper-

ties that can be assigned to the system are then repre-

sented by the Schmidt projectors Pk = λk|φSk 〉〈φSk |. Al-though all terms are present in the Schmidt expansion(that Dieks calls the “mathematical state”), the “phys-ical state” is postulated to be given by only one of theterms, with probability pk. A generalization of this ap-proach to a decomposition into any number of subsystemshas been described by Vermaas and Dieks (1995). In thissense, the Schmidt decomposition itself is taken to definean interpretation of quantum mechanics. Dieks (1995)suggested a physical motivation for the Schmidt decom-position in modal interpretations based on the assumedrequirement of a one-to-one correspondence between theproperties of the system and its environment. For a com-ment on the violation of the property composition prin-ciple in such interpretations, see the analysis of Clifton(1996).

A central problem associated with the approach oforthogonal decomposition is that it is not at all clearthat the properties determined by the Schmidt diago-

nalization represent the determinate properties of ourexperience. As outlined in Sec. III.E.4, the states se-lected by the (instantaneous) orthogonal decompositionof the reduced density matrix will in general differ fromthe robust “pointer states” chosen by the stability cri-terion of the decoherence program and may have dis-tinctly nonclassical properties. That this will be thecase especially when the states selected by the orthog-onal decomposition are close to degeneracy has alreadybeen indicated in Sec. III.E.4. It has also been exploredin more detail in the context of modal interpretationsby Bacciagaluppi et al. (1995) and Donald (1998), whoshowed that in the case of near degeneracy (as it typicallyoccurs for macroscopic systems with many degrees offreedom), the resulting projectors will be extremely sensi-tive to the precise form of the state (Bacciagaluppi et al.,1995). Clearly such sensitivity is undesired since the pro-jectors, and thus the properties of the system, will notbe well behaved under the inevitable approximations em-ployed in physics (Donald, 1998).

3. Property assignment based on decompositions of the

decohered density matrix

Other authors therefore have appealed to the or-thogonal decomposition of the decohered reduced den-sity matrix (instead of the decomposition of the in-stantaneous density matrix) which has led to notewor-thy results. When the system is represented by onlya finite-dimensional Hilbert space, a discrete model ofdecoherence, the resulting states were indeed found tobe typically close to the robust states selected by thestability criterion (for macroscopic systems, this typi-cally meant localization in position space), unless againthe final composite state was very nearly degenerate(Bacciagaluppi and Hemmo, 1996; Bene, 2001; see alsoSec. III.E.4). Thus, in sufficiently nondegenerate cases,decoherence can ensure that the definite properties se-lected by modal interpretations of the Dieks type will beappropriately close to the properties corresponding to theideal pointer states if the modal properties are based onthe orthogonal decomposition of the reduced decohereddensity matrix.

On the other hand, Bacciagaluppi (2000) showed thatin the more general and realistic case of an infinite-dimensional state space of the system, when one em-ploys a continuous model of decoherence (namely, thatof Joos and Zeh, 1985), the predictions of the modal ap-proach (Dieks, 1989; Vermaas and Dieks, 1995) and thoseof decoherence can differ significantly. It was demon-strated that the definite properties obtained from theorthogonal decomposition of the decohered density ma-trix were highly delocalized (that is, smeared out overthe entire spread of the state), although the coherencelength of the density matrix itself was shown to be verysmall, so that decoherence indicated very localized prop-erties. Thus, based on these results (and similar ones

Page 28: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

28

of Donald, 1998), decoherence can be used to argue forthe physical inadequacy of the rule for the assignmentof definite properties as proposed by Dieks (1989) andVermaas and Dieks (1995).

More generally, if the definite properties selected by themodal interpretation fail to mesh with the results of de-coherence (in particular, when they also lack the desiredclassicality and correspondence to the determinate prop-erties of our experience), we are given reason to doubtwhether the proposed rules for property assignment havesufficient physical motivation, legitimacy, or generality.

4. Concluding remarks

There are many different proposals that can begrouped under the heading of modal interpretations.They all share the problem of motivating and verifyinga consistent system of property assignment. Using therobust pointer states selected by interaction with the en-vironment and by the stability criterion is a step in theright direction, but the difficulty remains to derive a gen-eral rule for property assignment from this method thatwould yield explicitly the sets of possibilities in everysituation. In certain cases, for example, close to degen-eracy and in Hilbert-state spaces of infinite dimension,the simpler approach of deriving the possible propertiesfrom the orthogonal decomposition of the decohered re-duced density matrix fails to yield the sharply localized,quasiclassical pointer states as selected by environmentalrobustness criteria. These are the cases in which decoher-ence can play a vital role in helping to identify inadequaterules for property assignment in modal interpretations.

E. Physical collapse theories

The basic idea of physical collapse theories is to in-troduce an explicit modification of the Schrodinger timeevolution to achieve a physical mechanism for state-vector reduction (for an extensive recent review, seeBassi and Ghirardi, 2003). This is in general motivatedby a “realist” interpretation of the state vector, that is,the state vector is directly identified with a physical state,which then requires reduction to one of the terms in thesuperposition to establish equivalence to the observed de-terminate properties of physical states, at least as far asthe macroscopic realm is concerned.

The first proposals for theories of this type were madeby Pearle (1976, 1982, 1979) and Gisin (1984), who de-veloped dynamical reduction models that modify unitarydynamics such that a superposition of quantum statesevolves continuously into one of its terms (see also thereview by Pearle, 1999). Typically, terms represent-ing external white noise are added to the Schrodingerequation, causing the squared amplitudes |cn(t)|2 in thestate-vector expansion |Ψ(t)〉 =

∑n cn(t)|ψn〉 to fluctu-

ate randomly in time, while maintaining the normaliza-

tion condition∑

n |cn(t)|2 = 1 for all t. This processis known as stochastic dynamical reduction. Eventuallyone amplitude |cn(t)|2 → 1, while all other squared co-efficients → 0 (the “gambler’s ruin game” mechanism),where |cn(t)|2 → 1 with probability |cn(t = 0)|2 (thesquared coefficients in the initial precollapse state-vectorexpansion) in agreement with the Born probability inter-pretation of the expansion coefficients.

These early models exhibit two main difficulties. First,the preferred-basis problem: What determines the termsin the state-vector expansion into which the state vectorgets reduced? Why does reduction lead to precisely thedistinct macroscopic states of our experience and not su-perpositions thereof? Second, how can one account forthe fact that the effectiveness of collapsing superpositionsincreases when going from microscopic to macroscopicscales?

These problems motivated spontaneous localization

models, initially proposed by Ghirardi, Rimini, and We-ber (GRW; Ghirardi et al., 1986). Here state-vector re-duction is not implemented as a dynamical process (i.e.,as a continuous evolution over time), but instead oc-curs instantaneously and spontaneously, leading to aspatial localization of the wave function. To be pre-cise, the N -particle wave function ψ(x1, . . . ,xN ) is atrandom intervals multiplied by a Gaussian of the formexp

[−(X−xk)

2/2∆2]

(this process is often called a “hit”or a “jump”), and the resulting product is subsequentlynormalized. The occurrence of these hits is not explained,but rather postulated as a new fundamental physicalmechanism. Both the coordinate xk and the “center ofthe hit” X are chosen at random, but the probability fora specific X is postulated to be given by the squared innerproduct of ψ(x1, . . . ,xN ) with the Gaussian (and there-fore hits are more likely to occur where |ψ|2, viewed as afunction of xk only, is large). The mean frequency ν ofhits for a single microscopic particle is chosen so as to ef-fectively preserve unitary time evolution for microscopicsystems, while ensuring that for macroscopic objects con-taining a very large number N of particles the localiza-tion occurs rapidly (on the order of Nν), in such a way asto preclude the persistence of spatially separated macro-scopic superpositons (such as the pointer’s being in asuperpositon of “up” and “down”) on time scales shorterthan realistic observations could resolve. Ghirardi et al.

(1986) chose ν ≈ 10−16 s−1, so a macrosopic system withN ≈ 1023 particles undergoes localization on average ev-ery 10−7 s. Inevitable coupling to the environment can ingeneral be expected to lead to a further drastic increaseof N and therefore to an even higher localization rate.Note, however, that the localization process itself is in-dependent of any interaction with environment, in sharpcontrast to the decoherence approach.

Subsequently, the ideas of the stochastic dynamicalreduction and GRW theory were combined into con-

tinuous spontaneous localization models (Ghirardi et al.,1990; Pearle, 1989) in which localization of the GRW typecan be shown to emerge from a nonunitary, nonlinear Ito

Page 29: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

29

stochastic differential equation, namely, the Schrodingerequation augmented by spatially correlated Brownianmotion terms (see also Diosi, 1988, 1989). The partic-ular choice of stochastic term determines the preferredbasis. Frequently, the stochastic term has been based onthe mass density which yields a GRW-type spatial local-ization (Diosi, 1989; Ghirardi et al., 1990; Pearle, 1989),but stochastic terms driven by the Hamiltonian, leadingto a reduction on an energy basis, have also been stud-ied (Adler, 2002; Adler et al., 2001; Adler and Horwitz,2000; Bedford and Wang, 1975, 1977; Fivel, 1997;Hughston, 1996; Milburn, 1991; Percival, 1995, 1998). Ifwe focus on the first type of term, the Ghirardi-Rimini-Weber theory and continuous spontaneous localizationbecome phenomenologically similar, and we shall refer tothem jointly as “spontaneous localization” models in thefollowing discussion whenever it is unnecessary to distin-guish them explicitly.

