You are on page 1of 12

1

Tomasz Bigaj

Making trouble for Lewis in the quantum world

Abstract

Lewis’s account of counterfactuals has been accused of failing in indeterministic worlds. If this accusation is

justified, it seems natural that his analysis should be expected to fail in the quantum world too. In this paper an

analysis of a common quantum-mechanical experimental situation (known as the singlet-spin state) is presented

which makes trouble for Lewis not due to indeterminism, but due to its apparent non-locality. It is shown that

according to Lewis’s criteria of similarity, worlds that include non-local interactions have to be judged more

distant from the actual world than worlds with small miracles, which leads to the incorrect valuations of

common-sense counterfactuals. Several possibilities of rectifying Lewis’s original proposal will be analyzed,

including one that eliminates possible worlds with miracles (law-breaking events).

David Lewis’s analysis of counterfactuals, today commonly accepted as the textbook one,

consists of two components. Its first component contains the formal semantics for the

counterfactual conditional, based on the structure of the set of possible worlds together with

the binary relation of relative similarity. The central point of this formal semantics is Lewis’s

famous truth condition for counterfactuals, which prescribes that sentence A  B (“If it

were A, than it would be B”) is (non-vacuously) true at a given world w0 (called “the actual

world”) iff A and B are jointly true at some possible world w, and there is no possible A-and-

not-B-world which is more or equally similar to w0 than w (Lewis 1973). The second

component of Lewis’s conception is an informal one; it gives a set of qualitative criteria of

relative similarity between possible worlds, which enable the user to apply Lewis’s formal

semantics to real-life counterfactual statements. While there is little disagreement regarding

the claim that the formal analysis of the counterfactual correctly captures the semantic

intricacies of contrary-to-fact conditionals, the second, informal part of Lewis’s approach


2

gives rise to an ongoing controversy. For instance, it is claimed that his multi-tiered set of

criteria of similarity fails to produce correct valuations for common counterfactuals in the case

of indeterminism (Percival 1999).

In what follows, I will argue that Lewis’s account of similarity is prone to yet another

objection. I will consider a simple example of a counterfactual statement describing a

common experimental situation in quantum physics, which nevertheless cannot be correctly

valuated in the Lewis approach. The reason for this failure is not so much quantum

indeterminism, but another non-classical feature of the quantum world, namely its non-

locality.1 Next, several possibilities of rectifying Lewis’s original proposal will be analyzed.

1. Lewis proposes to take into account two factors when comparing possible worlds with

respect to their relative similarity to the actual world. These factors are: differences in

particular facts and differences in laws. The key point of his conception is that these two

factors are not meant to be weighed directly against each other; rather they create a hierarchy

of comparisons in which a particular level is taken into account only if the preceding one does

not yield a definite answer to the query. Lewis presents his ranking of respects of similarity as

follows:

(1) It is of the first importance to avoid big, widespread, diverse violations of law.

(2) It is of second importance to maximize the spatio-temporal region throughout

which perfect match of particular fact prevails.

(3) It is of the third importance to avoid even small, localized, simple violations of

law.

(4) It is of little or no importance to secure approximate similarity of particular fact.

(Lewis 1986, pp. 47-48).

1
An extensive counterfactual analysis of various versions of quantum non-locality can be found in (Bigaj 2006,
especially chapter 6).
3

The vagueness of the terms used in the above formulations, such as “big”, “small”, etc., does

not bother Lewis, for he claims that our counterfactual judgments are indeed vague, and any

formalization thereof should retain this vagueness. Nevertheless, he maintains that under the

standard resolution of vagueness his proposal gives expected valuations of many common-

sense counterfactuals, including the famous example of president Nixon pressing the button of

the doomsday machine, and therefore triggering a nuclear holocaust (Fine 1975).

Consider, however, the following example involving two spin-½ particles (e.g.

electrons), prepared in the singlet spin state, in which the total spin equals 0. Because of the

principle of the conservation of angular momentum, the total spin of these electrons cannot

change, as long as the particles are isolated from external influences. Hence, if the actually

performed measurement of the x spin component of the left-hand side particle L yields value

xL = +1, the value of the same spin component of the right-hand side particle R is bound to be

xR = 1. This correlation is independent of the relative location of the particles, so we can

assume that both measurements are space-like separated.

