This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

QUANTUM 𝔽𝐮𝐧\bm{\mathbb{F}_{\mathrm{un}}} :
THE q=1q=1 LIMIT OF GALOIS FIELD QUANTUM MECHANICS, PROJECTIVE GEOMETRY & THE FIELD WITH ONE ELEMENT

LAY NAM CHANG Department of Physics, Virginia Tech, Blacksburg, VA 24061, USA
laynam@vt.edu
   ZACHARY LEWIS Department of Physics, Virginia Tech, Blacksburg, VA 24061, USA
zlewis@vt.edu
   DJORDJE MINIC Department of Physics, Virginia Tech, Blacksburg, VA 24061, USA
dminic@vt.edu
   TATSU TAKEUCHI Department of Physics, Virginia Tech, Blacksburg, VA 24061, USA , and
Kavli Institute for the Physics and Mathematics of the Universe (WPI),
The University of Tokyo, Kashiwa-shi, Chiba-ken 277-8583, Japan
takeuchi@vt.edu
Abstract

We argue that the q=1q=1 limit of Galois Field Quantum Mechanics, which was constructed on a vector space over the Galois Field 𝔽q=G​F​(q)\mathbb{F}_{q}=GF(q), corresponds to its ‘classical limit,’ where superposition of states is disallowed. The limit preserves the projective geometry nature of the state space, and can be understood as being constructed on an appropriately defined analogue of a ‘vector’ space over the ‘field with one element’ 𝔽1\mathbb{F}_{1}.

keywords:
Field with one element; quantum mechanics; classical mechanics; projective geometry, Galois field
\ccode

03.65.Aa,03.65.Ta,03.65.Ud

1 Introduction - the Classical Limit of Quantum Mechanics

How deterministic classical mechanics (CM) emerges from probabilistic quantum mechanics (QM) is a conundrum which is yet to be resolved. As first noticed by Dirac[1], the Poisson brackets of CM can be considered the ℏ→0\hbar\rightarrow 0 limit of commutators in QM, and the Heisenberg equation goes over to Hamilton’s equation in that limit:

d​A^d​t=1i​ℏ​[A^,H^]→ℏ→0d​Ad​t={A,H}.\dfrac{d\hat{A}}{dt}\;=\;\dfrac{1}{i\hbar}[\,\hat{A},\,\hat{H}\,]\qquad\xrightarrow{\hbar\rightarrow 0}\qquad\dfrac{dA}{dt}\;=\;\{\,A,\,H\,\}\;. (1)

Despite this formal correspondence, however, it is far from clear how the full classical theory would be recovered from QM in such a limit. In particular, what happens to the Hilbert space of QM in the ℏ→0\hbar\rightarrow 0 limit is a difficult question.111The singularity of the ℏ→0\hbar\rightarrow 0 limit has been emphasized in an illuminating review by Berry[2]. Also, ℏ\hbar being a dimensionful quantity, the limit ℏ→0\hbar\rightarrow 0 itself is not particularly well-defined.222Given the dimensionfulness of ℏ\hbar, the ℏ→0\hbar\rightarrow 0 limit should be understood as ℏ≪S\hbar\ll S where SS is the classical action that is used to describe the quantum formulation as, say, in the path integral approach.

In order to see the CM↔\leftrightarrowQM correspondence more clearly, several attempts have been made to make CM look more like QM via the introduction of wave-functions, operators, and probability distributions. The WKB approximation[3, 4, 5] to the Schrödinger equation has been reinterpreted as the wave-function of CM by Van Vleck[6] and subsequently by Schiller,[7, 8, 9] with observables represented by commuting hermitian operators with continuous eigenvalues. However, utilizing the rewriting of the Schrödinger equation by de Broglie in his pilot-wave theory,[10] which was later elaborated on by Bohm to discuss a hidden variable interpretation of QM,[11, 12] Rosen has shown that taking the ℏ→0\hbar\rightarrow 0 limit of the Schrödinger equation does not necessarily recover the Hamilton-Jacobi equation for superpositions of states.[13, 14, 15]