1. The preferred-basis problem

Physical reduction theories typically remove wave-function collapse from the restrictive context of the or-thodox interpretation (where the external observer arbi-trarily selects the measured observable and thus deter-mines the preferred basis), and rather understand reduc-tion as a universal mechanism that acts constantly onevery state vector regardless of an explicit measurementsituation. In view of this it is particularly important toprovide a definition for the states into which the wavefunction collapses.

As mentioned before, the original stochastic dynamicalreduction models suffer from this preferred-basis prob-lem. Taking into account environment-induced supers-election of a preferred basis could help resolve this is-sue. Decoherence has been shown to occur, especially formesoscopic and macroscopic objects, on extremely shorttime scales, and thus would presumably be able to bringabout basis selection much faster than the time requiredfor dynamical fluctuations to establish a “winning” ex-pansion coefficient.

In contrast, the GRW theory solves the preferred-basisproblem by postulating a mechanism that leads to re-duction to a particular state vector in an expansion on aposition basis, i.e., position is assumed to be the univer-sal preferred basis. State-vector reduction then amountsto simply modifying the functional shape of the pro-jection of the state vector |ψ〉 onto the position basis〈x1, . . . ,xN |. This choice can be motivated by the in-sight that essentially all (human) observations must begrounded in a position measurement.17

17 This measurement may ultimately occur only in the brainof the observer; see the objection to the GRW model byAlbert and Vaidman (1989). With respect to the general pref-erence for position as the basis of measurements, see also the

On the one hand, the selection of position as the pre-ferred basis is supported by the decoherence program,since physical interactions frequently are governed bydistance-dependent laws. Given the stability criterionor a similar requirement, this leads to position as thepreferred observable. In this sense, decoherence providesa physical motivation for the assumption of the GRWmodel. On the other hand, it makes this assumption ap-pear as too restrictive as it cannot account for cases inwhich position is not the preferred basis—for instance,in microscopic systems where typically energy is the ro-bust observable, or in the superposition of (macroscopic)currents in SQUIDs. The GRW model simply excludessuch cases by choosing the parameters of the spontaneouslocalization process such that microscopic systems re-main generally unaffected by any state vector reduction.The basis selection approach proposed by the decoher-ence program is therefore much more general and alsoavoids the ad hoc character of the GRW theory by allow-ing for a range of preferred observables and motivatingtheir choice on physical grounds.

A similar argument can be made with respect to thecontinuous spontaneous localization approach. Here,one essentially preselects a preferred basis through theparticular choice of the stochastic terms added to theSchrodinger equation. This allows for a greater range ofpossible preferred bases, for instance by combining termsdriven by the Hamiltonian and by the mass density, lead-ing to a competition between localization in energy andposition space (corresponding to the two most frequentlyobserved eigenstates). Nonetheless, any particular choiceof terms will again be subject to the charge of possessingan ad hoc flavor, in contrast to the physical definitionof the preferred basis derived from the structure of theunmodified Hamiltonian as suggested by environment-induced selection.

2. Simultaneous presence of decoherence and spontaneous

localization

Since decoherence can be considered as an omnipresentphenomenon that has been extensively verified both theo-retically and experimentally, the assumption that a phys-ical collapse theory holds means that the evolution of asystem must be guided by both decoherence effects and

the reduction mechanism.

Let us first consider the situation in which decoherenceand the localization mechanism act constructively in thesame direction, i.e., towards a common preferred basis.This raises the question in which order these two effectsinfluence the evolution of the system (Bacciagaluppi,2003a). If localization occurs on a shorter time scale thanenvironment-induced superselection of a preferred basis

comment by Bell (1982).

Page 30: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

30

and suppression of local interference, decoherence will inmost cases have very little influence on the evolution ofthe system, since typically the system will already haveevolved into a reduced state. Conversely, if decoherenceeffects act more quickly on the system than the local-ization mechanism, the interaction with the environmentwill presumably lead to the preparation of quasiclassicalrobust states that are subsequently chosen by the local-ization mechanism. As pointed out in Sec. III.D, decoher-ence usually occurs on extremely short time scales, whichcan be shown to be significantly smaller than the actionof the spontaneous localization process for most cases(for studies related to the GRW model, see Tegmark,1993 and Benatti et al., 1995). This indicates that deco-herence will typically play an important role even in thepresence of physical wave-function reduction.

The second case occurs when decoherence leads to theselection of a different preferred basis than the reduc-tion basis specified by the localization mechanism. Asremarked by Bacciagaluppi (2003a,b) in the context ofthe GRW theory, one might then imagine the collapse ei-ther to occur only at the level of the environment (whichwould then serve as an amplifying and recording devicewith different localization properties than the system un-der study), or to lead to an explicit competition betweendecoherence and localization effects.

3. The tails problem

The clear advantage of physical collapse models overthe consideration of decoherence-induced effects alone fora solution to the measurement problem lies in the factthat an actual state reduction is achieved such that onemay be tempted to conclude that at the conclusion ofthe reduction process the system actually is in a deter-minate state. However, all collapse models achieve onlyan approximate (“for all practical purposes”) reductionof the wave function. In the case of dynamical reduc-tion models, the state will always retain small interfer-ence terms for finite times. Similarly, in the GRW the-ory the width ∆ of the multiplying Gaussian cannot bemade arbitrarily small, and therefore the reduced wavepacket cannot be made infinitely sharply localized in po-sition space, since this would entail an infinitely large en-ergy gain by the system via the time-energy uncertaintyrelation, which would certainly show up experimentally(Ghirardi et al., 1986, chose ∆ ≈ 10−5 cm). This needfor only an approximate reduction leads to wave function“tails” (Albert and Loewer, 1996), that is, in any regionin space and at any time t > 0, the wave function willremain nonzero if it has been nonzero at t = 0 (beforethe collapse), and thus there will be always a part of thesystem that is not “here.”

Physical collapse models that achieve reduction only“for all practical purposes” require a modification,namely, a weakening, of the orthodox e-e link to allowone to speak of the system’s actually being in a defi-

nite state, and thereby to ensure the objective attribu-tion of determinate properties to the system.18 In thissense, collapse models are as much “fine for all practi-cal purposes” (to paraphrase Bell, 1990) as decoherenceis, where perfect orthogonality of the environment statesis only attained as t → ∞. The severity of the conse-quences, however, is not equivalent for the two strate-gies. Since collapse models directly change the state vec-tor, a single outcome is at least approximately selected,and it only requires a “slight” weakening of the e-e linkto make this state of affairs correspond to the (objec-tive) existence of a determinate physical property. In thecase of decoherence, the lack of a precise destruction ofinterference terms is not the main problem; even if ex-act orthogonality of the environment states were ensuredat all times, the resulting reduced density matrix wouldrepresent an improper mixture, with no outcome havingbeen singled out according to the e-e link. This would bethe case regardless of whether the e-e link is expressed inthe strong or weakened form, and we would still have tosupply some additional interpretative framework to ex-plain our perception of outcomes (see also the commentby Ghirardi et al., 1987).

4. Connecting decoherence and collapse models

It was realized early that there exists a striking formalsimilarity of the equations that govern the time evolu-tion of density matrices in the GRW approach and inmodels of decoherence. For example, the GRW equationfor a single free mass point reads [Ghirardi et al., 1986,Eq. (3.5)]

i∂ρ(x, x′, t)

∂t=

1

2m

[∂2

∂x2− ∂2

∂x′2

]ρ− iΛ(x− x′)2ρ, (4.1)

where the second term on the right-hand side accountsfor the destruction of spatially separated interferenceterms. A simple model for environment-induced decoher-ence yields a very similar equation [Joos and Zeh, 1985,Eq. (3.75); see also the comment by Joos, 1987]. Thus thephysical justification for an ad hoc postulate of an explicitreduction-inducing mechanism could be questioned (ofcourse modulo the important interpretive difference be-tween the approximately proper ensembles arising fromcollapse models and the improper ensembles resultingfrom decoherence; see also Ghirardi et al., 1987). Moreconstructively, the similarity of the governing equationsmight enable one to choose the free parameters in collapsemodels on physical grounds rather than on the basis ofempirical adequacy. Conversely, this similiarity can also

18 It should be noted, however, that such “fuzzy” e-e links may inturn lead to difficulties, as the discussion of Lewis’s “countinganomaly” has shown (Lewis, 1997).

Page 31: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

31

be viewed as leading to a “protection” of physical col-lapse theories from empirical disproof. This is so becausethe inevitable and ubiquitous interaction with the envi-ronment will always, for all practical purposes of obser-vation (that is, of statistical prediction), result in (local)density matrices that are formally very similar to those ofcollapse models. What is measured is not the state vectoritself, but the probability distribution of outcomes, i.e.,values of a physical quantity and their frequency, andthis information is equivalently contained in the statevector and the density matrix. Measurements with theirintrinsically local character will presumably be unable todistinguish between the probability distribution given bythe decohered reduced density matrix and the probabilitydistribution defined by an (approximately) proper mix-ture obtained from a physical collapse. In other words, aslong as the free parameters of collapse theories are cho-sen in agreement with those determined from decoher-ence, models for state-vector reduction can be expectedto be empirically adequate since decoherence is an effectthat will be present with near certainty in every realistic(especially macroscopic) physical system.

