配色: 字号:
Are quantum particles objects
2023-03-20 | 阅:  转:  |  分享 
  
Analysis, 66 (2006) pp.52-63.

Are quantum particles objects?

Simon Saunders

It is widely believed that particles in quantum mechanics are metaphysically

strange; they are not individuals (the view of Cassirer 1956), in some sense

of the term, and perhaps they are not even objects at all, a suspicion raised

by Quine(1976a, 1990). In parallel it is thought that this di¤erence, and es-

pecially the status of quantum particles as indistinguishable, accounts for the

di¤erence between classical and quantum statistics - a view with long historical

credentials.1

?Indistinguishable?here mean permutable; that states of a¤airs di¤ering only

in permutations of particles are the same - which, satisfyingly, are described by

quantum entanglements, so clearly in a way that is conceptually new. And,

indeed, distinguishable particles in quantum mechanics, for which permutations

yield distinct states, do obey classical statistics, so there is something to this

connection.

But it cannot be the whole story if, as I will argue, at least in one notable

tradition, classical particle descriptions may also be permutable (so classical

particles may also be counted as indistinguishable); and if, in that same tradi-

tion, albeit with certain exceptions, quantum particles are bona ?de objects.

1

I will follow Quine in a number of respects, ?rst, with respect to the formal,

metaphysically thin notion of objecthood encapsulated in the use of singular

terms, identity, and quanti?cation theory; second (Quine 1970), in the appli-

cation of this apparatus in a ?rst-order language L, and preferably one with

only a ?nite non-logical alphabet; and third (Quine 1960), in the use of a weak

version of the Principle of Identity of Indiscernibles (PII). Applying the latter

requires a listing of the allowable predicates (the non-logical vocabulary of L);

for our present purposes this should be dictated on theoretical and experimental

grounds, grounds internal to the physics - for example, that only predications

of measurable properties and relations should be allowed. Our minimal, logical

question is then: whether indistinguishable quantum particles are L-discernible

by their measurable properties and relations.

But as a criterion for membership in L, measurability may be somewhat too

restrictive; it threatens to settle our question, negatively, solely on the basis that

quantum particles are unobservable. Better is a condition that is both precise

and more general, namely that only predicates invariant under the symmetries

of the theory qualify. This condition implicitly or explicitly underlines a good

many recent debates in the philosophy of physics over symmetry principles,

1It has been called the ?received view?by French and Rickles (2003).

1

and in important cases (for a number of space-time symmetries) it is physically

uncontroversial.2

In our case the symmetry is the permutation group. Our criterion, then,

is that L-predicates should be invariant under permutations (I shall then call

them symmetric or symmetrized). Whatever the metaphysical questions that

accompany the idea of a ?loss of identity?in quantum mechanics (for indistin-

guishable particles), its sole mathematical signature is permutation symmetry,

the syntactical expression of which (in terms of a regimented formal language)

is surely that predicates be symmetrized. So if it is true that in the words of

an early contributor to quantum statistics ?the conception of atoms as parti-

cles losing their identity cannot be introduced into the classical theory without

contradiction?(Stern 1949: 535) - and if the di¢ culty does not concern the

details of the classical theory but its basic concepts - one would expect it to

show up in our elementary framework of a ?nitary language restricted to to-

tally symmetrized predicates. Bach (1997) indeed takes it as self-evident that

a description of particles having de?nite coordinates can only be permutation

invariant in so far as it is incomplete (specifying only the statistical properties

of a particle ensemble, not the microscopic details).

But is there really any such conceptual impediment? If not - and from

what follows it seems not - the case for metaphysical novelty following on from

particle indistinguishability in quantum mechanics remains unmotivated. Add

to this the argument (Huggett 1999) that classical statistics is every bit as

consistent with permutation symmetry as is quantum statistics, and the claim

that indistinguishability explains quantum statistics looks threadbare indeed.

2

Let F be an nary predicate; the symmetrized language LS that we envisage

must be such that if F 2LS, then in any valuation Fx1:::xn can be replaced by

Fx(1):::x(n) without change of truth value, for any permutation of f1;:::;ng:

We imagine this as our procedure: we start from some language L, based

in part on other physical theories, which lacks permutation symmetry; and

we examine the e¤ects of implementing it, de?ning totally symmetrized LS-

predicates as complex predicates in L (e.g.

