You are on page 1of 10

Quantum Darwinism

Wojciech Hubert Zurek


Theory Division, MS B213, LANL Los Alamos, NM, 87545, U.S.A.

Quantum Darwinism describes the proliferation, in the environment, of multiple records of selected
states of a quantum system. It explains how the fragility of a state of a single quantum system can
lead to the classical robustness of states of their correlated multitude; shows how effective ‘wave-
packet collapse’ arises as a result of proliferation throughout the environment of imprints of the
states of quantum system; and provides a framework for the derivation of Born’s rule, which relates
probability of detecting states to their amplitude. Taken together, these three advances mark
considerable progress towards settling the quantum measurement problem.

The quantum principle of superposition implies that I. DECOHERENCE AND EINSELECTION


arXiv:0903.5082v1 [quant-ph] 29 Mar 2009

any combination of quantum states is also a legal state.


This seems to be in conflict with everyday reality: States Decoherence turns one of the two problems we noted
we encounter are localized. Classical objects can be ei- above – fragility of quantum states – into a solution of the
ther here or there, but never both here and there. Yet, the other. Environment-induced decoherence recognizes that
principle of superposition says that localization should be if a measurement can put a state at risk and re-prepare
a rare exception and not a rule for quantum systems. it, so can accidental information transfers that happen
Fragility of states is the second problem with quantum- whenever a system interacts with its environment.
classical correspondence: Upon measurement, a general Decoherence is by now well understood [3, 4, 5]:
preexisting quantum state is erased – it “collapses” into Fragility of states makes quantum systems very difficult
an eigenstate of the measured observable. How is it then to isolate. Transfer of information (which has no effect on
possible that objects we deal with can be safely observed, classical states) has dramatic consequences in the quan-
even though their basic building blocks are quantum? tum realm. So, while fundamental problems of classical
physics were always solved in isolation (it sufficed to pre-
To bypass these obstacles Bohr [1] followed Alexander vent energy loss) this is not so in quantum physics (leaks
the Great’s example: Rather than try disentangling the of information are much harder to plug).
Gordian Knot at the beginning of his conquest, he cut When a quantum system gives up information, its own
it. The cut separates the quantum from the classical. state becomes consistent with the information that was
Bohr’s Universe consists of two realms, each governed by disseminated. “Collapse” in measurements is an extreme
its own laws. Fragile superpositions were banished from example, but any interaction that leads to a correlation
the classical realm deemed more fundamental and indis- can contribute to such re-preparation: Interactions that
pensable to interpret or even practice quantum theory. depend on a certain observable correlate it with the en-
Thus, instead of trying to understand Universe (includ- vironment, so its eigenstates are singled out, and phase
ing “the classical”) in quantum terms one “quantized” relations between such pointer states are lost [6].
this and that, always starting from the classical base. Negative selection due to decoherence is the essence of
This was a brilliant tactical move: Physicists could environment-induced superselection, or einselection [7]:
conquer the quantum realm without getting distracted by Under scrutiny of the environment, only pointer states
interpretational worries. In those days only gedankenex- remain unchanged. Other states decohere into mixtures
periments like the famous Schrödinger cat [2] were truly of stable pointer states that can persist, and, in this sense,
disturbing: Real experiments dealt with electrons, pho- exist: They are einselected.
tons, atoms, or other microscopic systems. Bohr’s rule of These ideas can be made precise. The basic tool is the
thumb – that the macroscopic is classical – was enough. reduced density matrix ρS . It represents the state of the
Moreover, many (including Einstein) believed that quan- system that obtains from the composite state ΨSE of S
tum physics is just a step on a way to a deeper theory and E by tracing out the environment E:
that will solve or bypass interpretational conundrums.
ρS = T rE |ΨSE ihΨSE | . (1)
That did not happen. Instead, old gedankenexperi-
ments were carried out. They confirmed validity of quan- Evolution of ρS reveals preferred states: It is most pre-
tum laws on scales that have, of recent, begun to infringe dictable when the system starts in a pointer state. To
on “the macroscopic”. Quantum theory is here to stay. quantify this one can use (von Neumann) entropy, HS =
It is also increasingly clear that its weirdest predictions H(ρS ) = −T rρS lg ρS , as a function of time. Pointer
– superpositions and entanglement – are experimental states result in smallest entropy increase. By contrast,
facts, in principle relevant also for macroscopic objects. their superpositions produce entropy rapidly, at decoher-
Therefore, questions about the origin of “the classical”, ence rates, especially when S is macroscopic.
with its restriction to localized states that are robust, un- When pure states of the system are sorted by pre-
perturbed by measurements, can no longer be dismissed. dictability, according to entropy of the evolved ρS ,
2

pointer states are at the top. This criterion – the pre-


dictability sieve [4, 8, 9] – yields a short list of candidates
for effectively classical states: A cat can persist in one
of the two obvious stable states, but their superposition
would deteriorate into a mixture of |deadi and |alivei
when initiated in a way envisaged by Schrödinger [2].
The special role of position is traced to the nature of
the SE interactions: They tend to depend on distance.            
   

'  (  )  *
   

Hence, information about position is most readily passed


                    

%     &  #   "         

!
 "  # $ %     &  #   "

+ ,
         &  #   " 

on to the environment. This is why localized states sur-


vive while nonlocal superpositions decay into their mix-
tures. For example, in a weakly damped harmonic os-
cillator the minimum uncertainty wavepackets – familiar
coherent states, best quantum approximation of classical
points in phase space – are einselected [9, 10, 11].