One might of course speculate that the simultaneouspresence of both decoherence and reduction effects mightactually allow for an experimental disproof of collapsetheories by preparing states that differ in an observ-able manner from the predictions of the reduction mod-els.19 If we acknowledge the existence of interpretationsof quantum mechanics that employ only decoherence-induced suppression of interference to explain the per-ception of apparent collapses (as is, for example, claimedby the “existential interpretation” of Zurek, 1993, 1998;see Sec. IV.C.3), we will not be able to distinguish exper-imentally between a “true” collapse and a mere suppres-sion of interference as explained by decoherence. Instead,an experimental situation is required in which the col-lapse model predicts a collapse, but in which no suppres-sion of interference through decoherence arises. Again,the problem in the realization of such an experiment isthat it is very difficult to shield a system from decoher-ence effects, especially since we will typically require amesoscopic or macroscopic system in which the reductionis efficient enough to be observed. For example, based onexplicit numerical estimates, Tegmark (1993) has shownthat decoherence due to scattering of environmental par-ticles such as air molecules or photons will have a muchstronger influence than the proposed GRW effect of spon-taneous localization (see also Bassi and Ghirardi, 2003;Benatti et al., 1995; for different results for energy-drivenreduction models, cf. Adler, 2002).

19 For proposed experiments to detect the GRW collapse, see forexample Squires (1991) and Rae (1990). For experiments thatcould potentially demonstrate deviations from the predictionsof quantum theory when dynamical state-vector reduction ispresent, see Pearle (1984, 1986).

5. Summary and outlook

Decoherence has the distinct advantage of being de-rived directly from the laws of standard quantum me-chanics, whereas current collapse models are required topostulate their reduction mechanism as a new fundamen-tal law of nature. On the other hand, collapse modelsyield, at least for all practical purposes, proper mixtures,so they are capable of providing an “objective” solutionto the measurement problem. The formal similarity be-tween the time evolution equations of the collapse anddecoherence models nourishes hopes that the postulatedreduction mechanisms of collapse models could possiblybe derived from the ubiquituous and inevitable interac-tion of every physical system with its environment andthe resulting decoherence effects. We may therefore re-gard collapse models and decoherence not as mutuallyexclusive alternatives for a solution to the measurementproblem, but rather as potential candidates for a fruitfulunification. For a vague proposal along these lines, seePessoa (1998); cf. also Diosi (1989) and Pearle (1999)for speculations that quantum gravity might act as acollapse-inducing universal “environment.”

F. Bohmian mechanics

Bohm’s approach (Bohm, 1952; Bohm and Bub,1966; Bohm and Hiley, 1993) is a modification ofde Broglie’s (1930) original “pilot-wave” proposal. InBohmian mechanics, a system containing N (nonrela-tivistic) particles is described by a wave function ψ(t)and the configuration Q(t) =

(q1(t), . . . ,qN (t)

)∈ R

3N

of particle positions qi(t), i.e., the state of the system isrepresented by (ψ,Q) for each instant t. The evolution ofthe system is guided by two equations. The wave functionψ(t) is transformed as usual via the standard Schrodinger

equation, i~(∂/∂t)ψ = Hψ, while the particle positionsqi(t) of the configuration Q(t) evolve according to the“guiding equation”

dqidt

= vψi (q1, . . . ,qN ) ≡ ~

miIm

ψ∗∇qiψ

ψ∗ψ(q1, . . . ,qN ),

(4.2)where mi is the mass of the ith particle. Thus the par-ticles follow determinate trajectories described by Q(t),with the distribution of Q(t) being given by the quantumequilibrium distribution ρ = |ψ|2.

1. Particles as fundamental entities

Bohm’s theory has been critized for ascribing funda-mental ontological status to particles. General argumentsagainst particles on a fundamental level of any relativis-tic quantum theory have been frequently given (see, forinstance, Malament, 1996, and Halvorson and Clifton,

Page 32: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

32

2002).20 Moreover, and this is the point we would liketo discuss in this section, it has been argued that the ap-pearance of particles (“discontinuities in space”) could bederived from the continuous process of decoherence, lead-ing to claims that no fundamental role need be attributedto particles (Zeh, 1993, 1999, 2003). Based on decohereddensity matrices of mesoscopic and macroscopic systemsthat essentially always represent quasi-ensembles of nar-row wave packets in position space, Zeh (1993, p. 190)holds that such wave packets can be viewed as represent-ing individual “particle” positions:21

All particle aspects observed in measurements of

quantum fields (like spots on a plate, tracks in

a bubble chamber, or clicks of a counter) can be

understood by taking into account this decoher-

ence of the relevant local (i.e., subsystem) density

matrix.

The first question is then whether a narrow wave packetin position space can be identified with the subjective ex-perience of a “particle.” The answer appears to be yes:our notion of “particles” hinges on the property of local-izability, i.e., the definition of a region of space Ω ∈ R

3

in which the system (that is, the support of the wavefunction) is entirely contained. Although the nature ofthe Schrodinger dynamics implies that any wave func-tion will have nonvanishing support (“tails”) outside ofany finite spatial region Ω and therefore exact localizat-ibility will never be achieved, we only need to demandapproximate localizability to account for our experienceof particle aspects.

However, to interpret the ensembles of narrow wavepackets resulting from decoherence as leading to the per-ception of individual particles, we must embed standardquantum mechanics (with decoherence) into an addi-tional interpretive framework that explains why only oneof the wavepackets is perceived;22 that is, we do needto add some interpretive rule to get from the improperensemble emerging from decoherence to the perceptionof individual terms, so decoherence alone does not nec-essarily make Bohm’s particle concept superfluous. Butit suggests that the postulate of particles as fundamentalentities could be unnecessary, and taken together withthe difficulties in reconciling such a particle theory with

20 On the other hand, there are proposals for a “Bohmian mechan-ics of quantum fields,” i.e., a theory that embeds quantum fieldtheory into a Bohmian-style framework (Durr et al., 2003, 2004).

21 Schrodinger (1926) had made an attempt into a similar directionbut had failed since the Schrodinger equation tends to continu-ously spread out any localized wavepacket when it is consideredas describing an isolated system. The inclusion of an interact-ing environment and thus decoherence counteracts the spreadand opens up the possibility of maintaining narrow wave packetsover time (Joos and Zeh, 1985).

22 Zeh himself, like Zurek (1998), adheres to an Everett-stylebranching to which distinct observers are attached (Zeh, 1993);see also the quote in Sec. IV.C.

a relativistic quantum field theory, Bohm’s a priori as-sumption of particles at a fundamental level of the theoryappears seriously challenged.

2. Bohmian trajectories and decoherence

A well-known property of Bohmian mechanics is thefact that its trajectories are often highly nonclassical (see,for example, Appleby, 1999b; Bohm and Hiley, 1993;Holland, 1993). This poses the serious problem of howBohm’s theory can explain the existence of quasiclassicaltrajectories on a macroscopic level.

Bohm and Hiley (1993) considered the scattering of abeam of environmental particles on a macroscopic sys-tem, a process that is known to give rise to decoherence(Joos and Zeh, 1985; Joos et al., 2003). The authorsdemonstrate that this scattering yields quasiclassical tra-jectories for the system. It has further been shown thatfor isolated systems, the Bohm theory will typically notgive the correct classical limit (Appleby, 1999b). It wasthus suggested that the inclusion of the environment andof the resulting decoherence effects might be helpful in re-covering quasiclassical trajectories in Bohmian mechanics(Allori, 2001; Allori et al., 2002; Allori and Zanghı, 2001;Appleby, 1999a; Sanz and Borondo, 2003; Zeh, 1999).

We mentioned before that the interaction between amacroscopic system and its environment will typicallylead to a rapid approximate diagonalization of the re-duced density matrix in position space, and thus to spa-tially localized wave packets that follow (approximately)Hamiltonian trajectories. [This observation also providesa physical motivation for the choice of position as the fun-damental preferred basis in Bohm’s theory, in agreementwith Bell’s (1982) well-known comment that “in physicsthe only observations we must consider are position ob-servations, if only the positions of instrument pointers.”]The intuitive step is then to associate these trajectorieswith the particle trajectories Q(t) of the Bohm theory.As pointed out by Bacciagaluppi (2003b), a great ad-vantage of this strategy lies in the fact that the sameapproach would allow for a recovery of both quantumand classical phenomena.

However, a careful analysis by Appleby (1999a) showedthat this decoherence-induced diagonalization in the po-sition basis alone will in general not suffice to yield quasi-classical trajectories in Bohm’s theory; only under certainadditional assumptions will processes that lead to deco-herence also give correct quasiclassical Bohmian trajecto-ries for macroscopic systems (Appleby described the ex-ample of the long-time limit of a system that has initiallybeen prepared in an energy eigenstate). Interesting re-sults were also reported by Allori and co-workers (Allori,2001; Allori et al., 2002; Allori and Zanghı, 2001). Theydemonstrated that decoherence effects can play the roleof preserving classical properties of Bohmian trajecto-ries. Furthermore, they showed that while in standardquantum mechanics it is important to maintain narrow

Page 33: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

33

wave packets to account for the emergence of classicality,the Bohmian description of a system by both its wavefunction and its configuration allows for the derivationof quasiclassical behavior from highly delocalized wavefunctions. Sanz and Borondo (2003) studied the double-slit experiment in the framework of Bohmian mechan-ics and in the presence of decoherence and showed thateven when coherence is fully lost, and thus interference isabsent, nonlocal quantum correlations remain that influ-ence the dynamics of the particles in the Bohm theory,demonstrating that in this example decoherence does notsuffice to achieve the classical limit in Bohmian mechan-ics.

In conclusion, while the basic idea of employingdecoherence-related processes to yield the correct classi-cal limit of Bohmian trajectories seems reasonable, manydetails of this approach still need to be worked out.