Under the assumption that the actually obtained outcomes are xL = +1 and xR = 1,

let us now analyze the following counterfactual conditional

(C) xL = 1  xR = +1.

which can be presented in words as “If the result of the measurement of spin in direction x for

the left-hand side particle were –1, the outcome of the spin measurement in the same direction

for the right-hand side particle would be +1”. Intuitively, this counterfactual expresses our

belief that if the result of one of the measurements were different, the distant outcome would

have to change too in order to keep the total spin unchanged, and therefore (C) should be

rendered true. However, it can be argued that Lewis’s complex set of criteria (1)-(4) produces
4

the opposite answer. In order to see this, we have to consider two possible worlds w1 and w2

such that in both of them the outcome of the left-hand side measurement is xL = 1. Their

main difference is that in w1 the law of the conservation of angular momentum is upheld

exactly as in w0, and hence the outcome of the other measurement is xR = +1. However, in w2

a temporary suspension of the law occurs, allowing for the other outcome to remain

unchanged: xR = 1. Now the issue of what the logical value of the analyzed counterfactual is

boils down to the question which of the worlds w1 and w2 should be seen as closer to w0. If w1

is more similar to w0 than w2, then (C) comes out true as predicted. But if w2 is equally or

more similar to w0 than w1, the value of the counterfactual is “false”.

It is not difficult to see, however, that Lewis’s criteria seem to favor w2 over w1. First

of all, it may be argued that in none of these worlds is there a big, widespread violation of

laws, mentioned in criterion (1). The violation of law that is allowed to happen in w2 is quite

well located and limited, so it may arguably count as a “small miracle” in Lewis’s

terminology, and hence can be taken into account no sooner than in the third step. But before

that criterion can be applied there comes condition (2), which demands that the world in

which the region of perfect match in terms of particular facts is bigger should be counted as

more similar. But, as it is clearly depicted on Fig. 2 and 3, world w2 has a significantly smaller

region of divergence of particular facts in comparison with w1. World w1 differs from w0 in

causal consequences of the outcome of the left-hand side measurement, and the right-hand

side measurement alike. These causal consequences include electromagnetic radiation

spreading from the location of the experiments and carrying the information about the

outcomes, the records written on a computer’s disk, the experimenters’ memories, etc.2 In

world w2 those diverging causal consequences are constrained to the future light cone of the

2
It is irrelevant whether we agree that these differences will be slowly fading away with time, or that they may
accumulate over time to produce situations much different qualitatively from those in the actual world. After all,
according to Lewis an approximate match between a possible world and the actual one is the weakest criterion of
all four.
5

left-hand side measurement only, and hence the exact match in this part of FR which is disjoint

from FL (note that FR  FL is potentially infinite) is bought at the price of one small miracle.

And, according to Lewis’s conditions, this trade-off brings a net profit. Thus, w2 is more

similar to w0 than w1, and counterfactual (C) becomes false, in spite of our strong inclination

to the contrary.3

Fig. 1. The actual world w0

Fig.2. Possible “no-miracle” world w1

3
Notice that this troublesome situation would not take place if the right-hand side measurement occurred in the
absolute future of the left-hand side measurement. In such a case keeping the actual result of the xR-
measurement intact (at the expense of the small miracle) would not increase the area of perfect match between
possible world w2 and the actual world w0, as all the differences between w2 and w0 would be confined to the
future light cone of the xL-measurement anyhow. Consequently, w2 would be less similar to w0 in comparison
with w1 in virtue of criterion (3).
6

Fig. 3. Possible “miracle” world w2

In subsequent sections I will consider possible remedies to this trouble that we are

making for Lewis’s theory, laid out in an increasing order of plausibility. Accordingly, the last

solution presented is the one that I am ultimately inclined to accept.

2. Small or big miracle? The first conceivable solution turns on the apparent vagueness of the

distinction between big and small miracles. It may be argued that the violation of the law

required for the right-hand side outcome to remain unchanged, while the left-hand side

outcome was counterfactually modified, actually counts as a big miracle. If this strategy

succeeded, world w2 would become more distant from w0 than w1 thanks to criterion (1), and

the right valuation of (C) would be achieved. In order to make this solution at least prima

facie plausible we may point out that according to Lewis the number of violated laws has very

little to do with whether the violation is big or small (Lewis 1986, p. 55), so it is of no concern

here that the only law broken is the principle of the conservation of angular momentum.

