[PDF] Revisiting the Einstein-Bohr Dialogue



Previous PDF Next PDF







Revisiting the Einstein-Bohr Dialogue

Revisiting the Einstein-Bohr Dialogue Don Howard Einstein and Bohr – No names loom larger in the history of twentieth-century physics, and rightly so, Albert Einstein and Niels Bohr being the figures most prominently associated with the relativity and quantum revolutions 1 Their names dominate, likewise, the history of philosophical



The Bohr-Einstein dialogue : a rhetorical and genre analysis

present study examines the final phase of the Bohr-Einstein dialogue, in which Niels Bohr answers a challenge co-authored by Albert Einstein with a text that is, at once, generic enough to pass as a legitimate reply to Einstein and filled with unconventional generic forms borrowed from a neighbouring sphere of scientific activity



Chapter 4: Dialogue - Original Thinking

between Einstein and Bohr was a dialogue in which each listened deeply to what the other was saying for the purpose of understanding, not to persuade the other of the correctness of a particular belief If either Bohr or Einstein had been able to suspend their judgments and listen



Dokument1 - univieacat

The dialogue between Einstein and Bohr reached its culmination in 1935 with the publication of the famous Einstein-PodoIsky-Rosen paper, where the question whether quantum mechanics provides a complete description of physical reality was raised in a most succinct way 5 The novel states were called by Schrödinger



The Experiments of Einstein and Rupp on the Particle vs

“Dear Einstein”, Niels Bohr’s letter of 27 April 1927 begins, “[b]efore his holiday trip to the Bavarian mountains, Heisenberg asked me to send you a copy of the proofs that he was expecting, which he hoped might interest you ” Bohr sent Einstein the Albert proofs of Heisenberg’s article on the uncertainty relations



Dokument1 - univieacat

Ever since the famous Bohr-Einstein dialogue I it has been known that it is not pos- Sible in an interference experiment to have a maximum visibility interference pattern and path information at the same time This feature of quantum mechanics, neces- sary for its consistency, has been elevated by Feynman2 to a principle: whenever it is



How ideas became knowledge: The light-quantum hypothesis 1905

Isis, 58 (1967), 37–55 Martin J Klein, “The first phase of the Bohr-Einstein dialogue,” HSPS, 2 (1970), 1–39 Christa Jungnickel and Russell McCormmach, Intellectual mastery of nature: Theoretical physics from Ohm to Einstein, vol 2, The now mighty theoretical physics 1870–1925 (Chicago, 1986) 4



PATH INFORMATION IN QUANTUM INTERFEROMETRY

Ever since the famous Bohr-Einstein dialogue, it has been known that it is not pos­ sible in an interference experiment to have a maximum visibility interference pattern and path information at the same time This feature of quantum mechanics, neces­ sary for its consistency, has been f:'evated by Feynman2 to a principle: whenever it is

[PDF] intrication quantique communication

[PDF] bell inequality

[PDF] dieu ne joue pas aux dés einstein

[PDF] intrication quantique expérience

[PDF] décohérence quantique pdf

[PDF] paradoxe epr

[PDF] intrication quantique

[PDF] alain aspect prix nobel

[PDF] ens cachan

[PDF] non localité quantique

[PDF] bohm bijoux

[PDF] paradoxe epr pour les nuls

[PDF] mondialisation sur le plan politique

[PDF] la mondialisation politique pdf

[PDF] l'aspect politique de la mondialisation

R evisiting the Einstein-Bohr Dialogue Don H oward Einstein and Bohr - No names loom larger in the history of twentieth-century physics, andrig htly so, Albert Einstein and Niels Bohr being the figures most prominently associated with therela tivity and quantum revolutions. Their names dominate, likewise, the history of philosophical1re actions to the new physics of the twentieth century, Bohr for having identified complementaritya

s the chief novelty in the quantum description of nature, Einstein for having found vindication in2rela

tivity theory for either positivism or realism, depending upon whom one asks. Famous as is each3in his own domain, they

are famous also, together, for their decades-long disagreement over thefuture of fundamental physics, their respective embrace and rejection of quantum indeterminacybeing only the most widely-known point of contention.