G. Consistent histories interpretations

The consistent- (or decoherent-) histories approachwas introduced by Griffiths (1984, 1993, 1996) and fur-ther developed by Omnes (1988a,b,c, 1990, 1992, 1994,2002), Gell-Mann and Hartle (1990, 1991a, 1993, 1991b),Dowker and Halliwell (1992), and others. Reviews ofthe program can be found in the papers by Omnes(1992) and Halliwell (1993, 1996), as well as in therecent book by Griffiths (2002). Thoughtful critiquesinvestigating key features and assumptions of the ap-proach have been given, for example, by d’Espagnat(1989), Dowker and Kent (1995, 1996), Kent (1998),and Bassi and Ghirardi (1999). The basic idea of theconsistent-histories approach is to eliminate the funda-mental role of measurements in quantum mechanics, andinstead study quantum histories, defined as sequences ofevents represented by sets of time-ordered projection op-erators, and to assign probabilities to such histories. Theapproach was originally motivated by quantum cosmol-ogy, i.e., the study of the evolution of the entire universe,which, by definition, represents a closed system. There-fore no external observer (which is, for example, an indis-pensable element of the Copenhagen interpretation) canbe invoked.

1. Definition of histories

We assume that a physical system S is described bya density matrix ρ0 at some initial time t0 and define asequence of arbitrary times t1 < t2 < · · · < tn with t1 >t0. For each time point ti in this sequence, we consider an

exhaustive set P(i) = P (i)αi

(ti) |αi = 1 · · ·mi, 1 ≤ i ≤ n,of mutually orthogonal Hermitian projection operators

P(i)αi

(ti), obeying∑

αi

P (i)αi

(ti) = 1, P (i)αi

(ti)P(i)βi

(ti) = δαi,βiP (i)αi

(ti),

(4.3)and evolving, using the Heisenberg picture, according to

P (i)αi

(t) = U †(t0, t)P(i)αi

(t0)U(t0, t), (4.4)

where U(t0, t) is the operator that dynamically propa-gates the state vector from t0 to t.

A possible, “maximally fine-grained” history is definedby the sequence of times t1 < t2 < · · · < tn and by thechoice of one projection operator in the set P(i) for eachtime point ti in the sequence, i.e., by the set

Hα = P (1)α1

(t1), P(2)α2

(t2), . . . , P(n)αn

(tn). (4.5)

We also define the set H = Hα of all possible historiesfor a given time sequence t1 < t2 < · · · < tn. The naturalinterpretation of a history Hα is then to take it as aseries of propositions of the form “the system S was, at

time ti, in a state of the subspace spanned by P(i)αi

(ti).”

Maximally fine-grained histories can be combined toform “coarse-grained” sets which assign to each timepoint ti a linear combination

Q(i)βi

(ti) =∑

αi

π(i)αiP (i)αi

(ti), π(i)αi

∈ 0, 1 (4.6)

of the original projection operators P(i)αi

(ti).

So far, the projection operators P(i)αi

(ti) chosen at acertain instant ti in time in order to form a history Hα

were independent of the choice of the projection opera-tors at earlier times t0 < t < ti in Hα. This situationwas generalized by Omnes (1988a,b,c, 1990, 1992) to in-clude “branch-dependent” histories of the form (see alsoGell-Mann and Hartle, 1993)

Hα = P (1)α1

(t1), P(2,α1)α2

(t2), . . . , P(n,α1,...,αn−1)αn

(tn).(4.7)

2. Probabilities and consistency

In standard quantum mechanics, we can always assignprobabilities to single events, represented by the eigen-

states of some projection operator P (i)(t), via the rule

p(i, t) = Tr[P (i)†(t)ρ(t0)P

(i)(t)]. (4.8)

The natural extension of this formula to the calculationof the probability p(Hα) of a history Hα is given by

p(Hα) = D(α, α), (4.9)

where the so-called decoherence functional D(α, β) is de-fined by (Gell-Mann and Hartle, 1990)

D(α, β) = Tr[P (n)αn

(tn) · · · P (1)α1

(t1)ρ0P(1)β1

(t1) · · · P (n)βn

(tn)].

(4.10)If we instead work in the Schrodinger picture, the deco-herence functional is

Page 34: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

34

D(α, β) = Tr[P (n)αn

U(tn−1, tn) · · · P (1)α1ρ(t1)P

(1)β1

· · ·U †(tn−1, tn)P(n)βn

(tn)]. (4.11)

Consider now the coarse-grained history that arises from a combination of the two maximally fine-grained historiesHα and Hβ,

Hα∨β = P (1)α1

(t1) + P(1)β1

(t1), P(2)α2

(t2) + P(2)β2

(t2), . . . , P(n)αn

(tn) + P(n)βn

(tn). (4.12)

We interpret each combined projection operator P(i)αi

(ti) + P(i)βi

(ti) as stating that, at time ti, the system was in the

range described by the union of P(i)αi

(ti) and P(i)βi

(ti). Accordingly, we would like to require that the probability for ahistory containing such a combined projection operator be equivalently calculable from the sum of the probabilities

of the two histories containing the individual projectors P(i)αi

(ti) and P(i)βi

(ti), that is,

Tr[P (n)αn

(tn) · · ·(P (i)αi

(ti) + P(i)βi

(ti))· · · P (1)

α1(t1)ρ0P

(1)α1

(t1) · · ·(P (i)αi

(ti) + P(i)βi

(ti))· · · P (n)

αn(tn)

]

!= Tr

[P (n)αn

(tn) · · · P (i)αi

(ti) · · · P (1)α1

(t1)ρ0P(1)α1

(t1) · · · P (i)αi

(ti) · · · P (n)αn

(tn)]

+ Tr[P (n)αn

(tn) · · · P (i)βi

(ti) · · · P (1)α1

(t1)ρ0P(1)α1

(t1) · · · P (i)βi

(ti) · · · P (n)αn

(tn)].

It can be easily shown that this relation holds if and only if

ReTr

[P (n)αn

(tn) · · · P (i)αi

(ti) · · · P (1)α1

(t1)ρ0P(1)α1

(t1) · · · P (i)βi

(ti) · · · P (n)αn

(tn)]

= 0 if αi 6= βi. (4.13)

Generalizing this two-projector case to the coarse-grainedhistory Hα∨β of Eq. (4.12), we arrive at the (sufficientand necessary) consistency condition for two historiesHα and Hβ (Griffiths, 1984; Omnes, 1990, 1992),

Re[D(α, β)] = δα,βD(α, α). (4.14)

If this relation is violated, the usual sum rule for calcu-lating probabilities does not apply. This situation ariseswhen quantum interference between the two combinedhistories Hα and Hβ is present. Therefore, to ensurethat the standard laws of probability theory also hold forcoarse-grained histories, the set H of possible historiesmust be consistent in the above sense.

However, Gell-Mann and Hartle (1990) have pointedout that when decoherence effects are included to modelthe emergence of classicality, it is more natural to require

D(α, β) = δα,βD(α, α). (4.15)

Condition (4.14) has often been referred to as weak de-

coherence, and Eq. (4.15) as medium decoherence (fora proposal of a criterion for strong decoherence, seeGell-Mann and Hartle, 1998). The set H of historiesis called consistent (or decoherent) when all its mem-bers Hα fulfill the consistency condition, Eqs. (4.14) or(4.15), i.e., when they can be regarded as independent(noninterfering).

3. Selection of histories and classicality

Even when the stronger consistency criterion (4.15) isimposed on the set H of possible histories, the number ofmutually incompatible consistent histories remains rela-tively large (d’Espagnat, 1989; Dowker and Kent, 1996).It is not at all clear a priori that at least some of thesehistories should represent any meaningful set of proposi-tions about the world of our observation. Even if a col-lection of such “meaningful” histories is found, it leavesopen the question how to select such histories and whichadditional criteria would need to be invoked.

Griffith’s (1984) original aim in formulating the con-sistency criterion was only to allow for a consistentdescription of sequences of events in closed quan-tum systems without running into logical contradic-tions.23 Commonly, however, consistency has alsobeen tied to the emergence of classicality. For ex-ample, the consistency criterion corresponds to thedemand for the absence of quantum interference—aproperty of classicality—between two combined histo-ries. It has become clear that most consistent histo-ries are in fact flagrantly nonclassical (Albrecht, 1993;Dowker and Kent, 1995, 1996; Gell-Mann and Hartle,1990, 1991b; Paz and Zurek, 1993; Zurek, 1993). For in-

23 However, Goldstein (1998) used a simple example to argue thatthe consistent-histories approach can lead to contradictions withrespect to a combinination of joint probabilities, even if the con-sistency criterion is imposed; see also the subsequent exchangeof letters in the February 1999 issue of Physics Today.

Page 35: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

35

stance, when the projection operators P(i)αi

(ti) are chosento be the time-evolved eigenstates of the initial densitymatrix ρ(t0), the consistency condition will automaticallybe fulfilled, yet the histories composed of these projectionoperators have been shown to result in highly nonclassicalmacroscopic superpositions when applied to standard ex-amples such as quantum measurement or Schrodinger’scat. This demonstrates that the consistency conditioncannot serve as a sufficient criterion for classicality.

4. Consistent histories of open systems

Various authors have appealed to interaction with theenvironment and the resulting decoherence effects indefining additional criteria that would select quasiclassi-cal histories and would also lead to a physical motivationfor the consistency criterion (see, for example, Albrecht,1992, 1993; Anastopoulos, 1996; Dowker and Halliwell,1992; Finkelstein, 1993; Gell-Mann and Hartle, 1990,1998; Halliwell, 2001; Paz and Zurek, 1993; Twamley,1993b; Zurek, 1993). This approach intrinsically requiresthe notion of local, open systems and the split of the uni-verse into subsystems, in contrast to the original aim ofthe consistent-histories approach to describe the evolu-tion of a single closed, undivided system (typically theentire universe). The decoherence-based studies then as-sume the usual decomposition of the total Hilbert spaceH into a space HS , corresponding to the system S, andHE of an environment E . One then describes the histo-ries of the system S by employing projection operatorsthat act only on the system, i.e., that are of the form

P(i)αi

(ti) ⊗ IE , where P(i)αi

(ti) acts only on HS and IE isthe identity operator in HE .