II. ENVIRONMENT AS A WITNESS

Monitoring by the environment means that informa- - 

+ ,
     &  #   "  . /  0  1 2    

tion about S is deposited in E. What role does it play,     3  4 5 #   "   6 .   . 0 6 6 .  

and what is its fate? Decoherence theory ignores it. En-


  . / 7 8 9 0  1 : 7      ;  / 1 .     <

vironment is “traced out”. Information it contains is FIG. 1: Quantum Darwinism and the structure of the envi-
treated as inaccessible and irrelevant: E is a “rug to sweep ronment. Decoherence theory distinguishes between a system
under” the data that might endanger classicality. (S) and its environment (E) as in (a), but makes no further
Quantum Darwinism recognizes that “tracing out” is recognition of the structure of E; it could as well be mono-
not what we do: Observers eavesdrop on the environ- lithic. In Quantum Darwinism the focus is on redundancy.
ment. Vast majority of our data comes from fragments We recognize the subdivision of E into subsystems, as in (b).
of E. Environment is a witness to the state of the system. The only requirement for a subsystem is that it should be
For example, this very moment you intercept a fraction individually accessible to measurements; observables of dif-
of the photon environment emitted by a screen or scat- ferent subsystems commute. To obtain information about S
from E one can then measure fragments F of the environ-
tered by a page. We never access all of E. Tiny fractions
ment – non-overlapping collections of subsystems of E, (c).
suffice to reveal the state of various “systems of interest”. ically, there are many copies of the information about S in E
This insight captures the essence of Quantum Darwin- – “progeny” of the “fittest observable” that survived monitor-
ism: Only states that produce multiple informational off- ing by E proliferates throughout E. This proliferation of the
spring – multiple imprints on the environment – can be multiple informational offspring defines Quantum Darwinism.
found out from small fragments of E. The origin of the The environment becomes a witness with redundant copies of
emergent classicality is then not just survival of the fittest information about the preferred observable. This leads to the
states (the idea already captured by einselection), but objective existence of pointer states: Many can find out the
their ability to “procreate”, to deposit multiple records state of the system independently, without prior information,
– copies of themselves – throughout E. and they can do it indirectly, without perturbing S.
Proliferation of records allows information about S to
be extracted from many fragments of E (in the example
of the system was the basic tool of decoherence. To study
above, photon E). Thus, E acquires redundant records of
Quantum Darwinism we focus on correlations between
S. Now, many observers can find out the state of S in-
fragments of the environment and the system. The rele-
dependently, and without perturbing it. This is how pre-
vant reduced density matrix ρSF is given by:
ferred states of S become objective. Objective existence
– hallmark of classicality – emerges from the quantum ρSF = T rE/F |ΨSE ihΨSE | . (2)
substrate as a consequence of redundancy.
Decoherence theory was focused on the system. Its aim Above, trace is over “E less F”, or E/F – all of E except
was to determine what states survive information leaks for the fragment F. How much F knows about S can be
to E. Now we ask: What information about the system quantified using mutual information:
can be found out from fragments of E? This change of
focus calls for a more realistic model of the environment I(S : F) = HS + HF − HS,F , (3)
(Fig. 1): Instead of a monolithic E we recognize that envi-
ronments consist of subsystems that comprise fragments defined as the difference between entropies of two sys-
independently accessible to observers. tems (here S and F) treated separately and jointly. For
The reduced density matrix ρS representing the state example, the mutual information between an original and
3

and indirectly – without perturbing S.


Rapid rise and gradual leveling of I(S : Ff ), Fig. 2,
implies redundancy. The information in Ff allows one
to determine the state of S as it reaches redundancy
plateau. Observables of different F’s commute – such
measurements are independent. Yet, underlying corre-
lations mean that their outcomes imply the same state
of the system, as if S were classical: The redundancy
plateau is a classical plateau. Its level HS is the classical
information accessible from a small fraction of E.
Redundancy allows for objective existence of the state
of S: It can be found out indirectly, so there is no danger
of perturbing S with a measurement. Error correction al-
FIG. 2: Information about S stored in E and its redundancy. lowed by redundancy is also important: Fragility of quan-
Mutual information is monotonic in f . When global state of tum states means that copies in F’s are damaged by mea-
SE is pure, I(S : Ff ) in a typical fraction f of the environ- surements (we destroy photons!), and may be measured
ment is antisymmetric around f = 0.5 [13]. For pure states in a “wrong” basis. One cannot access records in E with-
picked out at random from the combined Hilbert space HSE , out endangering their existence. But with many (Rδ )
there is little mutual information between S and a typical F copies, state of S can be found out by ∼ Rδ observers
smaller than half of E. However, once a threshold f = 12 is who can get their information independently, and with-
attained, nearly all information is in principle at hand. Thus, out prior knowledge about S. Consensus between copies
such random states (green line) exhibit no redundancy. By suggests objective existence of the state of S.
contrast, states of SE created by decoherence (where the en-
The mutual information I(S : Ff ) computed in mod-
vironment monitors preferred observable of S) contain almost
all (all but δ) of the information about S in small fractions els of decoherence exhibits behavior illustrated by the red
fδ of E. The corresponding I(S : Ff ) (red line) quickly rises plot of Fig. 2. In the family of models representing spin
to HS (entropy of S due to decoherence), which is all of the S surrounded by environments of many spins [12, 13, 14]
information about S available from either E or S. (More, up the same number of spins suffices to reach the plateau:
to 2HS , can be obtained only through global measurements Adding more spins to E only extends length of the plateau
on S and nearly all E). HS is therefore the classically acces- measured in “absolute units” – in the number of the en-
sible information. As (1 − δ)HS of information is contained vironment spins. In this model (that can be viewed as
in fδ = 1/Rδ of E, there are Rδ such fragments in E: Rδ a simplified model of a photon environment) redundancy
is the redundancy of the information about S. Large redun- is then proportional to the number of the environment
dancy implies objectivity: The state of the system can be
subsystems that interact with the system of interest.
found out indirectly and independently by many observers,
who will agree about their conclusions. Thus, Quantum Dar- Quantum Brownian motion – harmonic oscillator sur-
winism accounts for the emergence of objective existence. rounded by many environmental oscillators – is the other
well known model of decoherence. It is exactly solvable,
and the case of an underdamped oscillator yields sur-
a perfect copy (of, say, a book) is equal to the entropy of prisingly simple results [15, 16]: (i) Mutual information
f
the original, as either contains the same text. So, every is approximately given by I(S : F) ≈ HS + 21 ln (1−f ),
bit of information in the first copy reveals a bit of infor- and; (ii) Redundancy for an initially squeezed state of S
mation in the original. However, having extra copies does reaches Rδ ≈ s2δ , where s, the squeeze factor, quantifies
not increase the information about the original. Yet, it delocalization of the state. Similar equation should hold
determines how many can independently access this in- for more general “Schrödinger cat” states, with s quan-
formation. The number of copies defines redundancy. tifying the separation of the two localized alternatives.
Similar ideas apply to the quantum case. Initially, ev- These results confirm intuitions that originally moti-
ery bit of information gained from a fraction f  1 of vated Quantum Darwinism [4, 17]: Monitoring of the
E that was pure before it monitored (and decohered) the system by the environment can deposit multiple records
system is a bit about S. The red plot in Fig. 2 starts with of preferred states of S in E. States of SE that arise from
this steep “bit for bit” slope, but moderates as I(S : Ff ) decoherence are special [13, 14], as I(S : Ff ) for a typ-
approaches redundancy plateau at HS , where additional ical pure state selected with Haar measure in the whole
bits only confirm what is already known. Hilbert space of SE (green plot in Fig. 2) shows. In
Redundancy is the number of independent fragments such random states small fragments reveal almost noth-
of the environment that supply almost all classical infor- ing about the rest of the state. Only when half of E is
mation about S, i.e., (1 − δ)HS . In other words; found out the whole state is suddenly revealed.
Rδ = 1/fδ . (4) States that arise from decoherence are then far from
random. Roughly speaking, they have a branch structure.
Rδ is the number of times one can acquire (1 − δ) of the This is why the rest of such a branch including the state
information about S independently (from distinct F’s) of the system – the “bud” from which this branch has
4