However, not much more can be said in support of this scheme. Lewis explicitly says that the

extent of a law violation should be measured with the help of a number and spatiotemporal

size of unlawful events participating in this violation (pp. 55-56). So it is unlikely that we

could resolve the vagueness associated with this criterion so that the occurrence of two almost

point-like unlawful events (xL = 1 and xR = 1 )4 would count as a big miracle. The
4
Note that taken separately, none of these events is unlawful; only their simultaneous occurrence counts as a law
violation.
7

physical explanation of this apparent “containment” of the miracle occurring in w2 may be that

the lawful correlation between elements of the quantum entangled system is non-local and

discriminating (meaning, roughly, that only the other particle can “feel” the underlying non-

local influence from the first particle; see Maudlin 1994, p. 23), so in order to “cut off” the

connection between the left- and right-hand side particles we don’t have to adjust any other

phenomena taking place in the vicinity of both measurements at w1. This fact, as I believe,

undermines the feasibility of this solution to our problem.

3. The common cause. When a discrepancy occurs between formal and intuitive semantic

analyses of a given expression of natural language, it is always possible to point out that our

intuitions are not sacred, and that sometimes it is necessary to choose the formal but

unintuitive analysis (a typical example: material implication versus indicative conditional).

However, there have to be good reasons for doing so; otherwise this strategy would give an

easy victory to virtually any ill-conceived formalization of natural language. In our example

there should be some hidden factors which would account for the fact that although the

correlation between distant outcomes exists, a counterfactual altering of one of them does not

force the other to be changed. Those hidden factors are known jointly as a common cause of

the correlation.

The way the common cause works in cutting off the counterfactual dependence

between correlated events may be explained with a simple example. Let us imagine that two

billiard balls are arranged on a table in such a way that every shot that sends one of them into

a pocket has to push the other ball simultaneously into another pocket. If that is the case, then

we have a case of a perfect correlation: ball A lands in one pocket if and only if ball B lands in

the other pocket. And yet the appropriate counterfactual is false: if we imagined a situation in

which after the shot my hand prevented ball A from falling into its pocket, this by itself should
8

not affect the trajectory of ball B. Here we have a clear case of a non-causal correlation which

can be screened off once an appropriate common factor is taken into account (in this case the

cue ball hitting the first ball).

In the case of two entangled particles the falsity of (C) can be secured by postulating

that in the common past of both measurements an event E occurs which causally determines

both outcomes to be xL = +1 and xR = 1. Considering a counterfactual situation in which xL

= 1, we imagine that right before the left-hand side measurement a tiny miracle occurred,

flipping the spin of the particle. But this miracle could in no way affect the past event E, and

hence the right-hand side outcome is bound to remain xR = 1. The counterfactual

dependence between distant outcomes has been broken by the common cause.

Unfortunately, the main disadvantage of this solution is that it relies on the assumption

that such a common cause exists, which is highly improbable in the light of Bell’s no-go

results (Bell 1987, van Fraassen 1982). Moreover, even if this assumption were probable, it is

still no good that Lewis’s approach does not seem to work in a scenario in which there is no

common cause at all. General semantic analyses of this sort should not depend on

extralinguistic facts, such as whether the physical world is local or non-local, deterministic or

indeterministic, etc.

4. The asymmetry by fiat. Lewis himself considered an alternative to his preferred analysis

(1986, p. 39). Rather than applying the complex set of guidelines (1)-(4), we could do as

follows. When evaluating counterfactual (C), let us consider a possible world which is exactly

as w0 up to the moment directly preceding the measurement of xL, at which point a small

miracle occurs, permitting the outcome of the measurement to be xL = 1. The miracle is

assumed to be “minimal” required to do the job with no unnecessary divergences from the

actual world being permitted. Finally, after the measurement the world evolves according to
9

usual laws, with no miracles whatsoever. If the consequent of the analyzed counterfactual

turns out to be true in all such worlds, the entire counterfactual comes out true; if not—the

counterfactual is false.