Koopman[16] and von Neumann[17, 18], and later Sudarshan[19], developed a complete formulation of CM on a Hilbert space, which was again characterized by commuting hermitian operators as observables. Recent work on extending the Koopman-von Neumann-Sudarshan formalism with the introduction of path integrals, etc. include Refs. \refciteGozzi:2001he,Mauro:2001rm,Mauro:2003,Gozzi:2003sh,Carta:2005fq. It should be emphasized that the Koopman-von Neumann-Sudarshan theory is not the ℏ→0\hbar\rightarrow 0 limit of QM. Indeed, the operators that correspond to position and momentum in their formalism commute with each other without the taking of any limit, and respectively have canonically conjugate operators with which they do not commute but are deemed unobservable. The superposition of states also correspond to ensembles of classical states, and not macroscopic Schrödinger-cat like states. Thus, though the Koopman-von Neumann-Sudarshan theory is a formulation of CM on a Hilbert space, it does not (yet) provide much insight on how CM can emerge from QM, or vice versa.

Given the apparent absence of macroscopic Schrödinger-cat like states, they must somehow vanish in the classical limit. Several approaches have been used to address this problem, the most prominent being that of ‘environmental decoherence’ reviewed in Refs. \refciteZurek:2003,Zurek:2003zz by Zurek. There, the system under observation and its environment (rest of the universe) are both treated quantum mechanically, and it has been shown that the interaction between the two suppresses the off-diagonal terms in the density matrix in a preferred basis. The statistical nature of the theory remains, however, and leads to the many-worlds interpretation of QM developed from the pioneering work of Everett[27]. Other approaches to the system-environment interaction problem treat the environment classically, and the interaction with the quantum system leads to classical-quantum ‘hybrid’ theories, the properties of which are still under investigation.333See Refs. \refciteSudarshan:1976,Sherry:1978ea,Sherry:1979vc,Gautam:1978ws,Peres:2001,Terno:2004ti,Elze:2011hi,Elze:2012fk,Elze:2013una,Jauslin:2011,Salcedo:2012qx,Barcelo:2012ja and references therein.

Thus, despite impressive developments in our understanding of QM during the past century, the CM/QM divide remains and bridging that gap is still an intense area of investigation. What seems probable is that recovering CM from QM would require more than taking the ℏ→0\hbar\rightarrow 0 limit of QM in some yet to be discovered way, or the rewriting of CM or QM to look more like the other in some fashion. A more generic theory which encompasses both CM and QM may be necessary to understand the CM↔\leftrightarrowQM correspondence.

As with any difficult problem, finding a toy analog which simplifies the situation while maintaining the essence of the quandary is often instructive. In Refs. \refciteChang:2012eh and \refciteChang:2012gg we constructed a toy analog of NN-level QM on the vector space 𝔽qN\mathbb{F}_{q}^{N}, where 𝔽q=G​F​(q)\mathbb{F}_{q}=GF(q) is the Galois field of order q=pnq=p^{n}, with pp prime and n∈ℕn\in\mathbb{N}, which we dubbed “Galois Field Quantum Mechanics” (GQM).444See also Refs. \refciteMQT,MQT2,MQT3,MQT4,James:2011,Willcock:2011,Hanson:2011,Hanson:2012,Hanson:2013,Chang:2012we,Takeuchi:2012mra,Chang:2013joa,Ellerman:2013,Ellerman:2014. GQM was necessarily different from canonical QM in many ways, due to the vector space 𝔽qN\mathbb{F}_{q}^{N} not possessing an inner product. Consequently, observables were not represented by operators, hermiticity being difficult to define without an inner product555It is possible to define analogs of hermitian operators using biorthogonal systems. See Refs. \refciteChang:2012we,Curtright:2005zk,Curtright:2006zv., and without operators there were no commutators, or ℏ\hbar for that matter. However, in it, physical states were still represented by elements of a projective geometry, and the theory still predicted probabilities of the outcomes of a measurement which could not be mimicked by any hidden variable theory. Thus GQM, despite being constructed on a discrete and finite vector space 𝔽qN\mathbb{F}_{q}^{N} without an inner product, nevertheless captured some of the quantum-ness of canonical QM. A natural question to ask then is: does GQM have a ‘classical’ limit in which this quantum-ness is lost and replaced by classical-ness? An answer to this could help us understand how canonical QM becomes CM also.