originated – can be deduced from its fragment. We shall not interact with each other. This is why light deliv-
see how such branches grow in the next section. ers most of our information. Moreover, photons emitted
Plots of I(S : Ff ) for pure SE are antisymmetric by the usual sources (e.g., sun) are far from equilibrium
around the point {HS , f = 21 } for typical fragments of with our surroundings. Thus, even when decoherence is
E [13]. Thus, rapid rise for small f must be matched at dominated by other environments (e.g., air) photons are
the other end, for f ∼ 1. This is a signature of entan- much better in passing on information they acquire while
glement that allows state to be known “as the whole”, “monitoring the system of interest”: Air molecules scat-
while states of subsystems are unknown. The joint state ter from one another, so that whatever record they may
of SE is then pure, so that HS,F =E = 0, and I(S : Ff ) have gathered becomes effectively undecipherable.
must rise to HS + HE = 2HS when f approaches 1. Stability of the level of the redundancy plateau at HS ,
This is a very quantum aspect of information. In clas- even for mixed E’s, is a compelling reason to think of it as
sical physics knowing a composite object implies knowing “classical”. The question we shall now address concerns
each of its subsystems. This is not so in quantum physics, the nature of that information – what does the environ-
where composite states are given by tensor (rather than ment know about the system, and why?
Cartesian) products of their constituents. Thus, one can
know perfectly quantum state of the whole, but know
nothing about states of parts. We shall see in Section IV III. FROM COPYING TO QUANTUM JUMPS
how this feature can be used to derive Born’s rule [18]
that relates probabilities with wavefunctions. Quantum Darwinism leads to appearance, in the en-
To reveal this latent quantumness one would have to vironment, of multiple copies of the state of the system.
measure the right global observable on all of SE. For However, the no-cloning theorem [20, 21] prohibits copy-
example, when mutual information, Eq. (3), is defined ing of unknown quantum states. If cloning is outlawed,
using Shannon entropy with probabilities corresponding how can redundancy seen in Fig. 2 be possible?
to optimal observables in S and in E, the resulting Shan- Quick answer is that cloning refers to (unknown) quan-
non I(S : Ff ) graph for small f would look very similar tum states. So, copying of observables evades the theo-
to Fig. 2. However, using Shannon entropy involves lo- rem. Nevertheless, the tension between the prohibition
cal probabilities (precluding global observables), so such on cloning and the need for copying is revealing: It leads
Shannon I(S : Ff ) never exceeds HS , antisymmetry is to breaking of unitary symmetry implied by the super-
lost, and the plateau continues until the end at f ∼ 1. position principle, accounts for quantum jumps, and sug-
Effective unattainability of the f ∼ 1 part of the plot gests origin of the “wavepacket collapse”, setting stage for
also shows why decoherence is so hard to undo: Correla- the study of quantum origins of probability in Section IV.
tions that reveal coherence can be usually detected only Quantum physics is based on several “textbook” pos-
by such global measurements of whole SE. We intercept tulates [22]. The first two; (i) States are represented by
small fractions of E, and never have the luxury of perfect vectors in Hilbert space, and; (ii) Evolutions are unitary –
global measurements needed to undo decoherence. Yet, give complete account of mathematics of quantum theory,
because of redundancy, we get ∼ HS information with but make no connection with physics. For that one needs
“sloppy” measurements of f  1. to relate calculations made possible by the superposition
Quantum Darwinism does not require pure E. Mixed principle of (i) and unitarity of (ii) to experiments.
environment is a noisy communication channel: Its initial Postulate (iii) Immediate repetition of a measurement
entropy of h per bit can still increase after interaction yields the same outcome starts this task. This is the only
with S, reflecting mutual information buildup. However, uncontroversial measurement postulate (even if it is diffi-
now a bit gained from E yields only 1−h of a bit about S. cult to approximate in the laboratory): Such repeatability
So, a completely mixed E (h = 1) is useless (even though or predictability is behind the very idea of “a state”.
it can still induce decoherence!). For a partly mixed E In contrast to (i)-(iii), collapse postulate (iv) Outcomes
mutual information will increase more slowly, pure case correspond to eigenstates of the measured observable, and
“bit per bit” rate tempered to ∼ 1 − h. Yet, it can still only one of them is detected in any given run of the ex-
climb the same redundancy plateau at HS [19]. periment, is inconsistent with (i) and (ii). Conflict arises
These conclusions apply when E is initially mixed, but for two reasons: Restriction to a preferred set of outcome
are also relevant when this channel is noisy for other rea- states seems at odds with with the egalitarian principle
sons (e.g., imperfect measurements). In all such cases one of superposition, embodied in (i). This restriction pre-
can still reach the same redundancy plateau, although vents one from finding out unknown quantum states, so
now a proportionally larger fragment of the environment it is responsible for their fragility. And a single outcome
is needed to get the same information about S. per run is at odds with unitarity (and, hence, linearity)
Suitability of the environment as a channel depends of quantum dynamics that preserves superpositions.
on whether it provides a direct and easy access to the The last axiom; (v) Probability of an outcome is given
records of the system. This depends on the structure by the square of the associated amplitude, pk = |ψk |2 ,
and evolution of E. Photons are ideal in this respect: is known as Born’s rule [18]. It completes the relation
They interact with various systems, but, in effect, do between mathematics of (i) and (ii) and the experiments.
5

a) b) c)
1.0 50 1.0
0.8 40 0.8

R0.1 (σ)