A well-entrenched narrative tells the story of the Einstein-Bohr debate as one in whichEinstein's tries, from 1927 throug

h 1930, to prove the quantum theory incorrect via thoughtex

periments exhibiting in-principle violations of the Heisenberg indeterminacy principle, only tohave B

ohr find the flaw in each, after which Einstein shifts his direction of attack, faulting thequant um theory now not as incorrect, but incomplete. In 1935, the Einstein, Podolsky, and Rosen(

EPR) paper, "Can Quantum-Mechanical Description of Physical Reality Be Considered Complete?"(Einstein, Podolsky

, Rosen 1935) represents the high-water mark of this critique. It is met by Bohr'sdeep a

nd devastating reply (Bohr 1935), after which Bohr grows ever more in stature and influenceas the sag

e of Copenhagen, while Einstein slips into senility in Princeton, meddling, perhapscommendab ly, in the politics of the atomic bomb, but no longer capable of constructive contributionsto phy sics, itself.4 -2- That something is seriously wrong with this triumphalist narrative has been remarked uponby

various authors for more than twenty years. More than anyone else, however, it was Mara Beller,5in her

Quantum Dialogue (Beller 1999) who forced a reassessment. One need not agree with every6detail in Be ller's own account of the way in which the community around Bohr in Copenhagenachie ved consensus on questions of interpretation in order to appreciate the point that the writing orrew

riting of history is, itself, one of the many tools with which communities define themselves andconstruct c

onsensus. Reason enough always to be just a bit suspicious of any community's tellingi ts own story. Some aspects of Beller's new history I commend; others I do not. I commend Beller'sstressing , like James Cushing before her (Cushing 1994), the role of contingent social and historicalci rcumstance in the achievement of consensus. More than Beller, I find cogency in Bohr'sarg uments, but there can be no doubt that the victory of a "Copenhagen" point of view oninterpre tation is explained, in no small measure, by such factors as Bohr's control over financialresour

ces through his institute in Copenhagen, by his personal prestige, and even by a conversationalmanner

that some regarded as persistent and others as bullying. But Beller identifies Bohr as thechief enforcer of a Copenhagen orthodoxy, whereas I think that what later came to be regarded asCopenhag en orthodoxy owed more to Werner Heisenberg than to Bohr. On my reading, Bohr wasne

ver a positivist, did not endorse wave-packet collapse, and did not demand ideological conformityamong

his followers. It was not by compulsion, but by the example of his dogged pursuit of the deep7philosophical lessons of the quantum that B

ohr created and sustained not a unitary "Copenhageninterpre tation," but what Leon Rosenfeld called the "Copenhagen spirit" (Rosenfeld 1957) -3- That the standard, triumphalist Copenhagen narrative requires significant revision is,ho

wever, as noted, a point upon which I wholeheartedly agree with Beller. It needs revision if onlyto redre

ss the insult to Einstein. The present paper is a contribution to that revision. I argue that thestandard history

is wrong not only for the kinds of reasons cited by Beller, but also for the reason thatit m

ore or less completely misses the real point at issue between Einstein and Bohr. Simply put, bothBohr

and Einstein understood early and clearly that the chief novelty of the quantum theory was whatwe

, today, call "entanglement," the non-factorizability of the joint states of previously interactingqu

antum systems. Bohr embraced entanglement, seeing in it the roots of complementarity. Einsteinreje

cted entanglement as incompatible with the principle of the spatial separability of systems, aprinciple that he

thought not only a necessary feature of any field theory like general relativity butalso a ne cessary condition for the very intelligibility of science. Everything else is derivative,including B ohr's defense of complementarity and Einstein's charge of the incompleteness ofquantum mecha nics, a charge unsustainable without the assumption of what Einstein termed the"sepa ration principle." On this way of revising the history, Bohr still emerges with the better arguments. Here, too,B

eller and I disagree. But Einstein's legacy is rehabilitated, his dissent being seen for what it was:principled, we

ll-motivated, based upon deep physical insight, and informed by a sophisticatedph ilosophy of science. Why later apologists for Copenhagen orthodoxy preferred the senile Einsteinas B ohr's antagonist is hard to fathom, for Bohr's defense against Einstein's critique makes moresense a nd is more interesting when read as a reply to a good argument rather than a bad one. Puttinga

debate about entanglement at center stage has the additional salubrious effect of highlighting a stillm

ore general failure of the received history of quantum mechanics in the early twentieth century. It -4- was not just Einstein and Bohr who were arguing about entanglement. Everyone was. That the jointstates of pre viously interacting quantum systems do not factorize and that the quantum mechanicalstory about multi-particle systems differs, thus, in a fundamental way from the classical story werefac s "entanglement" in 1935, and long before quantum information theory madeentang

lement the hot, new topic in the foundations literature in the 1990s. Why the standard8histories of quantum mecha

nics have not emphasized this is as puzzling as their penchant for lettingcar icature take the place of a sympathetic and accurate account of Einstein's critique of the quantumt heory.Einstein, Bohr, and Entanglement before the

1927 Solvay Meeting

Standard histories of the Einstein-Bohr debate have it beginning in earnest during thepersona l encounter between Einstein and Bohr at the 1927 Solvay meeting. This was an importantc

onfrontation, but to understand what Einstein and Bohr were arguing about then and later, one mustunderstand f

irst their prior confrontations with the issue of entanglement. Einstein had been thinkingabout the proble

m for over twenty years, Bohr for nearly ten.