_



Fx(1):::x(n), call Fdis; of course

there are plenty of other constructions too - we shall spell out a general one

shortly). Clearly L S L.

How limited is LS? The answer depends on L. Of particular interest is the

case when L has no names, so that singular reference is by means of bound

quanti?cation only (paraphrasing contexts involving names by Russellian de?-

nite descriptions). In fact, restricted to descriptions of particle distributions in

space - as used in specifying the coordinates of particles (or initial or ?nal data

more generally) - it would seem that inde?nite descriptions are enough to be

going on with, of the form ?particles of such-and-such a kind have such and such

properties and relations?. It is not at all clear that in giving such descriptions

2For a review see (Saunders 2003b).

2

one must single out one arrangement of particles, and no permutation of it; for

one has the logical equivalence:

9x1:::9xnFx1:::xn  9x1:::9xnFdisx1:::xn: (1)

It is unlikely that LS and L can di¤er very much for uses like this: what more

does one want to say in L, after listing all the relations among the mentioned

objects, in the form of a purely existential statement? - other that is than which

object (value of variable), for each k, is ak (an illegitimate question, if L has no

names).

In fact when L is devoid of names use of LS involves no restriction at all,

at least in the case of sentences with only ?nite models of a ?xed cardinality

(sayN). Given any suchLsentenceT, one can construct a logically equivalent

LS-sentence TS, true in all and only the same models.

The claim is su¢ ciently surprising to warrant at least a sketch of the proof.

We may suppose, with no loss of generality, that T is given in prenex normal

form (all the quanti?ers, say n in number, to the left). Now construct a sen-

tence T1 in a language L+, which is L supplemented by N names a1;::;aN,

by replacing each rightmost quanti?er and the complex predicate that follows

in T by a disjunction (in the case of 9) or conjunction (in the case of 8) of

formulas in each of which xn is replaced by a name (yielding

N_

k=1

Fx1:::xn1ak

or

N^

k=1

Fx1:::xn1ak respectively). ). Repeat, removing each innermost quanti-

?er, obtaining at each step a complex predicate completely symmetric in the N

names (ensured only because T has no names); on the ?nal step one obtains a

sentence T2, in which the xks do not occur. Now replace every occurrence of ak

by xk to obtain a totally symmetrized Nary L-predicate, which is therefore in

LS . Prefacing withN existential quanti?ers (and conjoining with a cardinality-

?xing sentence) one obtains the sentence TS; by construction it has the same

truth conditions as T:

The two languages, insofar as they are used to describe a ?nite collection of

objects, are in this sense strictly equivalent; under this condition, symmetriza-

tion makes no di¤erence to truth values of sentences. I suggest this is evidence

enough that indistinguishability in itself indicates nothing metaphysically unto-

ward, or otherwise strange.

Why might one have thought any otherwise? But the constraint is certainly

prohibitive applied to ordinary language; take the predicate ?... is in the kitchen,

not....?, as in:

(i) Bob is in the kitchen, not Alice.

To symmetrize and say instead:

(ii) Bob is in the kitchen and not Alice, or Alice is in the kitchen and not Bob

3

doesn?t tell us what we want to know. But in a language su¢ ciently rich in

predicates to replace ?Bob?and ?Alice?by de?nite descriptions, the situation is

quite di¤erent. We then have a sentence like:

(iii) There is someone Bob-shaped who is in the kitchen, and someone Alice-

shaped who is not

(where ?Bob-shaped?etc. is shorthand for some purely geometric, anatomical

description). Symmetrizing, in LS we say instead:

(iv) There is x1 and x2, where x1 is Bob-shaped and in the kitchen and x2 is

Alice-shaped and not, or x2 is Bob-shaped and in the kitchen and x1 is

Alice-shaped and not.

Unlike the passage from (i) to (ii), there is no di¤erence between (iii) and (iv);

they are an instance of the equivalence (1)). One might of course reintroduce

the question of which of x1 and x2 is which (say, which of two persons, speci?ed

independent of their appearance), but that is only to invite further de?nite

descriptions, whereupon we will be back to the same equivalence.

3

This argument would all by itself settle the matter, were it not for the worry

that the objects that we end up with - the values of x1 and x2; that only

contingently have the bodies or personalities (or what have you) that they do -

are themselves rather strange. It may be they are just as problematic, when it

comes to questions of identity, as quantum particles.