I(σ : e)
IˆN (σ)
0.6 30 0.6
0.4 20 0.4
µ = 0.23
0.2 10 0.2
0 0
0 0 0
0
π/4
µ π/4 π/4 40 50
π/8 µ π/4 π/4 µ 20 30
π/2 0 a π/8 π/2 0 10 m
π/2 0 a

FIG. 3: Quantum Darwinism in a simple model of decoherence [12]. The spin- 12 S interacts with N = 50 spin- 12 subsystems of E
Ek
with an Ising Hamiltonian HSE = N S √1 (|0i+|1i)⊗|0iE1 ⊗. . .⊗|0iEN . Couplings gk are
P
k=1 gk σz ⊗σy . The initial state of S⊗E is 2
distributed randomly in the interval (0,1]. All the plotted quantities are a function of the observable σ(µ) = cos(µ)σz +sin(µ)σx ,
where µ is the angle between its eigenstates and the pointer states of S – eigenstates of σzS . a) Information acquired by the
optimal measurement on the whole environment, IˆN (σ), as a function of the inferred observable σ(µ) and the average interaction
action hgk ti = a. A lot of information is accessible in the whole E about any observable σ(µ) except when a is so small that
there was no decoherence. b) Redundancy of the information about S as a function of the inferred observable σ(µ) and the
average action hgk ti = a. Rδ=0.1 (σ) counts the number of times 90% of the total information can be “read off” independently
by measuring distinct fragments of E. It is sharply peaked around the pointer observable: Redundancy is a very selective
criterion – the number of copies of relevant information is high only for the observables σ(µ) inside the theoretical bound (see
Ref.[12]) indicated by the dashed line. c) Information about σ(µ) extracted by local random measurements on m environmental
subsystems. Because of redundancy, pointer states – and only pointer states – can be found out through this far-from-optimal
strategy. Information about any other observable σ(µ) is restricted to what can be inferred from the pointer observable [12].