This method would obviously yield the required value for (C). The alteration of the

outcome of the xR-measurement would come out as a result of keeping standard laws,

including the principle of the conservation of angular momentum. A similar strategy has been

already adopted almost unanimously by the authors working with quantum counterfactuals

(Redhead 1987, Clifton et al. 1990, Bedford & Stapp 1995, Finkelstein 1998, 1999). In order

to make it relativistically invariant, some of these authors interpret the phrase “up to the

moment preceding the measurement” as denoting the past light cone with its apex in close

proximity of the xL-measurement.

However, Lewis raised two powerful arguments against the above solution. Firstly, it

arbitrarily renders all (non-trivial) backtracking counterfactuals false.5 And although Lewis is

definitely not a fan of backtracking counterfactuals, he would like to leave open the possibility

that some of them may actually be true (1986, p. 40). After all, the fact that we don’t typically

encounter cases when the past depends counterfactually on the future is supposed to be a

contingent feature of our world; in an alternative world this might be different. Secondly, the

proposed analysis has a limited scope of applicability, as it is applicable only to

counterfactuals with antecedents referring to spatiotemporally localized events. This objection

can be, to a certain extent, dealt with, as it appears that the “asymmetry by fiat” solution is

amenable to some generalizations, either with the help of a specially prepared overall

similarity relation (Finkelstein 1999, Bigaj 2004) or on a case-by-case basis. As for the first

objection, it brings us to the last considered answer to our main problem.

5
By „non-trivial backtracking counterfactuals” I understand counterfactuals that state that if a given present
event was different, some past event would be different too. A trivial backtracking counterfactual is a statement
announcing that if the present was different, the past would still be the same.
10

5. What do we need miracles for? The asymmetry-by-fiat solution was explicitly built for the

case of strict determinism, so it required miracles in order to graft smoothly the counterfactual

supposition onto the unchanged past. But the problem considered in this article arose in the

context of quantum mechanics, where indeterministic events are abundant. Considering a

different outcome of the measurement may not require any modification of the past

whatsoever. This fact might suggest that we could eliminate the troublesome notion of law

breaking once and for all.6 The idea would be to retain criterion (2) as the only condition of

similarity between possible worlds. This means that in order to evaluate (C) we would have to

consider a possible world in which all ordinary laws are obeyed, and which departs from the

actual world at the latest possible moment in order to accommodate for the counterfactual

outcome of the xL-measurement. Obviously in such a world the right-hand side outcome will

be exactly as predicted by (C).

The problem of the arbitrariness of the temporal asymmetry of counterfactuals is to a

certain degree taken care of in this approach. Backtracking counterfactuals with antecedents

describing chance events will typically be false7, but those with antecedents referring to

deterministic events will sometimes come out true. The number of true backtracking

counterfactuals may be slightly too big to Lewis’s taste, but at least we are not forced to

believe that such counterfactuals will stretch indefinitely back in time, because a causal chain

of deterministic events may have its beginning in a chance occurrence.8

6
The main problem with the miracle approach is that it makes laws only contingently true: the statement „It is
possible that law L is false” has to be seen as true in the actual world for all L.
7
Note that in normal discourse we treat human actions as not being uniquely determined by the past, which
explains our inclination to reject counterfactuals of the sort „If I did this rather than that, the past would have
been different”.
8
H. Field in (2003, p. 455) noted correctly that even Lewis’s approach doesn’t eliminate backtracking altogether,
but merely limits it. For it is true on this account that if the present were different, there would have been a small
miracle in the past. I would strengthen this observation by noting that depending on the contrary-to-fact
supposition we are considering, the miracle required to make this supposition true may be necessary to occur
way back in the past, and not just before the required change. If we consider a possible situation at time t that
departs radically from the actual situation, then the only choice is either to use a big miracle to make this
situation happen, or to go back in time sufficiently far to make all the necessary arrangements for it to occur.
11

However, this proposal is open to an immediate criticism. Recall that in section 3 we

have distinguished “genuine” non-local correlations from “common cause” correlations.

Intuitively, only the former support the appropriate counterfactual, while the latter do not.

Unfortunately, in the no-miracle approach in both cases counterfactuals of the (C)-type come

out true. If the correlation between A and B occurs as a result of common cause C, then in

order to consider a counterfactual situation in which A does not happen, we have to “go back”

in time, adjust C accordingly, and in effect this adjustment will bring about a change in the

occurrence of B, securing the truth of the counterfactual “non-A  non-B”.

