C:\Documents and Settings\Frank\My Documents\psapqm.wpd 1 IS QUANTUM MECHANICS POINTLESS? FRANK ARNTZENIUS RUTGERS UNIVERSITY ABSTRACT There exist well-known conundrums, such as measure theoretic paradoxes and problems of contact, which, within the context of classical physics, can be used to argue against the existence of points in space and space-time. I examine whether quantum mechanics provides additional reasons for supposing that there are no points in space and space-time. 2 IS QUANTUM MECHANICS POINTLESS? FRANK ARNTZENIUS RUTGERS UNIVERSITY 1. Introduction Our standard account of regions and their sizes, has some bizarre features. In the first place one can not cut a region exactly in two halves. For if one of the two regions includes its boundary (is closed), then the other does not include it (is open). One might reasonably think that this difference between open and closed regions is an artifact of our mathematical representation of regions which does not correspond to a difference in reality. Secondly, regions of finite size are composed of points, each of which have zero size. One might think it rather strange that when one gathers together countably many points one must have a region of size zero, while if one gathers together uncountably many points, one can form a region of any size. Thirdly, finite sized regions must have parts which have no well-defined size, i.e. are unmeasurable. One might swallow parts that have zero size, but parts that can not have any well-defined size, this could lead to gagging. Fourthly, Banach and Tarski have shown that one can break any finite sized region into finitely many parts which can then be re-assembled, without stretching or squeezing, to form a larger (or smaller) region. And then there are also problems about contact: physical objects which occupy closed regions can never touch, indeed they must always be a finite distance apart. Now, I do not say that problems such as these are a decisive argument against the standard account. But I do say that they form a good reason to devise a geometry that does not have these problems, and to see whether modern physics can plausible be set in such a geometry. Caratheodory, and others following him, have devised such “pointless geometries”. (See Caratheodory 1963, Skyrms 1993). Let me give an example of such a pointless geometry. Start 3 by designating the collection of all open intervals on the real line as regions. Then declare that the union of any countable set of regions is a region, declare that the intersection of any two regions is a region, declare that the complement of any region is a region, and declare that these are all the regions that there are. This is collection of regions is the so-called “Borel algebra” of regions. Now this collection of regions includes point-sized regions, regions that differ only in being open or closed, and more generally distinct regions whose differences have size 0. Let us get rid of all such differences by regarding as equivalent any regions such that the differences between those regions have size 0. I.e. let us declare Regions to be equivalence classes of regions that differ at most by (Lebesque) measure 0. This collection of Regions, and their sizes, comprises an example of a pointless geometry. Since any distinct points differ by measure 0, all points will correspond to one and the same Region, namely the “Null Region”, which is the complement of the Region consisting of the entire space. Any other Region has well-defined finite size (measure). Breaking up and re-assembling never changes the size of a Region. Regions can always be cut exactly in half. And there are no problems about contact between objects since there are no differences between open and closed Regions. This seems very pleasing. It therefore seems worthwhile to examine whether physics can be done in such a setting. In this paper I will take a look at quantum mechanics. I will argue that the formalism of quantum mechanics strongly suggests that its value spaces, including physical space and space-time, are pointless spaces. 2 Continuous observables in quantum mechanics. It is will known that, strictly speaking, on the standard account of the state-space of 4 quantum mechanics as a separable Hilbert space, continuous observables do not have eigenstates. For instance, there exists no quantum mechanical state |x=5> which is an eigenstate of the position operator X corresponding to the point x=5 in physical space. Indeed there exist no quantum mechanical state such that a measurement of position in that state will, with probability equal to one, yield a particular value. For if there were position eigenstates there would have to be uncountably many mutually orthogonal states, but a separable Hilbert space has only countably many dimensions. What is not often noted is that there is a more general conclusion that can be drawn from the assumption that the quantum mechanical state-space is a separable Hilbert space, namely that wave-functions are functions on pointless spaces. To be more precise, it is a consequence of the fact that wave-functions are representations of states in a separable Hilbert space that each wave- function is not simply a square integrable function, but rather an equivalence class of square integrable functions which differ in their values at most on a set of (Lebesque) measure 0. The reason for this is pretty straightforward. One of the axioms of the theory of Hilbert spaces is that there is a unique vector whose norm (inner product with itself) is zero. In the position representation, the norm of a wave-function f(x) is I|f(x)|2dx. But there are many different functions for which I|f(x)|2dx=0. So, in order for wave-functions to be representations of vectors in a Hilbert space one needs to assume that wave-functions correspond to equivalence classes of (square integrable) functions that differ at most on a set of measure 0. Now one can show mappings (homomorphisms) on pointless spaces correspond exactly to equivalence classes of functions that differ at most on a set of measure 0 (see Skyrms 1993). Thus wavefunctions are functions on pointless spaces. Quantum mechanics thus provides us with evidence that the 5 value-space for any continuous observable is a pointless space. However, let me now turn to two ways in which point values for continuous observables can be re-introduced into quantum mechanics. 