Bohr bypassed conflict of (i) and (ii) with (iv) by insist- demand”: As in cloning, one asks for “two (or more) of
ing that apparatus is classical, so unitarity and the prin- the same”. Its conflict with linearity of quantum the-
ciple of superposition need not apply to measurements. ory can be resolved only by restricting states that can
But this is an excuse, not an explanation. We are dealing be copied. Such pointer states then act as “buds” of
with a quantum environment, and redundancy of previ- branches that grow by reproducing, in E, multiple copies
ous section strengthened motivation for postulate (iii) – of the original in S. Interaction Hamiltonians do not per-
repeatability. Let us see where this demand takes us in turb observables that commute with them. So, buds of
a purely quantum setting of postulates (i), (ii), and (iii). branches coincide with the einselected pointer states.
Suppose there are states of S (say, |ui and |vi) that Evidence of such symmetry breaking is seen in Fig.
produce an imprint in a subsystem of E (which plays a 3. Mutual information and redundancy shown there are
role of an apparatus), but remain unperturbed (so they obtained using Eq. (3), but with Shannon (rather than
can produce more imprints). This repeatability implies: von Neumann) entropies of specific observables of S and
|ui|e0 i ⇒ |ui|eu i, |vi|e0 i ⇒ |vi|ev i in obvious notation. F, i.e., using probabilities of their eigenstates. While von
In a unitary process scalar product is preserved. Thus; Neumann-based I(S : Ff ) and Rδ characterized total
hu|vi = hu|viheu |ev i , (5) information, Shannon-based counterparts are well suited
to enquire: What observable is this information about?
where we have set he0 |e0 i = 1. This simple equation It turns out that the environment as a whole “knows”
can be satisfied only when; (a) either heu |ev i = 1 (which many observables of S, as is seen in Fig. 3a. By contrast,
means that copying was completely unsuccessful), or; (b) in Fig. 3b symmetry breaking is evident: The ridge of
hu|vi = 0, i.e., they are orthogonal. In that case heu |ev i redundancy appears abruptly only when test observable
is arbitrary – perfect record heu |ev i = 0 is also possible. σ(µ) and the preferred pointer observable σz (that re-
It follows that multiple (perfect or imperfect) copies mains unperturbed by the environment) nearly coincide.
of |ui and |vi can be imprinted in disjoint F’s. As a
consequence of unitarity, only sets of orthogonal states Why are pointer states favored? Commonsense says
(that define Hermitean observables [22]) can be so copied, that, to be reproduced, state must survive copying. This
explaining selection of a set of outcomes – terminal points leads to a theorem [12, 24] that only pointer states can be
of quantum jumps [23]. Before, they had to be postulated discovered from fractions of E. Other observables (such
by the first part of axiom (iv). We emphasize that this as σ(µ) in Fig. 3) can be deduced only to the extent they
result relies on just two values of the scalar product – 0 are correlated with the pointer observable. So, fragments
and 1 – and, thus, does not appeal to Born’s rule. of the environment offer a very narrow, projective point
This breaking of unitary symmetry (choice of preferred of view. Redundant imprinting of some observables hap-
states in an egalitarian Hilbert space) is induced by re- pens at the expense of their complements.
peatability of the information transfer. It is a “nonlinear Structure of branching state betrays its origin and fore-
6
Pn
shadows “collapse”. Starting from |ψS i = k ψk |sk i, Selection of the set of outcomes by the proliferation of
information essential for Quantum Darwinism parallels
n n
X (1) (N )
X Bohr’s insistence [1] that a “classical apparatus” should
|ΨSE i = ψk |sk i|ek i . . . |ek i = ψk |sk i|εk i (6) determine the outcomes. However, it follows from purely
k k
quantum Eq. (5), and is caused by a unitary evolution
branches grow to include N subsystems of E. Branch responsible for the information transfer. Nevertheless, as
(j) (j) classical apparatus would, preferred pointer states desig-
fragments can be nearly orthogonal; ΠJj=1 hek |ek0 i '
nate possible future outcomes, precluding measurements
δkk0 for large enough J. This means that a pointer state
of complementary observables or determining preexist-
|sk i of S can be determined (along with the rest of the
ing state of the system. Thus, information acquisition –
branch) from a sufficiently long fragment (which may still
a copying process – results in preferred states.
be short compared to the length of the branch, J  N ).
In the huge Hilbert space HSE branching state is a Consensus between records deposited in fragments of
very atypical minimally entangled superposition of only E looks like “collapse”. In this sense we have accounted
n product “branches” labelled by the pointer states of for postulate (iv) using only very quantum postulates (i)-
the system. This is tiny compared to the dimension of (iii). In particular, in deriving and analyzing Eq. (5) we
HSE that exceeds n by a factor exponential in N . This have not employed Born’s rule, axiom (v). We shall be
is why the two plots in Fig. 2 are so different: Branch- therefore able to use our results as a starting point for
ing state is, to a good approximation, a multi-system such a derivation in the next section.
Schmidt decomposition, with long branch fragments con- There was nothing nonunitary above – unitarity was
stituting “systems”. In a Schmidt decomposition, states the crux of our argument, and orthogonality of branch
of partners are in one-to-one correspondence. Thus, in seeds our main result. Relative states of Everett [26, 27,
Eq. (6), |sk i implies |εk i (and, vice versa), and measur- 28] come to mind. One could speculate about reality of
ing a branch fragment F can reveal the whole branch. branches with other outcomes. We abstain from this –
Initial part of I(S : Ff ), Fig. 2, represent buildup of our discussion is interpretation-free, and this is a virtue.
this correlation: When f = 0, observer is ignorant of Indeed, “reality” or “existence” of universal state vector
what branch he will find out, but the structure of the seems problematic. Quantum states acquire objective
correlations within |ΨSE i leaves no doubt of what these existence when reproduced in many copies. Individual
branches are. Using Born’s rule one could assign to them states – one might say with Bohr – are mostly informa-
probabilities pk = |ψk |2 and the corresponding entropy tion, too fragile for objective existence. And there is only
HS . Next section shows how one can deduce these prob- one copy of the Universe. Treating its state as if it really
abilities without axiom (v) – how symmetries of entan- existed [26, 27, 28] seems unwarranted and “classical”.
glement imply Born’s rule.
When observer measures enough of E, he finds out
the branch (and what the state of S is). Additional IV. PROBABILITIES FROM ENTANGLEMENT
data are redundant. They only confirm what is already
known. Probabilities associated with |ΨSE i are replaced Observer prepared S in a state |ψS i, but wants to mea-
with certainty of a branch. This transition from uncer- sure observable with eigenstates {|sk i}. This will lead to
tainty (initial presence of many branches – potential for entangled |ΨSE i with branch structure, Eq. (6). Pointer
multiple outcomes) to certainty (once a sufficiently long states {|sk i} define the outcomes, but, as yet, observer
branch fragment becomes known) accounts for percep- has not measured E, and does not know the result. Given
tion of “collapse”. The initial, steeply rising, part of |ΨSE i, what is the probability of, say, |s17 i?
I(S : Ff ) “resolves” it: Collapse is brief compared to To derive it we cannot use reduced density matrices,
the ensuing period of certainty about the outcome, as Eqs. (1,2). Tracing out is averaging [25, 29, 30] – it relies
fδ  1, but, nevertheless, not instantaneous. on pk = |ψk |2 , Born’s rule we want to derive. We have
Assumptions that lead from copying to preferred states imposed that ban while deriving and analyzing Eq. (5),
can be relaxed. Thus, E need not be initially pure [23]. but relaxed it to plot Fig. 3. Now we reimpose it again.
Moreover, it suffices that the records (e.g., in the appara- So, Born’s rule and standard tools of decoherence are
tus A) are “repeatably accessible”. Transfer of responsi- off limits – using them courts circularity. Our derivation
bility for repeatability from a quantum S to a (still quan- will rest instead on certainty and symmetry, cornerstones
tum) A allows one to model non-orthogonal measurement that mark two extremal cases of probability.
outcomes (POVM’s): A entangles with the system, and The case of certainty was just settled without Born’s
then acts as ancilla. Its orthogonal pointer states |Ak i rule using Eq. (5). When one re-measures an observable,
P
correlate with non-orthogonal |ςk i of S, k ψ̃k |ςk i|Ak i. the same outcome will be seen again. Thus, when {|sk i}
Interaction of A with the environment results in multiple includes |ψS i (e.g., |ψS i = |s17 i), newly added copies
copies of |Ak i. The usual projective measurement imple- just extend the branch already correlated with observer’s
mentation of POVM’s (see e.g. [25]) is now straightfor- state, and the outcome is certain; p17 = 1. Certainty of
ward. Branches are labelled by |Ak i. Indeed, we usually correlations between partners in Schmidt decomposition,
experience “quantum jumps” via an apparatus pointer. Eq. (6) is another important example.
7

a)
~

b)
=
| >| >+| >| >
S E S E

c) | >| >+| >| >


S E S E
| >| >+| >| >
S E S E

| >| >+| >| > = | >| >+| >| >


S E S E S E S E

FIG. 4: Probabilities and symmetry: (a) Laplace used subjective ignorance to define probability. Player who does not know face
values of the cards, but knows that one of them is a spade will infer probability p♠ = 12 for the top card. (b) The real physical
state of the system is however altered by the swap, illustrating subjective nature of Laplace’s approach, and demonstrating its
unsuitability for physics. (c) Perfectly known entangled states have objective symmetries that allow one to rigorously deduce
probabilities. When two systems are maximally entangled as above, probabilities of Schmidt partners are equal, p♥ = p♦ , and
p♠ = p♣ . After a swap uS = |♠ih♥| + |♥ih♠| in S, the resulting state |♠i|♦i + |♥i|♣i must have p0♠ = p♦ , and p0♥ = p♣ . (We
‘primed’ probabilities in S, as it was acted upon by a swap, so they might have changed.) A counterswap uE = |♦ih♣| + |♣ih♦|
in E restores the original entangled state, proving that p0♥ = p♥ and p0♠ = p♠ , after all (as counterswap uE leaves S untouched).
This sequence of equalities implies p♠ = p♦ = p♥ , so that p♠ = p♥ = 12 , as probabilities in S must add up to 1.