We should face this challenge head on. The account of identity that follows

applies to any ?rst-order language L without equality, for any ?nite non-logical

alphabet, whether or not symmetrized.

Any such L e¤ectively comes with identity, a point that Quine has often

emphasized. We get for free the de?ned sign:

s = t =

def

^

all primitive L-predicates

Fs$Ft (2)

(here s and t are L-terms, occupying the same predicate position in F). Un-

packing this schematic de?nition, and temporarily introducing the notation Fnk

for the k-th nary predicate symbol of L, we obtain on universal generalization

over free variables not in s, t:

s = t =

def

^

n

^

k

n^

j=1

8x1:::8xn(Fnk x1::xj1sxj+1::xn $Fnk x1::xj1txj+1::xn):

(3)

The RHS is of the form ^

88:::8(Fs$Ft) (4)

4

and not:

^

(88:::8(Fs) $ 88:::8(Ft)) (5)

the point that so often goes unstated.3 By construction the schemes variously

written as (2),(3),(4) (but not (5)) imply the usual axiom scheme for identity:

s = s; s = t! (s$ t)

(where  can be replaced by any L-predicate, primitive or otherwise); moreover

any other scheme with implies the latter yields an equality sign coextensive with

the one de?ned, so identity as given by (2), (3), (4) is essentially unique.

But isn?t this just the familiar PII? - yes, or it should be familiar. In fact

it has received surprisingly little attention, despite its endorsement by Quine.

A correct formal classi?cation of L-discernibles, according to this scheme, was

only given quite recently.4 And Quine made no applications of the principle to

physically problematic cases (nor, so far as I know, has anyone since).

Quine?s amended classi?cation is (I follow his earlier terminology for the ?rst

two cases):

Two objects are

 absolutely discernible inLif there is anL-formula in one free variable that

applies to one of them only

 relatively discernible inLif there is anL-formula in two free variables that

applies to them in one order only

 weakly discernible in L if they satisfy an irre?exive L-formula in two free

variables.

As stated each category contains the one before it (here I follow Quine), but

they are exhaustive: values of variables not even weakly discernible are counted

(in L) as the same.

The interesting cases are mere relative or weak discernibility. For example,

let the only non-logical symbol in L be an irre?exive and symmetric dyadic F;

then from the de?nition (3):

x = y $ 8z((Fxz $Fyz):

On any valuation in whichFxy is true, 8z(Fxz $Fyz) is false (asFxy^:Fyy

is true); it follows that x 6= y. Thus, to take Black?s famous example of two

spheres of iron, positioned in an otherwise empty universe, one mile apart in

space; they are weakly discerned by the symmetric and irre?exive relation ?one

mile apart in space?; but they are neither absolutely nor relatively discernible.5

3It is discussed at length by Quine (1976b).

4By Quine in 1976, amending, without comment, the classi?cation he gave in Word and

Object in 1960.

5For other physical examples, see Saunders (2003a) and, in mathematics, Ladyman (2005).

5

What now of names? Let L contain names, and the categories of relative

and weak discernibility seem to be obliterated. For named objects, if discernible

at all, are absolutely discernible.6 However, everything turns on the proviso, if

discernible; whether or not the names do in fact name di¤erent objects will still

depend, just as before, on predicates alone. In this sense, then, named objects

can still be classi?ed as absolutely, relatively, or only weakly discernible, just as

before.

4

We are ?nally ready to answer our question: Are permutable particles dis-

cernible? The answer depends, evidently, on LS, and speci?cally its non-logical

vocabulary. In quantum mechanics the state of a particle is speci?ed by a vector

'', up to a complex multiplicative constant, or phase, in the state-space of the

system (Hilbert space) . AnNparticle state is a sequence of 1-particle vectors,

or - this the essential di¤erence from classical theory - a sum of such sequences.

A state of a collection of quantum particles, if the particles are indistinguishable,

must be invariant under the permutation group. Among these are expressions

of the form (for a 3particle state):

const:('' +'' + ''+ ''+'' + '') (6)

where ''; ; are 1particle vectors. Pretty evidently, it does not specify which

particle is in which state - there is no such determinate rule here. It is like the

symmetrized triadic ?the ?rst particle is in the state ?const. ''?the second in

the state ?const. ?, the third in the state ?const. ?, or the ?rst particle is in

the state ?const. ''?, the second in the state ?const. ?, the third in the state

?const. ?, or ....?(evidently sequence positions are here functioning as names).