This raises the question of when, i.e., under which cir-cumstances, the reduced density matrix ρS = TrEρSE ofthe system S suffices to calculate the decoherence func-tional. The reduced density matrix arises from a nonuni-tary trace over E at every time point ti, whereas the de-coherence functional of Eq. (4.11) employs the full, uni-tarily evolving density matrix ρSE for all times ti < tfand only applies a nonunitary trace operation (over bothS and E) at the final time tf . Paz and Zurek (1993) haveanswered this (rather technical) question by showing thatthe decoherence functional can be expressed entirely interms of the reduced density matrix if the time evolu-tion of the reduced density matrix is independent of thecorrelations dynamically formed between the system andthe environment. A necessary (but not always sufficient)condition for this requirement to be satisfied is given bydemanding that the reduced dynamics be governed by amaster equation that is local in time.

When a “reduced” decoherence functional exists, atleast to a good approximation, i.e., when the reduceddynamics are sufficiently insensitive to the formationof system-environment correlations, the consistency ofwhole-universe histories, described by a unitarily evolv-ing density matrix ρSE and sequences of projection oper-

ators of the form P(i)αi

(ti)⊗ IE , will be directly related tothat of open-system histories, represented by a nonuni-tarily evolving reduced density matrix ρS(ti) and “re-

duced” projection operators P(i)αi

(ti) (Zurek, 1993).

5. Schmidt states vs pointer basis as projectors

The ability of the instantaneous eigenstates of thereduced density matrix (Schmidt states; see alsoSec. III.E.4) to serve as projectors for consistent histo-ries and possibly to lead to the emergence of quasiclassi-cal histories has been studied in much detail (Albrecht,1992, 1993; Kent and McElwaine, 1997; Paz and Zurek,1993; Zurek, 1993). Paz and Zurek (1993) have shown

that Schmidt projectors P(i)αi

, defined by their commuta-tivity with the evolved, path-projected reduced densitymatrix,

[P (i)αi, U(ti−1, ti)· · ·U(t1, t2)P

(1)α1ρS(t1)

× P (1)α1U †(t1, t2) · · · U †(ti−1, ti)

]= 0, (4.16)

will always give rise to an infinite number of sets of con-sistent histories (“Schmidt histories”). However, thesehistories are branchdependent [see Eq. (4.7)] and usuallyextremely unstable, since small modifications of the timesequence used for the projections (for instance by delet-ing a time point) will typically lead to drastic changesin the resulting history, indicating that Schmidt histo-ries are usually very nonclassical (Paz and Zurek, 1993;Zurek, 1993).

This situation is changed when the time points ti arechosen such that the intervals (ti+1 − ti) are larger thanthe typical decoherence time τD of the system over whichthe reduced density matrix becomes approximately di-agonal in the preferred pointer basis chosen throughenvironment-induced superselection (see also the discus-sion in Sec. III.E.4). When the resulting pointer states,rather than the instantaneous Schmidt states, are used todefine the projection operators, stable quasiclassical his-tories will typically emerge (Paz and Zurek, 1993; Zurek,1993). In this sense, it has been suggested that inter-action with the environment can provide the missing se-lection criterion that ensures the quasiclassicality of his-tories, i.e., their stability (predictability), and the corre-spondence of the projection operators (the pointer basis)to the preferred determinate quantities of our experience.

The approximate noninterference, and thus consis-tency, of histories based on local density operators (en-ergy, number, momentum, charge etc.) as quasiclassi-cal projectors (the so-called hydrodynamic observables,see Dowker and Halliwell, 1992; Gell-Mann and Hartle,1991b; Halliwell, 1998) has been attributed to the dy-namical stability exhibited by the eigenstates of the localdensity operators. This stability leads to decoherencein the corresponding basis (Halliwell, 1998, 1999). It hasbeen argued by Zurek (2003b) that this behavior and thus

Page 36: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

36

the special quasiclassical properties of hydrodynamic ob-servables can be explained by the fact that these observ-ables obey the commutativity criterion, Eq. (3.21), of theenvironment-induced superselection approach.

6. Exact vs approximate consistency

In the idealized case where the pointer states lead to anexact diagonalization of the reduced density matrix, his-tories composed of the corresponding pointer projectors

will automatically be consistent. However, under real-istic circumstances decoherence will typically lead onlyto approximate diagonality in the pointer basis. Thisimplies that the consistency criterion will not be ful-filled exactly and that hence the probability sum ruleswill only hold approximately—although usually, due tothe efficiency of decoherence, to a very good approxima-tion (Albrecht, 1992, 1993; Gell-Mann and Hartle, 1991b;Griffiths, 1984; Omnes, 1992, 1994; Paz and Zurek, 1993;Twamley, 1993b; Zurek, 1993). Hence, the consistencycriterion has been viewed both as overly restrictive, sincethe quasiclassical pointer projectors rarely obey the con-sistency equations exactly, and as insufficient, because itdoes not give rise to constraints that would single outquasiclassical histories.

A relaxation of the consistency criterion has thereforebeen suggested, leading to “approximately consistent his-tories” whose decoherence functional would be allowed tocontain nonvanishing off-diagonal terms (correspondingto a violation of the probability sum rules) as long as thenet effect of all the off-diagonal terms was “small” in thesense of remaining below the experimentally detectablelevel (see, for example, Dowker and Halliwell, 1992;Gell-Mann and Hartle, 1991b). Gell-Mann and Hartle(1991b) have even ascribed a fundamental role to suchapproximately consistent histories, a move that hassparked much controversy and has been considered asunnecessary and irrelevant by some (Dowker and Kent,1995, 1996). Indeed, if only approximate consistency isdemanded, it is difficult to regard this condition as a fun-damental concept of a physical theory, and the questionof how much consistency is required will inevitably arise.

7. Consistency and environment-induced superselection

The relationship between consistency andenvironment-induced superselection, and thereforethe connection between the decoherence functionaland the diagonalization of the reduced density matrixthrough environmental decoherence, has been investi-gated by various authors. The basic idea, promoted, forexample, by Zurek (1993) and Paz and Zurek (1993),is to suggest that if the interaction with the environ-ment leads to rapid superselection of a preferred basis,which approximately diagonalizes the local densitymatrix, coarse-grained histories defined in this basis will

automatically be (approximately) consistent.

This approach has been explored by Twamley (1993b),who carried out detailed calculations in the context ofa quantum optical model of phase-space decoherenceand compared the results with two-time projected phase-space histories of the same model system. It was foundthat when the parameters of the interacting environmentwere changed such that the degree of diagonality of thereduced density matrix in the emerging preferred pointerbasis was increased, histories in that basis also becamemore consistent. For a similar model, however, Twamley(1993a) also showed that consistency and diagonality ina pointer basis as possible criteria for the emergence ofquasiclassicality may exhibit a very different dependenceon the initial conditions.

Extensive studies on the connection between Schmidtstates, pointer states and consistent quasiclassical his-tories have also been made by Albrecht (1992, 1993),based on analytical calculations and numerical simula-tions of toy models for decoherence, including detailednumerical results on the violation of the sum rule forhistories composed of different (Schmidt and pointer)projector bases. It was demonstrated that the presenceof stable system-environment correlations (“records”),as demanded by the criterion for the selection of thepointer basis, was of great importance in making cer-tain histories consistent. The relevance of “records” forthe consistent-histories approach in ensuring the “perma-nence of the past” has also been emphasized by other au-thors, for example, by Paz and Zurek (1993) and Zurek(1993, 2003b), and in the “strong decoherence” criterionby Gell-Mann and Hartle (1998). The redundancy withwhich information about the system is recorded in theenvironment and can thus be found out by different ob-servers without measurably disturbing the system itselfhas been suggested to allow for the notion of “objec-tively existing histories,” based on environment-selectedprojectors that represent sequences of “objectively exist-ing” quasiclassical events (Paz and Zurek, 1993; Zurek,1993, 2003b, 2004b).

In general, damping of quantum coherence caused bydecoherence will necessarily lead to a loss of quantuminterference between individual histories (but not viceversa—see the discussion by Twamley, 1993b), sincethe final trace operation over the environment in thedecoherence functional will make the off-diagonal ele-ments very small due to the decoherence-induced approx-imate mutual orthogonality of the environmental states.Finkelstein (1993) has used this observation to propose anew decoherence condition that coincides with the origi-nal definition, Eqs. (4.10) and (4.11), except for restrict-ing the trace to E , rather than tracing over both S andE . It was shown that this condition not only implies theconsistency condition of Eq. (4.15), but also characterizesthose histories that decohere due to interaction with theenvironment and that lead to the formation of “records”of the state of the system in the environment.

Page 37: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

37

8. Summary and discussion

The core difficulty associated with the consistent-histories approach has been to explain the emergence ofthe classical world of our experience from the underlyingquantum nature. Initially, it was hoped that classical-ity could be derived from the consistency criterion alone.Soon, however, the status and the role of this criterionin the formalism and its proper interpretation becamerather unclear and diffuse, since exact consistency wasshown to provide neither a necessary nor a sufficient cri-terion for the selection of quasiclassical histories.

The inclusion of decoherence effects into the consis-tent histories approach, leading to the emergence of sta-ble quasiclassical pointer states, has been found to yielda highly efficient mechanism and a sensitive criterionfor singling out quasiclassical observables that simulta-neously fulfill the consistency criterion to a very goodapproximation due to the suppression of quantum coher-ence in the state of the system. The central questionis then: What is the meaning and the remaining roleof an explicit consistency criterion in the light of such“natural” mechanisms for the decoherence of histories?Can one dispose of this criterion as a key element of thefundamental theory by noting that for all “realistic” his-tories consistency will be likely to arise naturally fromenvironment-induced decoherence alone?