From the very beginning, that is, from his 1905 paper on the photon hypothesis, Einsteinunderstood that the e

merging quantum theory was likely to incorporate some compromise with thefundame ntally classical idea of the mutual independence of interacting systems. The reason was9sim ple and obvious. Recall that Einstein's route to the photon hypothesis was via a thermodynamicanalog y. Einstein showed that expression for the entropic behavior of radiation in the Wien (high-fr equency) regime exhibited the same functional form as the entropy of a classical, Boltzmann gas. -5-

That ana

logy sustained the inference that radiation in the Wien regime had a microstucture like thatof a Boltzmann gas, in the sense that it consisted of discrete, distinguishable, mutually independent,c

orpuscle-like carriers of the field energy that Einstein dubbed light quanta. In Einstein's own words:"Monochr

omatic radiation of low density (within the domain of validity of Wien's radiation formula)b ehaves from a thermodynamic point of view as if it consisted of mutually independent energyquanta of the magnitude Rßí /N " (Einstein 1905, 143). What did it mean to say that light quantawer

e "mutually independent"? It meant precisely that the Boltzmann formula applied or that entropyis an extensive pr

operty of the radiation gas, for the additivity of the entropy of a composite systemis equivalent to the fa

ctorizability of the joint probabilities for the differently spatially situatedcomponent subsy stems occupying given cells of phase space, this latter being the normal way ofe xpressing the probabilistic independence of two events. Again, in Einstein's own words:12 If we have two systems S and S that do not interact with each other, we can put11 If these two systems are viewed as a single system of entropy S and probability W, we have12 and12

W = W A W.

The last relation tells us that the states of the two systems are mutually independent events.(Einstein 1905, 140-141)

But this argument works only in the Wien limit. That such mutual independence is thussh

own only to hold in the Wien limit and that it is likely to fail outside of the Wien regime isobvious,

however much this obvious fact goes unremarked in much of the secondary literature on -6-

Einstein and t

he quantum. The simple fact is that, from the very beginning, Einstein understood thelimit ations of the argument and understood that, in general, the radiation field could not be modeledin s imple corpuscular terms.

If there be any doubt that Einstein understood that mutual independence would not obtainoutside of the Wien reg

ime, consider what he wrote to Hendrik Lorentz four years later: I must have expressed myself unclearly in regard to the light quanta. That is to say,I am not at all of the opinion that one should think of light as being composed of mutuallyindepende nt quanta localized in relatively small spaces. This would be the most convenientexplanation of the Wien end of the radiation formula. But already the division of a light raya t the surface of refractive media absolutely prohibits this view. A light ray divides, but al ight quantum indeed cannot divide without change of frequency. As I already said, in my opinion one should not think about constructing light out ofdiscr ete, mutually independent points. I imagine the situation somewhat as follows: . . . Iconce

ive of the light quantum as a point that is surrounded by a greatly extended vector field,that somehow diminishes with dist

ance. Whether or not when several light quanta are presentwith mutually overlapping fields one must imagine a simple superposition of the vectorfields, that I

cannot say. In any case, for the determination of events, one must have equationsof motion for the sing

ular points in addition to the differential equations for the vector field.(Einstein to L orentz, 23 May 1909, Einstein 1993, Doc. 163)I n other words, the idea of wave-particle duality is born out of Einstein's effort to exploit theme taphor of interference for describing the manner in which light quanta outside of the Wien regimedo not behave like mutually independent systems.

Over the next decade, many people in addition to Einstein struggled to understand what kindof statistics

was appropriate for quantum systems behaving in such a non-classical manner. Ane

specially illuminating moment in that history is recorded in a series of papers by Mieczys³awWolfke, who was Einstein's c

olleague in Zurich for a short time in 1913-1914. Wolfke was seekingto un derstand exactly what kind of independence among light quanta was assumed in Einstein's 1905 -7- arg

ument, and his remarks are interesting in part because they are based on direct communicationsfrom Einstein, himself. I

n the last of four papers, Wolfke wrote: In fact the Einsteinian light quanta behave like the individual, mutually independentmol ecules of a gas . . . . However, the spatial independence of the Einsteinian light quantaco mes out even more clearly from Einstein's argument itself. From the Wien radiationf ormula Einstein calculates the probability W that all n light quanta of the same frequency0 enclosed in a volume v find themselves at an arbitrary moment of time in the subvolume v0 of the volume v. The expression for this probability reads: 0