Such states (permuting and summing) are also called symmetrized; the particles

described by them are bosons.

But what are we to make of the allegedly 3-particle (and manifestly sym-

metrized) state ?const. ''''''?? Evidently, that there are three particles each in

exactly the same 1-particle state, and therefore exactly alike in every respect.

They are surely not absolutely discernible, hence, since relative discernibles re-

quire non-symmetric predicates, they are at most weakly discernible. But are

they even that? What (physical, invariant) relation does a boson enter into with

a boson in exactly the same state, supposed to be a complete description, that

it does not enter into with itself?

If the answer is none (as it appears), or none that can be sanctioned by the

physical theory, then either the PII, or the objectual status of quantum particles,

is in question. Were that the end of the story, either way our total system

would be in trouble. In fact there is another possibility - another prescription

under which the state is invariant under permutations: vectors may instead be

antisymmetric, changing sign on any odd number of interchanges of particles

6If discernible at all, then they are at least weakly discernible; so there is a totally symmetric

and totally irre?exive Nary predicate F such that the sentence Fa1:::aN is true. But then

Fa1::ak1xak+1:::aN absolutely discerns ak.

6

(the state itself - the vector up to phase - remaining unchanged). The particles

of antisymmetrized states are fermions. In place of (6) we have

const:('' '' + '' ''+''  ''):

Evidently antisymmetric states cannot assign two di¤erent particles to exactly

the same 1-particle state; antisymmetrizing ?''''''?produces the zero. The prob-

lem we encountered with bosons does not arise.

Antisymmetrization ensures Pauli?s exclusion principle (the principle that

fermions cannot have all their quantum numbers in common). The latter was

indeed early on considered, by Weyl among others, to vindicate the PII,7 but

the suggestion was squashed by Margenau in 1944 (and seems to have been

hardly advocated since). Margenau came up with a new argument to show

that fermions are indiscernible, namely, that all the 1-particle expectation values

(which may be taken as exhausting the 1-place predications) of any fermion in

an antisymmetrized state must be the same. This was thought to show that the

PII cannot apply.8

But discernibility does not require absolute discernibility; and if one con-

siders the remaining candidates, relative or weak discernibility, it seems that

Margenau was wrong and Weyl was right all along. For even in a situation of

maximal symmetry, for example in the singlet state of spin

= const. (''"''# ''#''") (7)

the two particles are still weakly discernible. Here ''"''# correspond to the two

opposite possible values (parallel or antiparallel) of the spin of the particle along

a given direction ". Here any direction can be chosen, without change of the

state - it is in this sense that (7) is a specially symmetric state, invariant under

rotations as well as permutations - still the two particles satisfy the symmetric

but irre?exive predicate ?... has opposite " component of spin to:::?.9

Why was this simple observation missed? The answer, presumably, is that

it would then seem that the particles must each have a de?nite and opposite

value for the " component of spin, implying some kind of hidden-variable

interpretation of quantum mechanics (contentious in itself, for entirely unrelated

reasons). But this is to fall back on our old habit of turning discernment on

the basis of relations into discernment by di¤erences in properties (?relational

properties?); it is to miss the logical categories of relative and weak discernibility.

Consider again Black?s two iron spheres, each exactly alike, but one mile

apart in space. They are weakly discerned by the irre?exive relation ?...one

mile apart from...?, but - on pain of begging the question against relationism

- it does not follow, because the spheres bear spatial relations to each other,

that they each have a particular position in space. Neither, if two lines are

weakly discerned by the irre?exive relation ?at right-angles?, does it follow that

each line has a particular direction in space. Two particles can have opposite "

7It was called the ?Leibniz-Pauli?principle by Weyl (1949: 247).

8Similar arguments have since been given by French and Redhead (1988) and Dieks (1990).

9For more formal details, see Saunders (2003a, 2006).

7

component of spin (they are anticorrelated as regards spin in the" direction)

without each having a particular value for the "-component of spin.

On the strength of this we can see, I think, the truth of the general case:

so long as the state of an Nfermion collective is antisymmetrized, there will

be some totally irre?exive and symmetric N-ary predicate that they satisfy.