The answer to this question may actually depend onthe viewpoint one takes with respect to the aim of theconsistent-histories approach. As we have noted be-fore, the original goal was simply to provide a formalismin which one could, in a measurement-free context, as-sign probabilities to certain sequences of quantum eventswithout logical inconsistencies. The more recent andrather opposite aim would be to provide a formalismthat selects only a very small subset of “meaningful” qua-siclassical histories, all of which are consistent with ourworld of experience, and whose projectors can be directlyinterpreted as objective physical events.

The consideration of decoherence effects that can giverise to effective superselection of possible quasiclassical(and consistent) histories certainly falls into the lattercategory. It is interesting to note that this approach hasalso led to a departure from the original “closed systemsonly” view to the study of local open quantum systemsand thus to the decomposition of the total Hilbert spaceinto subsystems, within the consistent-histories formal-ism. Besides the fact that decoherence intrinsically re-quires the openness of systems, this move might also re-flect the insight that the notion of classicality itself canbe viewed as only arising from a conceptual division ofthe universe into parts (see the discussion in Sec. III.A).

Therefore environment-induced decoherence and su-perselection have played a remarkable role in consistent-histories interpretations: a practical one by suggesting aphysical selection mechanism for quasiclassical histories;and a conceptual one by contributing to a shift in ourview of originally rather fundamental concepts, such as

consistency, and of the aims of the consistent-historiesapproach, like the focus on description of closed systems.

V. CONCLUDING REMARKS

We have presented an extensive discussion of the roleof decoherence in the foundations of quantum mechanics,with a particular focus on the implications of decoherencefor the measurement problem in the context of variousinterpretations of quantum mechanics.

A key achievement of the decoherence program is therecognition that openness in quantum systems is impor-tant for their realistic description. The well-known phe-nomenon of quantum entanglement had already, earlyin the history of quantum mechanics, demonstrated thatcorrelations between systems can lead to “paradoxical”properties of the composite system that cannot be com-posed from the properties of the individual systems.Nonetheless, the viewpoint of classical physics that theidealization of isolated systems is necessary to arrive atan “exact description” of physical systems has influencedquantum theory for a long time. It is the great merit ofthe decoherence program to have emphasized the ubiq-uity and essential inescapability of system-environmentcorrelations and to have established the important roleof such correlations as factors in the emergence of “clas-sicality” from quantum systems. Decoherence also pro-vides a realistic physical modeling and a generalization ofthe quantum measurement process, thus enhancing the“black-box” view of measurements in the standard (“or-thodox”) interpretation and challenging the postulate offundamentally classical measuring devices in the Copen-hagen interpretation.

With respect to the preferred-basis problem of quan-tum measurement, decoherence provides a very promis-ing definition of preferred pointer states via a physicallymeaningful requirement, namely, the robustness crite-rion, and it describes methods for selecting operationallysuch states, for example, via the commutativity crite-rion or by extremizing an appropriate measure such aspurity or von Neumann entropy. In particular, the factthat macroscopic systems virtually always decohere intoposition eigenstates gives a physical explanation for whyposition is the ubiquitous determinate property of theworld of our experience.

We have argued that, within the standard interpre-tation of quantum mechanics, decoherence cannot solvethe problem of definite outcomes in quantum measure-ment: We are still left with a multitude of (albeit indi-vidually well-localized quasiclassical) components of thewave function, and we need to supplement or other-wise to interpret this situation in order to explain whyand how single outcomes are perceived. Accordingly, wehave discussed how environment-induced superselectionof quasiclassical pointer states together with the localsuppression of interference terms can be put to great usein physically motivating, or potentially disproving, rules

Page 38: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

38

and assumptions of alternative interpretive approachesthat change (or altogether abandon) the strict orthodoxeigenvalue-eigenstate link and/or modify the unitary dy-namics to account for the perception of definite outcomesand classicality in general. For example, to name just afew applications, decoherence can provide a universal cri-terion for the selection of the branches in relative-stateinterpretations and a physical argument for the noninter-ference between these branches from the point of view ofan observer; in modal interpretations, it can be used tospecify empirically adequate sets of properties that canbe ascribed to systems; in collapse models, the free pa-rameters (and possibly even the nature of the reductionmechanism itself) might be derivable from environmentalinteractions; decoherence can also assist in the selectionof quasiclassical particle trajectories in Bohmian mechan-ics, and it can serve as an efficient mechanism for singlingout quasiclassical histories in the consistent-histories ap-proach. Moreover, it has become clear that decoherencecan ensure the empirical adequacy and thus empiricalequivalence of different interpretive approaches, whichhas led some to the claim that the choice, for exam-ple, between the orthodox and the Everett interpretationbecomes “purely a matter of taste, roughly equivalentto whether one believes mathematical language or hu-man language to be more fundamental” (Tegmark, 1998,p. 855).

It is fair to say that the decoherence program sheds newlight on many foundational aspects of quantum mechan-ics. It paves a physics-based path towards motivatingsolutions to the measurement problem; it imposes con-straints on the strands of interpretations that seek such asolution and thus makes them also more and more similarto each other. Decoherence remains an ongoing field ofintense research, in both the theoretical and experimen-tal domain, and we can expect further implications forthe foundations of quantum mechanics from such studiesin the near future.

Acknowledgments

The author would like to thank A. Fine for many valu-able discussions and comments throughout the process ofwriting this article. He gratefully acknowledges thought-ful and extensive feedback on this manuscript fromS. L. Adler, V. Chaloupka, H.-D. Zeh, and W. H. Zurek.

References

Adler, S. L., 2002, J. Phys. A 35, 841.Adler, S. L., 2003, Stud. Hist. Philos. Mod. Phys. 34, 135.Adler, S. L., D. C. Brody, T. A. Brun, and L. P. Hughston,

2001, J. Phys. A 34, 8795.Adler, S. L., and L. P. Horwitz, 2000, J. Math. Phys. 41,

2485.

Albert, D., and B. Loewer, 1996, in Perspectives on Quan-tum Reality, edited by R. Clifton (Kluwer, Dordrecht, TheNetherlands), pp. 81–91.

Albert, D. Z., and L. Vaidman, 1989, Phys. Lett. A 129, 1.Albrecht, A., 1992, Phys. Rev. D 46, 5504.Albrecht, A., 1993, Phys. Rev. D 48, 3768.Allori, V., 2001, Decoherence and the Classical Limit of Quan-

tum Mechanics, Ph.D. thesis, Physics Department, Univer-sity of Genova.

Allori, V., D. Durr, S. Goldstein, and N. Zanghı, 2002, J. Opt.B: Quantum Semiclass. Opt. 4, S482.

Allori, V., and N. Zanghı, 2001, biannual IQSA Meeting, Ce-sena, Italy, 2001, eprint quant-ph/0112009.

Anastopoulos, C., 1996, Phys. Rev. E 53, 4711.Anderson, P. W., 2001, Stud. Hist. Philos. Mod. Phys. 32,

487.Appleby, D. M., 1999a, Found. Phys. 29, 1885.Appleby, D. M., 1999b, Found. Phys. 29, 1863.Araki, H., and M. M. Yanase, 1960, Phys. Rev. 120, 622.Auletta, G., 2000, Foundations and Interpretation of Quan-

tum Mechanics in the Light of a Critical-Historical Analysisof the Problems and of a Synthesis of the Results (WorldScientific, Singapore).

Bacciagaluppi, G., 2000, Found. Phys. 30, 1431.Bacciagaluppi, G., 2003a, in The Stanford Encyclo-

pedia of Philosophy, edited by E. N. Zalta, URLhttp://plato.stanford.edu/archives/win2003/entries/qm-decoherence.

Bacciagaluppi, G., 2003b, talk given at theworkshop, Quantum Mechanics on a LargeScale, Vancouver, 23 April 2003. URLwww.physics.ubc.ca/∼berciu/PHILIP/CONFERENCES/PWI03/FILES/baccia.ps.

Bacciagaluppi, G., and M. Dickson, 1999, Found. Phys. 29,1165.

Bacciagaluppi, G., M. J. Donald, and P. E. Vermaas, 1995,Helv. Phys. Acta 68, 679.

Bacciagaluppi, G., and M. Hemmo, 1996, Stud. Hist. Philos.Mod. Phys. 27, 239.

Barnum, H., 2003, eprint quant-ph/0312150.Barnum, H., C. M. Caves, J. Finkelstein, C. A. Fuchs, and

R. Schack, 2000, Proc. R. Soc. London, Ser. A 456, 1175.Barvinsky, A. O., and A. Y. Kamenshchik, 1995, Phys. Rev.

D 52, 743.Bassi, A., and G. Ghirardi, 1999, Phys. Lett. A 257, 247.Bassi, A., and G. C. Ghirardi, 2003, Phys. Rep. 379, 257.Bedford, D., and D. Wang, 1975, Nuovo Cimento Soc. Ital.

Fis., A 26, 313.Bedford, D., and D. Wang, 1977, Nuovo Cimento Soc. Ital.

Fis., A 37, 55.Bell, J. S., 1975, Helv. Phys. Acta 48, 93.Bell, J. S., 1982, Found. Phys. 12, 989.Bell, J. S., 1990, in Sixty-Two Years of Uncertainty, edited

by A. I. Miller (Plenum, New York), pp. 17–31.Benatti, F., G. C. Ghirardi, and R. Grassi, 1995, in Advances

in Quantum Phenomena, edited by E. G. Beltrametti andJ.-M. Levy-Leblond (Plenum, New York), pp. 263–279.