W = (v/v).n

0

This probability may be interpreted as the product of the individual probabilities v/vthat an individual one of the

light quanta under consideration lies in the subvolume v at ana rbitrary moment of time. From the fact that the total probability W is expressed as the0 product of the individual probabilities v/v, one recognizes that it is a matter of individualmu

tually independent events. Thus we see that, according to Einstein's view, the fact that alight quantum li

es in a specific subvolume is independent of the position of the other lightqu anta. (Wolfke 1914, 463-464)But unde rstanding the independence of photons in the Wien limit is not the same thing asunderstanding the manner in which that independence fails outside of the Wien regime. It tookanother

ten years for Einstein to solve the quantum statistics puzzle thanks to the fortuitousintervention of the I

ndian physicist, Satyendra Nath Bose (1924). How deeply the problem of quantum statistics weighed upon Einstein's soul before 1924 isclea r from a far-too-little-known argument adduced by Einstein in 1920 for the purpose ofopposing( !) a field ontology of trackable, mutually independent carriers of field energy. The settingis the nstein's visiting professorship in Leiden, home to Lorentz, whom Einstein held in high esteem.I n what is partly a gesture of respect for Lorentz, Einstein evinces surprising sympathy for theconce pt of the ether, arguing that, in many respects, the space-time of general relativity has taken -8- over the

role of the electromagnetic ether, at least inasmuch as it is the abode of field energy. Toquestions about the microstruc

ture of fields, Einstein gives an interesting answer: The special theory of relativity does not compel us to deny the aether. We mayassume the e

xistence of an ether; only we must give up ascribing a definite state of motionto it, i.e., we must by

abstraction take from it the last mechanical characteristic whichL orentz had still left it. . . . Think of waves on the surface of water. Here we can describe two entirely differentthings. Either we may follow how the undulatory surface forming the boundary betweenwater and air alters in the course of time; or else-with the help of small floats, forinstance-w e can follow how the position of the individual particles of water alters in thecourse of time. If the existence of such floats for tracking the motion of the particles of afluid were a fundamental impossibility in physics-if, in fact, nothing else whatever weredis cernible than the shape of the space occupied by the water as it varies in time, we shouldha ve no ground for the assumption that water consists of movable particles. But all the samewe could characterize it as a medium. We have something like this in the electromagnetic field. For we may picture the fieldto o urselves as consisting of lines of force. If we wish to interpret these lines of force toourselves a s something material in the ordinary sense, we are tempted to interpret thedy namic processes as motions of these lines of force, such that each individual line of forceis tracke d through the course of time. It is well known, however, that this way of regardingthe electromagnetic field leads to contradictions. Generalizing we must say this: There may be supposed to be extended physicalobjects to which the ide a of motion cannot be applied. They may not be thought of asc

onsisting of particles that allow themselves to be individually tracked through time. InMinkowski's idiom t

his is expressed as follows: Not every extended structure in thefour- dimensional world can be regarded as composed of world-threads. The special theoryo

f relativity forbids us to assume the ether to consist of particles that can be tracked throughtime, but the hy

pothesis of the ether in itself is not in conflict with the special theory ofrela

tivity. Only we must be on our guard against ascribing a state of motion to the ether.(Einstein 1920, 9-10)

Savor the irony

. Here is Einstein, the inventor of the photon hypothesis, arguing against modelingthe electromagnetic field as being composed of distinguishable and, thus, trackable field quanta, andthi

s because such a model is inherently non-relativistic. One easily imagines the relief Einstein feltwhen, pr

odded by Bose, he realized in 1924 that he could rescue a quantum field ontology bydeny ing the distinguishability of quanta. Surely we see here another important reason for Einstein's -9- pr

ompt and warm embrace of Bose, along with the fact that Bose's new statistics yielded an elegant,first-princ

iples derivation of the Planck formula for black body radiation.