Certainty seems trivial but is important. Confirmation Figure 4 illustrates how this classical intuition yields –
that a state “is what it is” – postulate (iii) – is a part of far more convincingly — quantum probabilities.
standard quantum lore [22]. We re-affirmed it, but with Symmetry is probed by invariance. Transformations
a key insight: Redundancy allows observers to discover that respect it take system between states that exhibit
(and not just confirm) that S is in a certain pointer state. no measurable differences. For example, change of phase
We now turn to the opposite case of complete inde- Pnthe coefficients in the Schmidt decomposition |ΨSE i =
in
terminacy. Its connection with symmetry was noted by k ψk |sk i|εk i cannot influence the state of S: It is in-
Laplace. He wrote: “The theory of chance consists in re- duced by uS = eiφk |sk ihsk |, local unitary on S, that can
ducing all the events ... to a certain number of cases that be “undone” by uE = e−iφk |εk ihεk | on E, or;
are equally possible... The ratio of this number to that of
all the cases possible is the measure of probability” [31]. uS ⊗ 1E |ΨSE i = |ΦSE i; 1S ⊗ uE |ΦSE i = |ΨSE i (7)
8

So, phases of ψk cannot matter for a local state or influ- of S. However, this is done by a unitary “countertrans-
ence probabilities in S. This symmetry, Eq. (7), is the formation” acting solely on E. Hence, by fact (1), state
entanglement-assisted invariance or envariance [32, 33]. of S must have been unaffected by uS in the first place.
Such loss of phase significance for S entangled with E So, by fact (2), phases of ψk cannot change outcomes of
implies decoherence [33]. We arrived at its essence using any measurement on S. Equiprobability follows.
envariance, without reduced density matrices, trace, etc. One can now derive Born’s rule, pk = |ψk |2 , with
We now use phase envariance to show that equal ab- straightforward algebra from the above two simple cases
solute values of the coefficients ψk imply equal prob- of complete certainty (pk = 1) and equiprobability (pk =
1
abilities. For equal |ψk | any orthogonal basis of S n ): The general case can be always reduced to the case
is “Schmidt” (i.e., has an orthogonal partner in E). case of equal coefficients by “finegraining” (see Box).
Thus, |ϕ̄SE i =
|0iS |0iE +|1iS |1iE
√ =
|+iS |+iE +|−iS |−iE
√ , The origin of probability is a fascinating problem that
2 2 is older than quantum measurement problem, and is for-
|0i±|1i
where |±i = √
2
. Sign change induced by eiπ |−ih−| gotten primarily because it is so old. We have seen how
|+iS |+iE −|−iS |−iE quantum physics sheds a new, very fundamental, light
acting on S produces |η̄SE i = √
2
=
|1iS |0iE +|0iS |1iE on probability. We cannot do justice to the history of

2
In other words, one can swap |0iS with
. this subject here, but Ref. [34] provides a basic overview
|1iS by rotating phase in a |±i basis by π. Yet, we just and exhaustive set of references. In particular, envariant
saw that phases of Schmidt coefficients do not matter for derivation is very different from the classic proof of Glea-
the state of S, so probabilities of 0 and 1 in S must have son [35] in that it sheds light on the physical significance
remained the same. Moreover, probabilities of paired up of the resulting measure. Moreover, it does not assume
Schmidt states are equal, so that pS (0) = pE (0) in |ϕ̄SE i probabilities are additive (except to posit that probabil-
and pS (1) = pE (0) in |η̄SE i. Hence, pS (0) = pS (1) = 21 , ity of an event and its complement are certain, i.e., to
where we assumed that probabilities add up to 1. establish normalization; see Box and Ref. [33, 38]). By-
In contrast to Laplace’s subjective “ignorance-based” passing additivity of probabilities is essential when deal-
approach, we obtained objective probabilities for a com- ing with a theory with another principle of additivity
pletely known entangled state. Phase envariance implied – the quantum superposition principle – which trumps
equiprobability in S. To paraphrase Beatles, “All you additivity of probabilities or at least classical intuitiions
need is phase...”. We rotated phases of the coefficients to about it (e.g., in the double-slit experiment). Discus-
induce a swap in a complementary basis. Another proof sion of the implications of envariance has already started,
(that implements swap more directly) is given in Fig. 4. with [36, 37], and [5] providing insightful commentary.
This equiprobability case is the difficult part of the
BOX
proof. Instead of subjectivity (that undermined appli-
We show here how “finegraining” reduces the case of
cability of Laplace’s approach to physics) we relied on
arbitrary ψk to equiprobability. To illustrate general
objective symmetries of entangled quantum states. This
strategy consider state in a 2D Hilbert space HS of S
was made possible by the nature of quantum states of
spanned by orthonormal {|0i, |2i}q
and (at least) 3D HE :
composite systems. Classically, pure states have struc- q
2 1
ture of a Cartesian product – knowing the whole implies |ψSE i ∝ 3 |0iS |+iE + 3 |2iS |2iE .
knowledge of each subsystem. In quantum theory they The state |+iE = |0iE +|1iE
exists in (at least 2D) sub-

are tensor products – one can know state of the whole, 2
and thus know nothing about parts, as envariance shows. space of E orthogonal to |2iE , i.e., h0|1i = h0|2i = h1|2i =
h+|2i = 0. We know we can ignore phases.
This was the basis of our proof of equiprobability. We
assumed unitarity. Moreover, we assumed; (1) When a To reduce |ψSE i to equal coefficients case we “extend
system is not acted upon by a unitary transformation, its it” to a state |Ψ̄SEC i by letting E act on an ancilla C.
state remains unaffected. This state is a property of (S is not acted upon, so, by fact (1), probabilities for S
S alone, so; (2) Predictions regarding measurement out- cannot change.) This can be done by a generalization of
comes on S (including their probabilities) can be inferred controlled-not acting between E (control) and C (target),
from the state of S. Last not least; (3) When S is entan- so that (in obvious notation) |ki|00 i ⇒ |ki|k 0 i, leading to
gled with other systems (e.g., the environment) the state √ √ 0 0
of S alone is determined by the state of the whole SE. 2|0i|+i|00 i + |2i|2i|00 i ⇒ 2|0i |0i|0 i+|1i|1