Bene, G., 2001, eprint quant-ph/0104112.Blanchard, P., D. Giulini, E. Joos, C. Kiefer, and I. Sta-

matescu (eds.), 2000, Decoherence: Theoretical, Experi-mental, and Conceptual Problems, Lecture Notes in PhysicsNo. 538 (Springer, Berlin).

Bohm, D., 1952, Phys. Rev. 85, 166.Bohm, D., and J. Bub, 1966, Rev. Mod. Phys. 38, 453.

Page 39: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

39

Bohm, D., and B. Hiley, 1993, The Undivided Universe (Rout-ledge, London).

de Broglie, L., 1930, An Introduction to the Study of WaveMechanics (E. PhysicaDutton and Co., New York).

Bub, J., 1997, Interpreting the Quantum World (CambridgeUniversity, Cambridge, England), 1st edition.

Butterfield, J. N., 2001, in Quantum Mechanics: ScientificPerspectives on Divine Action, edited by R. J. Russell,P. Clayton, K. Wegter-McNelly, and J. Polkinghorne (Vat-ican Observatory and The Center for Theology and theNatural Sciences, Vatican City State), p. 55, eprint quant-ph/0105052.

Caldeira, A. O., and A. J. Leggett, 1983, Physica A 121, 587.Cisnerosy, C., R. P. Martinez-y-Romeroz, H. N. Nunez-Yepe,

and A. L. Salas-Brito, 1998, Eur. J. Phys. 19, 237.Clifton, R., 1994, in Proceedings of the Symposium on the

Foundations of Modern Physics 1994, 70 Years of Mat-ter Waves, Helsinki, Finland, 13–16 June 1994, edited byK. V. Laurikainen, C. Montonen, and K. Sunnarborg (Edi-tions Frontieres, Gif-sur-Yvette, France), pp. 45–60.

Clifton, R., 1996, Br. J. Philos. Sci. 47, 371.d’Espagnat, B., 1988, Conceptual Foundations of Quantum

Mechanics, Advanced Book Classics (Perseus, Reading),2nd edition.

d’Espagnat, B., 1989, J. Stat. Phys. 56, 747.d’Espagnat, B., 2000, Phys. Lett. A 282, 133.Deutsch, D., 1985, Int. J. Theor. Phys. 24, 1.Deutsch, D., 1996, Br. J. Philos. Sci. 47, 222.Deutsch, D., 1999, Proc. R. Soc. London, Ser. A 455, 3129.Deutsch, D., 2002, Proc. R. Soc. London, Ser. A 458, 2911.DeWitt, B. S., 1970, Phys. Today 23, 30.DeWitt, B. S., 1971, in Foundations of Quantum Mechanics,

edited by B. d’Espagnat (Academic Press, New York), pp.211–262.

DeWitt, B. S., and N. Graham, 1973, The Many-Worlds In-terpretation of Quantum Mechanics (Princeton University,Princeton).

Dieks, D., 1989, Phys. Lett. A 142, 439.Dieks, D., 1994a, Phys. Rev. A 49, 2290.Dieks, D., 1994b, in Proceedings of the Symposium on

the Foundations of Modern Physics, edited by I. Busch,P. Lahti, and P. Mittelstaedt (World Scientific, Singapore),pp. 160–167.

Dieks, D., 1995, Phys. Lett. A 197, 367.Diosi, L., 1988, Phys. Lett. A 129, 419.Diosi, L., 1989, Phys. Rev. A 40, 1165.Diosi, L., and C. Kiefer, 2000, Phys. Rev. Lett. 85, 3552.Donald, M., 1998, in The Modal Interpretation of Quantum

Mechanics, edited by D. Dieks and P. Vermaas (Kluwer,Dordrecht), pp. 213–222.

Dowker, F., and J. J. Halliwell, 1992, Phys. Rev. D 46, 1580.Dowker, F., and A. Kent, 1995, Phys. Rev. Lett. 75, 3038.Dowker, F., and A. Kent, 1996, J. Stat. Phys. 82, 1575.Durr, D., S. Goldstein, R. Tumulka, and N. Zanghı, 2003, J.

Phys. A 36, 4143.Durr, D., S. Goldstein, R. Tumulka, and N. Zanghı, 2004,

Phys. Rev. Lett. 93, 090402.Einstein, A., B. Podolsky, and N. Rosen, 1935, Phys. Rev.

47, 777.Eisert, J., 2004, Phys. Rev. Lett. 92, 210401.Elby, A., and J. Bub, 1994, Phys. Rev. A 49, 4213.Everett, H., 1957, Rev. Mod. Phys. 29, 454.Farhi, E., J. Goldstone, and S. Gutmann, 1989, Ann. Phys.

(N.Y.) 192, 368.

Finkelstein, J., 1993, Phys. Rev. D 47, 5430.Fivel, D. I., 1997, eprint quant-ph/9710042.van Fraassen, B., 1973, in Contemporary Research in the

Foundations and Philosophy of Quantum Theory, editedby C. A. Hooker (Reidel, Dordrecht), pp. 180–213.

van Fraassen, B., 1991, Quantum Mechanics: An EmpiricistView (Clarendon, Oxford).

Friedman, J. R., V. Patel, W. Chen, S. K. Yolpygo, and J. E.Lukens, 2000, Nature , 43.

Furry, W. H., 1936, Phys. Rev. 49, 393.Galindo, A., A. Morales, and R. Nunez-Lagos, 1962, J. Math.

Phys. 3, 324.Gell-Mann, M., and J. Hartle, 1990, in Proceedings of the 3rd

International Symposium on the Foundations of QuantumMechanics in the Light of New Technology (Tokyo, Japan,August 1989), edited by S. Kobayashi, H. Ezawa, Y. Mu-rayama, and S. Nomura (Physical Society of Japan, Tokio),pp. 321–343.

Gell-Mann, M., and J. Hartle, 1991a, in Proceedings of the25th International Conference on High Energy Physics(Singapore, August 2–8, 1990), edited by K. K. Phua andY. Yamaguchi (World Scientific, Singapore), volume 2, pp.1303–1310.

Gell-Mann, M., and J. Hartle, 1993, Phys. Rev. D 47, 3345.Gell-Mann, M., and J. B. Hartle, 1991b, in Complexity, En-

tropy, and the Physics of Information, edited by W. H.Zurek (Addison-Wesley, Redwood City), Santa Fe Instituteof Studies in the Science of Complexity, pp. 425–458.

Gell-Mann, M., and J. B. Hartle, 1998, in Quantum ClassicalCorrespondence: The 4th Drexel Symposium on QuantumNonintegrability, edited by D. H. Feng and B. L. Hu (In-ternational Press, Cambridge, Massachussetts), pp. 3–35.

Geroch, R., 1984, Nous 18, 617.Ghirardi, G. C., P. Pearle, and A. Rimini, 1990, Phys. Rev.

A 42, 78.Ghirardi, G. C., A. Rimini, and T. Weber, 1986, Phys. Rev.

D 34, 470.Ghirardi, G. C., A. Rimini, and T. Weber, 1987, Phys. Rev.

D 36, 3287.Gill, R. D., 2003, to appear in Quantum Probability and Infi-

nite Dimensional Analysis: From Foundations to Applica-tions, edited by M. Schurmann and U. Franz (World Sci-entific, Singapore), in press, eprint quant-ph/0307188.

Gisin, N., 1984, Phys. Rev. Lett. 52, 1657.Giulini, D., 2000, Lect. Notes Phys. 559, 67.Giulini, D., C. Kiefer, and H. D. Zeh, 1995, Phys. Lett. A

199, 291.Gleason, A. M., 1957, J. Math. Mech. 6, 885.Goldstein, S., 1998, Phys. Today 51, 42.Graham, N., 1973, in The Many-Worlds Interpretation of

Quantum Mechanics, edited by B. S. DeWitt and N. Gra-ham (Princeton University, Princeton), pp. 229–253.

Griffiths, R. B., 1984, J. Stat. Phys. 36, 219.Griffiths, R. B., 1993, Found. Phys. 23, 1601.Griffiths, R. B., 1996, Phys. Rev. A 54, 2759.Griffiths, R. B., 2002, Consistent Quantum Theory (Cam-

bridge University Press, Cambridge).Halliwell, J. J., 1993, eprint gr-qc/9308005.Halliwell, J. J., 1996, Ann. N.Y. Acad. Sci. 755, 726.Halliwell, J. J., 1998, Phys. Rev. D 58, 105015.Halliwell, J. J., 1999, Phys. Rev. Lett. 83, 2481.Halliwell, J. J., 2001, in Time in Quantum Mechanics, edited

by J. G. Muga, R. S. Mayato, and I. L. Egususquiza(Springer, Berlin), pp. 153–182.

Page 40: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

40

Halvorson, H., and R. Clifton, 2002, Philos. Sci. 69, 1.Harris, R. A., and L. Stodolsky, 1981, J. Chem. Phys. 74,

2145.Hartle, J. B., 1968, Am. J. Phys. 36, 704.Healey, R. A., 1989, The Philosophy of Quantum Mechan-

ics: An Interactive Interpretation (Cambridge University,Cambridge, England/New York).

Hemmo, M., 1996, Quantum Mechanics Without Collapse:Modal Interpretations, Histories and Many Worlds, Ph.D.thesis, University of Cambridge.