I see this problem as a serious challenge, but not as an insurmountable obstacle. To

begin with, we have to admit that the intuition of the counterfactual conditional underlying

this approach is slightly different than in Lewis’s original conception. The question whether A

 B is true is interpreted here, roughly, as “Does B hold in a world whose overall history is

adjusted in the least conspicuous way allowing for a lawful occurrence of A?”. This

interpretation reveals that counterfactuals of that sort may not be sufficient for explicating the

causal dependence between non-chance events. Lewis’s account of causality may require an

amendment in the form of an extra condition over and above the counterfactual dependence.9 I

am convinced that it is possible to introduce a condition which would pronounce that the

relation between correlated outcomes in the case when no common cause is present is that of a

causal counterfactual dependence, while in the case with the common cause the relation is a

non-causal counterfactual dependence. Also, this condition would classify most cases of

backtracking counterfactual dependence as non-causal, with a possibility left open that some

backtracking counterfactual dependencies may still be discovered to be causal. Roughly

9
This idea of amending Lewis’s counterfactual analysis of causality occurs in a slightly different context (and for
a different purpose) in (Field 2003, p. 452). Field advocates there a position according to which causality should
be defined in terms of conditional counterfactual dependence (with certain facts being fixed), rather than in terms
of counterfactual dependence simpliciter. Another proposal of emendating Lewis’s account of causality in order
to cope with a slightly different problem called the background condition problem can be found in (Bigaj 2005).
12

speaking, the said condition would require that the effect’s following the cause be relatively

insensitive to the conditions under which the purported cause occurs. In the case of typical

backtracking counterfactuals, the antecedent-event occurring in the present may be brought

about in many different ways, not involving the past occurrence of the consequent-event. On

the other hand, in normal, forward-looking causal relations it is relatively unimportant how

the cause has arisen—once it is there, the effect should follow. The details of this proposal

obviously have to be worked out, but I don’t see any reason why it could not be done.

References

Bedford, D and Stapp, H.P.: 1995, “Bell’s Theorem in an Indeterministic Universe”, Synthese 102, 139-164.
Bell, J.: 1987, Speakable and Unspeakable in Quantum Mechanics, Cambridge University Press, Cambridge.
Bigaj, T.: 2004, “Counterfactuals and Spatiotemporal Events”, Synthese, 142, 1, 1-20.
Bigaj, T.: 2005, “Causes, Conditions, and Counterfactuals”, Axiomathes, 15, 599-619.
Bigaj, T.: 2006, Non-locality and Possible Worlds. A Counterfactual Perspective on Quantum Entanglement,
Ontos Verlag, Frankfurt—Lancaster—New Brunswick, forthcoming.
Clifton R., Butterfield, J., and Redhead, M.: 1990, “Non-local Influences and Possible Worlds”, British Journal
for the Philosophy of Science, 41, 5-58.
Field, H.: 2003, “Causation in a Physical World”, in: M.J. Loux and D.W. Zimmerman (eds.) The Oxford
Handbook of Metaphysics, Oxford University Press, Oxford – New York, 435-460.
Fine, K.: 1975, review of D. Lewis, Counterfactuals, Mind 84, 451-458.
Finkelstein, J.: 1998, “Yet Another Comment on Non-local Character of Quantum Theory”, preprint
quant-ph/9801011.
Finkelstein, J.: 1999, “Space-time Counterfactuals”, Synthese 119, 287-298.
Lewis, D.: 1973, Counterfactuals, Harvard University Press, Cambridge (Mass.).
Lewis, D.: 1986, Philosophical Papers Vol II, Oxford University Press, Oxford.
Maudlin, T.: 1994, Quantum Non-locality and Relativity, Blackwell, Oxford.
Percival, P.: 1999, „A Note on Lewis on Counterfactual Dependence in a Chancy World”, Analysis 59.3, 165-
173.
Redhead, M.: 1987, Non-locality, Incompleteness and Realism, Oxford University Press, Oxford.
van Fraassen, B.C.: 1982, “The Charybdis of Realism: Epistemological Implications of Bell’s Inequality”,
Synthese 52, 25-38.

You might also like