2
i
+ |2i|2i|20 i.
These “facts of life” are accepted properties of systems
and states, but given the fundamental nature of our dis- Above, and from now on we skip subscripts: The state of
cussion it seems a good idea to make them explicit [33]. S will be listed first, and
√ the state of C will be primed.
For instance, to establish independence from phases of The cancellation of 2 yields an equal coefficient state:
the coefficients ψk we noted that the state of S is un-
affected by the unitaries uS diagonal in Schmidt basis |Ψ̄SCE i ∝ |0, 00 i|0i + |0, 10 i|1i + |2, 20 i|2i .
acting on S (like changes of Schmidt coefficient phases)
that would normally affect isolated S: The global state We have combined S and C in a single ket and (below)
ΨSE is restored by uE . Thus, by fact (3), so is local state we shall swap states of SC as if it was a single system.
9

Clearly, this is a Schmidt decomposition of (SC)E. “single idea” category. Several ideas, applied in the right
Three orthonormal product states have coefficients with order, led to advances described here. Logically, we may
the same absolute value. Therefore, they can be en- well have started with the derivation of Eq. (5) and the
variantly swapped. Thus, the probabilities of states analysis of quantum jumps. Their randomness leads to
|0i|00 i, |0i|10 i, and |2i|20 i are all equal. By normalization probabilities. And symmetries of entangled states (that
they are 13 . So, probability of detecting state |2i of S is arise in decoherence and Quantum Darwinism) allow one
1 to derive Born’s rule. As we have seen, phase envariance
3 . Moreover, |0i and |2i are the only two outcome states
for S. It follows that probability of |0i must be 23 ; is (nearly) “all you need”. With probabilities at hand
p0 = 23 ; p2 = 31 . one has then every right to use reduced density matrices
This is Born’s rule. We have just seen why the amplitudes to analyze Quantum Darwinism and decoherence.
in the initial |ψSE i “get squared” to yield probabilities. Our presentation was “historical”. We started with de-
Note that we have avoided assuming additivity of prob- coherence, and used it to introduce Quantum Darwinism.
abilities: p0 = 23 not because it is a sum of two fine- Analysis of copying essential to information flows in both
grained alternatives for SE, each with probability of 13 , of these phenomena led to quantum jumps. This in turn
but rather because there are only two (mutually exclu- motivated entangelment-based derivation of Born’s rule.
sive and exhaustive) alternatives for S; |0i and |2i, and Quantum Darwinism – upgrade of E to a communication
p2 = 13 . Therefore, by normalization, p0 = 1 − 13 . Prob- channel from a mundane role it played in decoherence –
abilities of Schmidt states can be added because of the tied together all of the other developments. This order
loss of phase coherence that follows directly from phase had the advantage of making motivations clear, but it is
envariance established earlier (see also Ref. [32, 33]). different from more logical presentation where postulates
Extension of this proof to the case where proba- (i)-(iii) are the starting point (strategy followed in [38]).
bilities are commensurate is conceptually straightfor- The collection of ideas discussed here allows one to un-
ward but notationally cumbersome. The case of non- derstand how “the classical” emerges from the quantum
commensurate probabilities is settled with an appeal to substrate staring from more basic assumptions than de-
continuity. Frequency of the outcomes can be also de- coherence. We have bypassed a related question of why is
duced, allowing one to establish connection with the fa- our Universe quantum to the core. The nature of quan-
miliar relative frequency approach to probabilities [32, tum state vectors is a part of this larger mystery. Our
33, 38], but in a quantum setting probability arises as a focus was not on what quantum states are, but on what
consequence of symmetries of a single entangled state. they do. Our results encourage a view one might describe
We end by noting that the finegraining discussed above (with apologies to Bohr) as “complementary”. Thus, |ψi
does not need to be carried out experimentally each time is in part information (as, indeed, Bohr thought), but
probabilities are discussed: Rather, it is a way to de- also the obvious quantum object to explain “existence”.
duce a measure that is consistent with the geometry of We have seen how Quantum Darwinism accounts for the
the Hilbert spaces using entanglement as a tool. Still, transition from quantum fragility (of information) to the
given fundamental implications of envariance experimen- effectively classical robustness. One can think of this
tal tests would be most useful. transition as “It from bit” of John Wheeler [39].
In the end one might ask: “How Darwinian is Quan-
tum Darwinism?”. Clearly, there is survival of the fittest,
V. DISCUSSION and fitness is defined as in natural selection – through
the ability to procreate. The no-cloning theorem implies
We derived the two controversial quantum postulates competition for resources – space in E – so that only
from the first three. We have thus seen how classical do- pointer states can multiply (at the expense of their com-
main of the Universe arises from the superposition princi- plementary competition). There is also another aspect
ple (postulate (i)) and unitarity (postulate (ii)) as well as of this competition: Huge memory available in the Uni-
rudimentary assumptions about information flows (pos- verse as a whole is nevertheless limited. So the question
tulate (iii)), and a few basic facts about states of com- arises: What systems get to be “of interest”, and imprint
posite quantum systems (including their tensor nature, their state on their obliging environments, and what are
often cited as additional “axiom (0)”). the environments? Moreover, as the Universe has a finite
The essence of the measurement problem – accounting memory, old events will be eventually “overwritten” by
for axioms (iv) and (v) – has been largely settled. It is of new ones, so that some of the past will gradually cease
course likely one may be able to clarify assumptions and to be reflected in the present record. And if there is no
simplify proofs. Much work remains to be done on Quan- record of an event, has it really happened? These ques-
tum Darwinism and envariance. Nevertheless, nature of tions seem far more interesting than deciding closeness
the quantum-classical correspondence has been clarified. of the analogy with natural selection [40]. They suggest
Physicists take it for granted that even hard problems one more question: Is Quantum Darwinism (a process of
are solved by a single good idea. Therefore, when a single multiplication of information about certain favored states
idea does not do the whole job, often our first instinct is to that seems to be a “fact of quantum life”) in some way
dismiss it. Measurement problem does not fall into this behind the familiar natural selection? I cannot answer
10

this question, but neither can I resist raising it.