Fermions are therefore invariably weakly discernible.

Not only are fermions secured; so too, concerning the atomic constituents of

ordinary matter, are bosons. For all but one of the stable bosons are composites

of fermions (the exception is the photon). In all these cases, the bosonic wave-

function (with its symmetrization properties) is an incomplete description, and

at a level of ?ner detail - irrelevant, to be sure, to the statistical properties of

a gas of such composites - we have a collection of weakly discernible particles.

By reference to the internal structure of atoms, if nothing else, we are assured

that atoms will be at least weakly discernible.

The only cases in which the status of quantum particles as objects is seriously

in question are therefore elementary bosons - bosons (supposedly) with no in-

ternal fermionic structure. The examples in physics (according to the Standard

Model) of truly elementary bosons are photons and the other gauge bosons (the

W and Z particles and gluons) and the conjectured (but yet to be observed)

Higgs boson. But in these cases there is a ready alternative to hand for object

position in sentences: the mode of the corresponding quantum ?eld. We went

wrong in thinking the excitation numbers of the mode, because di¤ering by

integers, represented a count of things; the real things are the modes.10

5

The answer to Quine?s question - Are quantum particles objects? - is

therefore: Yes, except for the elementary bosons.

Similar conclusions follow in the classical case. If indistinguishable, and

permutations are symmetries, we should speak of them using only symmetrized

predicates. If impenetrable they will be at least weakly discerned by the irre?ex-

ive relation ?...non-zero distance from...?; but even if one relaxes this assumption,

and allows classical particles to occupy the same points of space, they may still

be (relatively) discerned by their relative velocities. Problems only arise if rela-

tive distances and velocities are zero, in which case, if no more re?ned description

is available, they will remain structureless and forever combined, and we would

do better to say there is only a single particle present (with proportionately

greater mass). This, a classical counterpart to elementary bosons, makes the

similarities in the status of particles in classical and quantum mechanics only

the closer.

What of the more metaphysical question, of whether quantum particles are

individuals? But here it is not clear what more is required of an object if it is

to count as an individual: perhaps that it is not permutable, or that it is always

absolutely discernible, or discernible by intrinsic (state-independent) properties

and relations alone. But in all cases, one is no closer to an explanation, in

10A suggestion ?rst made by Erwin Schr?dinger in 1924.

8

logical terms, of the di¤erence between classical and quantum statistics, for

none of these distinctions cut along lines that demarcate the two.

The facts about statistics are these:11

 Distinguishable classical particles obey classical statistics.

 Indistinguishable classical particles obey classical statistics.

 Distinguishable quantum particles obey classical statistics.

 Indistinguishable quantumparticles obeyquantumstatistics(Bose-Einstein

or Fermi-Dirac statistics12).

Distinguishable particles in physics we may take to be absolutely discernible,

and in all cases they obey classical statistics; but indistinguishable particles,

particles ensured only to be weakly discernible, may or may not obey classical

or quantum statistics. No more does the discernible/indiscernible distinction

line up with the classical/quantum divide; it only serves to distinguish between

certain classical and quantum particles, on the one hand, and the elementary

bosons (and their classical analogues) on the other. And ?nally, names do not

capture the distinction; given the restriction to totally symmetrized predicates,

the presence or absence of names is irrelevant.

Is there some other dimension along which one might mark out a distinctive

status for indistinguishable quantum particles? Perhaps - say, in whether or not

quantum particles are re-identi?able over time (as argued by Feynman (1965)).

But this takes us away from permutation symmetry per se, and there are many

classical objects (shadows, droplets of water, patches of colour) that likewise

may not be identi?able over time. In the weakly interacting case, taking the 1-

particle states that enter into a symmetrized or antisymmetrized state as objects

instead, one may or may not have things reidenti?able over time,13 and yet the

statistics remain the same. But the overriding objection, in the present context,

is that in considerations like these we seem to be getting away from the purely

logical notion of identity.

It seems the only remaining alternative, if indistinguishability is to have the

explanatory signi?cance normally accorded it, is to deny that permutability is

intelligible at all as a classical symmetry - that it is simply a metaphysical mis-

take, on a traditional conception of objects, to think that particles can be really

indistinguishable. One would then be left with the clean equation: permutable

if and only if quantum mechanical.