Hepp, K., 1972, Helv. Phys. Acta 45, 327.Holland, P. R., 1993, The Quantum Theory of Motion (Cam-

bridge University, Cambridge, England).Hu, B. L., J. P. Paz, and Y. Zang, 1992, Phys. Rev. D 45,

2843.Hughston, L. P., 1996, Proc. R. Soc. London, Ser. A 452, 953.Joos, E., 1987, Phys. Rev. D 36, 3285.Joos, E., 2000, in Decoherence: Theoretical, Experimen-

tal, and Conceptual Problems, edited by P. Blanchard,D. Giulini, E. Joos, C. Kiefer, and I.-O. Stamatescu(Springer, Berlin), pp. 1–17.

Joos, E., and H. D. Zeh, 1985, Z. Phys. B 59, 223.Joos, E., H. D. Zeh, C. Kiefer, D. Giulini, J. Kupsch, and

I.-O. Stamatescu, 2003, Decoherence and the Appearanceof a Classical World in Quantum Theory (Springer, NewYork), 2nd edition.

Kent, A., 1990, Int. J. Mod. Phys. A 5, 1745.Kent, A., 1998, Phys. Scr. T 76, 78.Kent, A., and J. McElwaine, 1997, Phys. Rev. 55, 1703.Kiefer, C., and E. Joos, 1999, in Quantum Future: From Volta

and Como to the Present and Beyond (Proceedings of theXth Max Born Symposium Held in Przesieka, Poland, 24–27 September 1997), edited by P. Blanchard and A. Jadczyk(Springer, Berlin), p. 105.

Kochen, S., 1985, in Symposium on the Foundations of Mod-ern Physics: 50 Years of the Einstein-Podolsky-Rosen Ex-periment (Joensuu, Finland, 1985), edited by P. Lahti andP. Mittelstaedt (World Scientific, Singapore), pp. 151–169.

Kubler, O., and H. D. Zeh, 1973, Ann. Phys. (N.Y.) 76, 405.Landau, L. D., 1927, Z. Phys. 45, 430.Landsman, N. P., 1995, Stud. Hist. Philos. Mod. Phys. 26,

45.Lewis, P., 1997, Br. J. Philos. Sci. 48, 313.Lockwood, M., 1996, Br. J. Philos. Sci. 47, 159.Malament, D. B., 1996, in Perspectives on Quantum Reality,

edited by R. Clifton (Kluwer, Boston), 1st edition, pp. 1–10.

Mermin, N. D., 1998a, Pramana 51, 549.Mermin, N. D., 1998b, Am. J. Phys. 66, 753.Milburn, G. J., 1991, Phys. Rev. A 44, 5401.Mohrhoff, U., 2004, Int. J. Quantum Inf. 2, 221.von Neumann, J., 1932, Mathematische Grundlagen der

Quantenmechanik (Springer, Berlin).Ollivier, H., D. Poulin, and W. H. Zurek, 2003, eprint quant-

ph/0307229.Omnes, R., 1988a, J. Stat. Phys. 53, 893.Omnes, R., 1988b, J. Stat. Phys. 53, 933.Omnes, R., 1988c, J. Stat. Phys. 53, 957.Omnes, R., 1990, Ann. Phys. (N.Y.) 201, 354.Omnes, R., 1992, Rev. Mod. Phys. 64, 339.Omnes, R., 1994, The Interpretation of Quantum Mechanics

(Princeton University, Princeton).Omnes, R., 2002, Phys. Rev. A 65, 052119.Paz, J. P., and W. H. Zurek, 1993, Phys. Rev. D 48, 2728.

Paz, J. P., and W. H. Zurek, 1999, Phys. Rev. Lett. 82, 5181.Pearle, P., 1976, Phys. Rev. D 13, 857.Pearle, P., 1982, Found. Phys. 12, 249.Pearle, P., 1984, Phys. Rev. D 29, 235.Pearle, P., 1986, Phys. Rev. D 33, 2240.Pearle, P., 1989, Phys. Rev. A 39, 2277.Pearle, P. M., 1979, Int. J. Theor. Phys. 48, 489.Pearle, P. M., 1999, in Open Systems and Measurement in

Relativistic Quantum Theory, Proceedings of the WorkshopHeld at the Instituto Italiano per gli Studi Filosofici, editedby H.-P. Breuer and F. Petruccione (Springer, Berlin), pp.31–65, eprint quant-ph/9901077.

Percival, I., 1995, Proc. R. Soc. London, Ser. A 451, 503.Percival, I., 1998, Quantum State Diffusion (Cambridge Uni-

versity, Cambridge, England).Pessoa, O., 1998, Synthese 113, 323.Rae, A. I. M., 1990, J. Phys. A 23, L57.Rovelli, C., 1996, Int. J. Theor. Phys. 35, 1637.Sanz, A. S., and F. Borondo, 2003, eprint quant-ph/0310096.Saunders, S., 1995, Synthese 102, 235.Saunders, S., 1997, The Monist 80, 44.Saunders, S., 1998, Synthese 114, 373.Saunders, S., 2002, eprint quant-ph/0211138.Schlosshauer, M., and A. Fine, 2005, Found. Phys. 35, 197.Schrodinger, E., 1926, Naturwissenschaften 14, 664.Squires, E. J., 1990, Phys. Lett. A 145, 67.Squires, E. J., 1991, Phys. Lett. A 158, 431.Stapp, H. P., 1993, Mind, Matter, and Quantum Mechanics

(Springer, New York), 1st edition.Stapp, H. P., 2002, Can. J. Phys. 80, 1043.Stein, H., 1984, Nous 18, 635.Tegmark, M., 1993, Found. Phys. Lett. 6, 571.Tegmark, M., 1998, Fortschr. Phys. 46, 855.Tegmark, M., 2000, Phys. Rev. E 61, 4194.Twamley, J., 1993a, eprint gr-qc/9303022.Twamley, J., 1993b, Phys. Rev. D 48, 5730.Unruh, W. G., and W. H. Zurek, 1989, Phys. Rev. D 40,

1071.Vaidman, L., 1998, Int. Stud. Philos. Sci. 12, 245.Vermaas, P. E., and D. Dieks, 1995, Found. Phys. 25, 145.van der Wal, C. H., A. C. J. ter Haar, F. K. Wilhelm, R. N.

Schouten, C. J. P. M. Harmans, T. P. Orlando, S. Lloyd,and J. E. Mooij, 2000, Science 290, 773.

Wallace, D., 2002, Stud. Hist. Philos. Mod. Phys. 33, 637.Wallace, D., 2003a, Stud. Hist. Philos. Mod. Phys. 34, 87.Wallace, D., 2003b, Stud. Hist. Philos. Mod. Phys. 34, 415.Wick, G. C., A. S. Wightman, and E. P. Wigner, 1952, Phys.

Rev. 88, 101.Wick, G. C., A. S. Wightman, and E. P. Wigner, 1970, Phys.

Rev. D 1, 3267.Wightman, A. S., 1995, Nuovo Cimento B 110, 751.Wigner, E. P., 1952, Z. Phys. 133, 101.Wigner, E. P., 1963, Am. J. Phys. 31, 6.Zeh, H. D., 1970, Found. Phys. 1, 69.Zeh, H. D., 1973, Found. Phys. 3, 109.Zeh, H. D., 1993, Phys. Lett. A 172, 189.Zeh, H. D., 1995, eprint quant-ph/9506020.Zeh, H. D., 1997, in New Developments on Fundamental Prob-

lems in Quantum Physics (Oviedo II), edited by M. Ferreroand A. van der Merwe (Kluwer Academic, Dordrecht), pp.441–452, eprint quant-ph/9610014.

Zeh, H. D., 1999, Found. Phys. Lett. 12, 197.Zeh, H. D., 2000, in Decoherence: Theoretical, Experimen-

tal, and Conceptual Problems, edited by P. Blanchard,

Page 41: arXiv:quant-ph/0312059v4 28 Jun 2005 · arXiv:quant-ph/0312059v4 28 Jun 2005 ... Bacciagaluppi (2003a) in the Stanford Encyclopedia of ... several popular interpretations of quantum

41

D. Giulini, E. Joos, C. Kiefer, and I. Stamatescu (Springer,Berlin), Lecture Notes in Physics No. 538, pp. 19–42, eprintquant-ph/9905004.

Zeh, H. D., 2001, The Physical Basis of the Direction of Time(Springer, Berlin), 4th edition.

Zeh, H. D., 2003, Phys. Lett. A 309, 329.Zurek, W. H., 1981, Phys. Rev. D 24, 1516.Zurek, W. H., 1982, Phys. Rev. D 26, 1862.Zurek, W. H., 1991, Phys. Today 44, 36, see also the updated

version available as eprint quant-ph/0306072.Zurek, W. H., 1993, Prog. Theor. Phys. 89, 281.Zurek, W. H., 1998, Philos. Trans. R. Soc. London, Ser. A

356, 1793.Zurek, W. H., 2000, Ann. Phys. (Leipzig) 9, 855.

Zurek, W. H., 2003a, private communication.Zurek, W. H., 2003b, Rev. Mod. Phys. 75, 715.Zurek, W. H., 2003c, Phys. Rev. Lett. 90, 120404.Zurek, W. H., 2004a, eprint quant-ph/0405161.Zurek, W. H., 2004b, in Science and Ultimate Reality, edited

by J. D. Barrow, P. C. W. Davies, and C. H. Harper (Cam-bridge University, Cambridge, England), pp. 121–137.

Zurek, W. H., F. M. Cucchietti, and J. P. Paz, 2003, eprintquant-ph/0312207.

Zurek, W. H., S. Habib, and J. P. Paz, 1993, Phys. Rev. Lett.70, 1187.

Zurek, W. H., and J. P. Paz, 1994, Phys. Rev. Lett. 72, 2508.