[1] Bohr, N. The quantum Postulate and the recent devel- Oxford, 1958).
opment of atomic theory Nature 121, 580-590 (1928). [23] Zurek, W. H., Quantum origin of quantum jumps: Break-
[2] Schrödinger, E. Die gegenwärtige Situation in der ing of unitary symmetry induced by information transfer
Quantenmechanik. Naturwissenschaften 807-812; 823- and the transition from quantum to classical. Phys. Rev.
828; 844-849 (1935). A 76, 052110 (2007).
[3] Joos, E., Zeh, H. D., Kiefer, C., Giulini, D., Kupsch, [24] Ollivier, H., Poulin, D., and Zurek, W. H., Environment
J., and Stamatescu, I.-O., Decoherence and the Appear- as a Witness: Selective Proliferation of Information and
ancs of a Classical World in Quantum Theory, (Springer, Emergence of Objectivity in a Quantum Universe Phys.
Berlin, 2003). Rev. A72, 423113 (2005).
[4] Zurek, W. H. Decoherence, einselection, and the quan- [25] Nielsen, M. A., and I. L. Chuang, Quantum Computation
tum origins of the classical Rev. Mod. Phys. 75, 715-775 and Quantum Information, (Cambridge University Press,
(2003). 2000).
[5] Schlosshauer, M. Decoherence and the Quantum - to - [26] Everett III, H., Relative state formulation of quantum
Classical Transition (Springer, Berlin, 2007). theory. Rev. Mod. Phys. 29, 454-462 (1957).
[6] Zurek, W. H. Pointer basis of a quantum apparatus: Into [27] Everett III, H., 1957b, Ph. D. Dissertation, Princeton
what mixture does the wavepacket collapse? Phys. Rev. University.
D24, 1516-1525 (1981). [28] DeWitt, B. S., and Graham, N., eds., The Many - Worlds
[7] Zurek, W. H. Environment-induced superselection rules. Interpretation of Quantum Mechanics (Princeton Univer-
Phys. Rev. D26, 1862-1880 (1982). sity Press, Princeton, 1973).
[8] Paz, J.-P., and Zurek, W. H., Environment-induced deco- [29] Landau. L., Das Dämpfungsproblem in der Wellen-
herence and the transition from quantum to classical. pp. mechanik. Zeits. Phys. 45, 430-441 (1927).
533-614 in Coherent Atomic Matter Waves, Les Houches [30] von Neumann, J. 1932, Mathematical Foundations of
Lectures, R. Kaiser, C. Westbrook, and F. David, eds. Quantum Theory, translated from German original by R.
(Springer, Berlin, 2001). T. Beyer (Princeton University Press, Princeton, 1955).
[9] Zurek, W. H., Habib, S., and Paz, J.-P., Coherent states [31] Laplace, P. S,. 1820, A Philosophical Essay on Probabil-
via decoherence Phys. Rev. Lett. 70, 1187-1190 (1993). ities, English translation of the French original by F. W.
[10] Tegmark, M., and Shapiro, H. S., Decoherence produces Truscott and F. L. Emory (Dover, New York, 1951).
coherent states: An explicit proof for harmonic chains. [32] Zurek, W. H., Environment-assisted invariance, causal-
Phys. Rev. E50, 2538-2547 (1994). ity, and probabilities in quantum physics. Phys. Rev.
[11] Gallis, M. R., The emergence of classicality via decoher- Lett. 90, 120404 (2003).
ence described by Lindblad operators. Phys. Rev. A53, [33] Zurek, W. H., Probabilities from entanglement, Born’s
655 (1996). rule from envariance. Phys. Rev. A71, 052105 (2005).
[12] Ollivier, H., Poulin, D, and Zurek, W. H., Objective [34] Auletta, G., Foundations and Interpretation of Quantum
properties from subjective quantum states: Environment Theory (World Scientific, Singapore, 2000).
as a witness. Phys. Rev. Lett. 93, 220401 (2004). [35] Gleason, A. M., Measures on closed subspaces of Hilbert
[13] Blume-Kohout, R., and Zurek, W. H., A simple example space, J. Math. Mech. 6, 855-893 (1957).
of “Quantum Darwinism”: Redundant information stor- [36] Schlosshauer, M, and Fine, A., On Zurek’s derivation of
age in many-spin environments Found. Phys. 35, 1857 the Born rule. Found. Phys. 35(2), 197-213 (2005)
(2005). [37] Barnum, H., No-signalling-based version of Zurek’s
[14] Blume-Kohout, R., and Zurek, W. H., Quantum Darwin- derivation of quantum probabilities: A note on
ism: Entanglement, branches, and the emergent classi- “Environment-assisted invariance, entanglement,
cality of redundantly stored quantum information. Phys. and probabilities in quantum physics”, arXiv:quant-
Rev. A73, 062310 (2006). ph/0312150 (2003).
[15] Blume-Kohout, R., and Zurek, W. H., Quantum Darwin- [38] Zurek, W. H., Relative States and the Environment: Ein-
ism in quantum Brownian motion. Phys. Rev. Lett., 101, selection, Envariance, Quantum Darwinism, and the Ex-
240405 (2008). istential Interpretation, arXiv:0707.2832 (2007).
[16] J. P. Paz and A. Roncaglia, in preparation. [39] Wheeler, J. A., It from Bit. p. 3 in Complexity, Entropy,
[17] Zurek, W. H., Einselection and decoherence from an in- and the Physics of Information, Zurek, W. H., ed. (Ad-
formation theory perspective. Ann. Physik (Leipzig), 9, dison Wesley, Redwood City, 1990).
822 (2000). [40] Darwin, C., The Origin of the Species. (1859).
[18] Born, M., Zur Quantenmechanik der Stossvorgänge Acknowledgments: I am grateful to Robin Blume-
Zeits. Phys. 37, 863-867 (1926). Kohout, Fernando Cucchietti, Juan Pablo Paz, David
[19] M. Zwolak, H. T. Quan, and W. H. Zurek, in preparation. Poulin, Hai-Tao Quan, Michael Zwolak for stimulating
[20] Wootters, W. K., and Zurek, W. H., A single quantum discussions. This research was supported by an LDRD
cannot be cloned. Nature 299, 802-803 (1982).
grant at Los Alamos and, in part, by FQXi.
[21] Dieks, D., Communication by EPR devices. Phys. Lett.
92A, 271 (1982).
[22] Dirac, P. A. M., Quantum Mechanics (Clarendon Press,

You might also like