Accepting Bose's new statistics, however, meant facing up to the fact that the failure of themutual inde

pendence of quanta outside of the Wien limit, suspected by Einstein since at least 1909,wa s going be a deep and pervasive feature of a still not yet established quantum mechanics. Einsteine

xplained this point in his second paper on the extension of bosonic statistics to material particles:

Bose's theory of radiation and my analogous theory of ideal gases have been reprovedby Mr. Ehrenfest and other colleagues because in these theories the quanta or molecules arenot treate d as structures statistically independent of one another, without this circumstancebe ing especially pointed out in our papers. This is entirely correct. If one treats the quantaas being statistically independent of one another in their localization, then one obtains theWien ra diation law; if one treats the gas molecules analogously, then one obtains theclassica l equation of state for ideal gases, even if one otherwise proceeds exactly as Bose andI have. . . . It is easy to see that, according to this way of calculating [Bose-Einsteinsta tistics], the distribution of molecules among the cells is not treated as a statisticallyindepend

ent one. This is connected with the fact that the cases that are here called"complexions" would not be reg

arded as cases of equal probability according to thehy pothesis of the independent distribution of the individual molecules among the cells.As signing different probability to these "complexions" would not then give the entropyc

orrectly in the case of an actual statistical independence of the molecules. Thus, the formula[for the entropy

] indirectly expresses a certain hypothesis about a mutual influence of themolecules-fo r the time being of a quite mysterious kind-which determines precisely theequal statistical proba bility of the cases here defined as "complexions." (Einstein 1925, 5-6)Ei In the Bose statistics employed by me, the quanta or molecules are not treated as beingindepende nt of one another. . . . A complexion is characterized through giving the numbero f molecules that are present in each individual cell. The number of the complexions sodefine d should determine the entropy. According to this procedure, the molecules do nota

ppear as being localized independently of one another, but rather they have a preference tosit togethe

r with another molecule in the same cell. One can easily picture this in the case ofsma ll numbers. [In particular] 2 quanta, 2 cells: -10- Bose -statisticsindependent molecules1s t cell2nd cell1st cell2nd cell1st case !!-1s t caseI II -2n d caseI

II2nd case

3r d caseII I3r d case-!!

4th case-

I II According to Bose the molecules stack together relatively more often than according to thehy bruary 1925, EA 22-002)

xt several months through the lens of his ongoing correspondence with Einstein, it is hard to avoidthe conc

apable of expressing precisely the curious failure of the mutual independence of quantummecha

nical systems revealed in the new quantum statistics, just as Einstein, in 1909, deployed themetaphor

of wave interference to model the failure of mutual independence outside of the Wienreg ime. The key move of writing the n-particle wave function in 3n-dimensional configuration space,ra

ther than 3-space, is necessary because only thus does one have access to a state space rich enoughto c

er's regret, was the loss of visualizability, but that was precisely the price required byentang lement, as both Bohr and Einstein would quickly realize. -11- On Einstein's side there followed two years of intense theoretical and experimental workprobing

the subtleties of the new quantum mechanics of multi-particle systems. With advice fromEinstein, the Be

rlin experimentalist and master of the technology of coincidence counting, WaltherBothe , pursued a kind of proto-Bell experimental program, investigating a variety of differentc

orrelation phenomena (see, for example, Bothe 1926). On the theoretical front, Einstein's effortsculminated in May

of 1927-a few months before the first of the famous clashes with Bohr at theSolvay mechanics. Ironically, Einstein's model failed precisely because it included the verye

ntanglement Einstein was hoping to avoid. Though it reached the proof stage, Einstein's lecture tothe B

erlin Academy was never published because Einstein could not find a way around theentang lement. He explained in a "Note Added in Proof": I have found that the schema does not satisfy a general requirement that must beimp osed on a general law of motion for systems. Consider, in particular, a system Ó that consists of two energetically independent12 subsystems, Ó and Ó; this means that the potential energy as well as the kinetic energy is1 additively composed of two parts, the first of which contains quantities referring only to Ó,2 the second quantities referring only to Ó. It is then well known that 12 11

22where Ø depends only on the coordinates of Ó, Ø only on the coordinates of Ó. In this casewe

must demand that the motions of the composite system be combinations of possiblemot ions of the subsystems. The indicated scheme [Einstein's hidden variables model] does not satisfy this1 requirement. In particular, let ì be an index belonging to a coordinate of Ó, í an index2 ìíbelonging to a coordinate of Ó. Then Ø does not vanish. (Einstein 1927a) Here, then, we have Einstein on the eve of the 1927 Solvay meeting. For twenty-two yearshe had known that the full story of the quantum would involve some fundamental compromise withthe classical notion of the mutual independence of interacting systems. In 1924, Bose showed him -12- that this failure of mutual indepe ndence was not incidental but an essential feature of the quantumrea

nical formalism that relates the new quantum statistics to the symmetry properties of the two-particle

wave function. In the spring of 1927, Einstein's attempt at a hidden variables model of wavemechaquotesdbs_dbs23.pdfusesText_29