3 Rigged Hilbert spaces There is a standard way of re-introducing eigenstates of continuous observables in a rigorous way, namely the “rigged Hilbert space” formalism. Let me outline this formalism. (For more detail see Böhm 1978). Let’s use the simplest example, the harmonic oscillator. I will assume that the reader is familiar with the construction of the “ladder” of eigenstates Nn=(a +)nN0 /%n! of the number operator N, which starts “at the bottom” with the state N0 which has the feature that NN0=0. Let us now consider all and only the finite superpositions of these states, i.e. the states of form N='cnNn , where we superpose only finitely many Nn. Let us denote this linear space of states as Q. Using the standard scalar product (N,R) and norm |R|2=(R,R) one can then define the standard Hilbert space topology on the space Q, and the accompanying standard notion of convergence: Nk6N iff |Nk-N|60 as k64. Given this topology the space Q is not “complete”, i.e. there exist Cauchy sequences (converging sequences) that have no limit point in Q. If one now completes Q by adding all such limit points, one obtains the standard Hilbert space H of the harmonic oscillator. It is important to note that this has as a consequence that the Hilbert space H will contain “infinite energy” states: there will exists Cauchy sequences of states N1=c1E1, N2=d1E1+d2E2, N3=e1E1+e2E2+e3E3, ........., such that as n64, the expectation value of Energy=(1/3|ci| 2)(3|ci| 2Ei)64. (each Ei denotes an energy eigenstate). By the completeness of the 6 Hilbert space H there must exist a limit state corresponding to each Cauchy sequence. Hence there will exist a state that one can reasonably call an “infinite energy” state, even though this state, strictly speaking, is not in the domain of the energy operator. Let us now define a different topology, a “nuclear” topology, on Q and the accompanying different, “nuclear”, notion of convergence: Nk6N iff ((Nk-N),(N+1) p(Nk-N))60 as k64 for any p. Roughly speaking, the factor (N+1)p is a factor designed to weigh the higher number eigenstates heavier than the lesser number eigenstates, so that differences in the higher number coefficients have to converge to 0 very rapidly if the norm ((Nk-N),(N+1) p(Nk-N)) is to converge to 0 as k converges to infinity. Thus any sequence of states in Q that is a Cauchy sequence according to the nuclear topology is also a Cauchy sequence according to the Hilbert space topology, but not vice versa. Now let us complete Q according to the nuclear sense of convergence. Of course, this will add only a proper subset of the states that get added when one completes Q according to the Hilbert space topology. We then obtain a “linear topological” space of states M. It is interesting to note that M does not contain infinite energy states. The reason for this is that the coefficients of higher number (higher energy) states have to drop to 0 very rapidly (faster than any polynomial) in order for the sequence to be a Cauchy sequence according to the Nuclear topology. This might seem to be a rather appealing feature of space M. We need just a little more machinery in order to construct such point valued states. A so- called “anti-linear functional” F on a linear space 1 is a function F(2), often denoted as <2|F>, from elements 2 of 1 to complex numbers, such that =c1*<21|F>+c2*<22|F>. (Here the ci denote complex numbers, and * denotes complex conjugation.) The space 1 X of 7 linear functionals on a linear space 1 is linear itself, and is called the space “conjugate to” 1. It is easy to see that each vector f in a linear space 1 with a scalar product (2,0) defines an anti- linear functional F as follows: <2|F>=(2,f). It is also fairly easy to show that for a Hilbert space H, there is a 1-1 correspondence between anti-linear functionals |0> and vectors <0|, so that H and HX can be taken to be the same space. This, however, is not true for the space M that we obtained from Q by completing it according to the nuclear topology. Rather, one can show that MdHdMX. This triplet of spaces is known as a “rigged Hilbert space”, or a “Gelfand Triplet”. Corresponding to any continuous linear operator A on states in M there exist an adjoint operator AX on states in MX, which is defined by the demand that =def= for all . Now we can define so-called “generalized” eigenvectors of an operator A on M. A “generalized” eigenvector of A corresponding to the “generalized” eigenvalue 8 is an anti- linear functional F,MX such that: ==8* for all =8*|F>. One can then show that, for our harmonic oscillator system, there are a continuum of generalized eigenvalues and eigenvectors of both the X and P operators. And one can show, for our harmonic oscillator system, that any state |N> in MX which corresponds to a state of the position operator X, and a unique expansion in terms of a measure over the generalized eigenvectors |p> of the momentum operator P. This all seems great. Let us now consider some unappealing features of rigged Hilbert spaces. A rigged Hilbert space, i.e. a Gelfand triple MdHdMX, is a not as simple and natural a state-space as a Hilbert space. Just look at the machinery that I needed above in order to explain the basics of rigged Hilbert spaces, and compare it to the simplicity and naturalness of (the 8 axioms of) the normal (separable) Hilbert space formalism. Moreover, a rigged Hilbert space is a rather non-unified, cobbled together, state-space which consists of 3 quite distinct parts M, H and MX, where states in the distinct parts have distinct properties. For instance, given any two states N and R in H, one can take their scalar product , which is a complex number. But the scalar product of states f and g that are in MX but not in H, is not an ordinary complex number. The scalar product in MX exists only in a distributional sense, i.e. it is defined as the distribution which satisfies =Idg for all N in M. And there is the awkward, but essential, use of two distinct topologies, the one corresponding to the usual inner product, the other being the “nuclear” topology. It’s all rather messy. A more serious problem is the following. Since can not spectrally decompose a position eigenstate in terms of the eigenstates of such an observable, one can not make sense of probabilities of the results of a measurement of such an observable when the object is in a position eigenstate. More generally, in a state f one can only make sense of the ratios of expectation values / of ‘admissable’ observables A and B, where an observable A is said to be admissable iff A|f> belongs to the domain of