But the claim is implausible, as we are in a position to see. Finitary, cate-

gorical descriptions in LS, that are restricted to totally symmetrized predicates,

11In all cases the entropy is extensive (even for closed systems) if and only if the particles are

indistinguishable. For arguments that extensivity (and hence indistinguishability) is strictly

required for closed systems, even in classical thermodynamics, see Pniower 2006, Saunders

2006.

12There is also the possibility of parastatistics (not so far experimentally detected), involving

mixed boson and fermion transformations.

13Depending on whether or not the total state is a superposition of states of de?nite occu-

pation numbers.

9

are logically equivalent to those in L that omit only names. Descriptions of

the latter sort, whatever their philosophical inadequacies, can hardly be called

unintelligible.

In the face of this, our conclusion is rather that indistinguishability has

nothing at all to do with the quantum and classical divide, and that the reason

for quantum statistics, in the face of permutability, must be sought elsewhere.14

University of Oxford, 10, Merton St.,

Oxford OX1 4JJ

simon.saunders@philosophy.ox.ac.uk

Acknowledgements

My thanks to Guido Bacciagaluppi and Justin Pniower for helpful discussions,

and to an anonymous referee for a number of constructive criticisms.

References

Bach, A. 1997. Indistinguishable Classical Particles. Berlin: Springer-Verlag.

Cassirer, E. 1956. Determinism and Indeterminism in Modern Physics. New

Haven: Yale University Press.

Dieks, D. 1990. ?Quantum statistics, identical particles and correlations?, Syn-

these 82: 127-55.

Feynman, R. 1965. The Feynman Lectures on Physics, Vol. 3: Quantum Me-

chanics, R. Feynman, R. Leighton, and M. Sands, eds. Reading: Addison-

Wesley.

French, S., and M. Redhead 1988. ?Quantum physics and the identity of indis-

cernibles?, British Journal for the Philosophy of Science 39: 233-46.

French, S., and D. P. Rickles 2003. ?Understanding permutation symmetry?, in

K. Brading and E. Castellani, Symmetries in Physics: New Re?ections. Cam-

bridge: Cambridge University Press.

Huggett, N. 1999. ?Atomic metaphysics?, Journal of Philosophy 96: 5-24.

Ladyman, J. 2005. ?Mathematical structuralism and the identity of indis-

cernibles?, Analysis 65: 218-21.

Margenau, H. 1944. ?The exclusion principle and its philosophical importance?,

Philosophy of Science 11: 187-208.

Pniower, J. 2006. Particles, Objects, and Physics. D. Phil thesis, University of

Oxford.

Quine, W. V. 1960. Word and Object. Cambridge: Harvard University Press.

Quine, W. V, 1970. Philosophy of Logic. Cambridge: Harvard University Press.

Quine, W.V. 1976a. ?Whither physical objects??, in Essays in Memory of Imre

Lakatos, R. S. Cohen, P. Feyerabend, and M. Wartofsky, eds. Dordrecht: Reidel.

Quine, W.V. 1976b. ?Grades of discriminability?, Journal of Philosophy 73;

reprinted in 1981, Theories and Things. Cambridge: Harvard University Press.

Quine, W.V. 1990. The Pursuit of Truth. Cambridge: Harvard University

Press.

14For further discussion, see Saunders (2006).

10

Saunders, S. 2003a. ?Physics and Leibniz?s principles?, in K. Brading and E.

Castellani, Symmetries in Physics: New Re?ections. Cambridge: Cambridge

University Press.

Saunders, S. 2003b. ?Indiscernibles, general covariance, and other symmetries:

the case for non-reductive relationalism?, in Revisiting the Foundations of Rela-

tivistic Physics: Festschrift in Honour of John Stachel, A. Ashtekar, D. Howard,

J. Renn, S. Sarkar, and A. Shimony, eds. Amsterdam: Kluwer.

Saunders, S. 2006. ?On the explanation for quantum statistics?, Studies in the

History and Philosophy of Modern Physics, forthcoming.

Stern, O. 1949. ?On the term klnn! in the entropy?, Reviews of Modern Physics

21: 534-35.

Weyl, H. 1949. Philosophy of Mathematics and Natural Science. Princeton:

Princeton University Press.

11

献花(0)
+1
(本文系mc_eastian首藏)