In this paper, we present the observation that even though GQM does not have an ℏ\hbar in its formulation, its ‘classical’ limit can still be defined by taking the limit q→1q\rightarrow 1. In that limit, the Galois Field 𝔽q\mathbb{F}_{q} can be expected to become 𝔽1\mathbb{F}_{1}, aka 𝔽un\mathbb{F}_{\mathrm{un}},666Denoting 𝔽1\mathbb{F}_{1} as 𝔽un\mathbb{F}_{\mathrm{un}} is a French-English bilingual pun. See, e.g. Refs. \refciteLeBruyn:2009,Connes:2008a. the ‘field with one element,’ an exotic and somewhat nebulous mathematical concept first suggested by Tits in 1957.[58] 𝔽q\mathbb{F}_{q} in turn can be considered the ‘quantum’-deformation of 𝔽1\mathbb{F}_{1}.[59] Though dormant for many decades, the study of 𝔽1\mathbb{F}_{1}, and efforts to actually define what it is, has intensified since the 1990’s777See, for instance, Refs. \refciteManin:1995,Manin:2008,Zhu:2000,Soule:2003,Soule:2009,Kurokawa:2003,Kurokawa:2005,Koyama:2012,Deitmar:2005,Deitmar:2006a,Deitmar:2006b,Deitmar:2008,Borger:2009,Lorscheid:2009,LopezPena:2009,LeBruyn:2009,Connes:2008a,Connes:2008b,Connes:2009a,Connes:2009b,Connes:2010a,Connes:2010b,Lescot:2009,Lescot:2011,Lescot:2012. under the expectation that it would lead to a proof of the Riemann hypothesis.888According to Kurokawa and Koyama in Ref. \refciteKurokawa:2010, the proof for the analogue of the Riemann hypothesis on projective algebraic varieties over 𝔽q\mathbb{F}_{q} is known (one of the Weil conjectures proved by Deligne in 1974). The hope is that reinterpreting the integers ℤ\mathbb{Z} as a projective algebraic variety over 𝔽1\mathbb{F}_{1} will lead to a proof of the Riemann hypothesis utilizing similar techniques. In a separate popular book[86], Kurokawa and Koyama state that Mochizuki’s recent work on the ABC conjecture[87] also uses ideas based on 𝔽1\mathbb{F}_{1}.999See also Ref. \refciteBerryKeating:1998 for a quantum mechanical approach to the Riemann hypothesis.

This paper is organized as follows. In section 2 we review how GQM is constructed on 𝔽qN\mathbb{F}_{q}^{N}. We will look at what happens to the theory if we let q=1q=1 for the case N=2N=2, and discuss its ‘classical’ properties. In section 3 we review the notion of 𝔽1\mathbb{F}_{1}, the ‘field with one element,’ following the treatment of Kurokawa and Koyama,[85] and show that the q=1q=1 limit of GQM can be constructed directly on the ‘vector’ space 𝔽1N\mathbb{F}_{1}^{N}. The resulting state space is a projective geometry, just as with GQM and canonical QM. The theory on 𝔽1N\mathbb{F}_{1}^{N} also prohibits the superposition of states, precisely the property one expects in a ‘classical’ theory. Discussion on what this teaches us is given in section 4.

2 Galois Field Quantum Mechanics

2.1 The Model

Since the details of GQM can be found in Refs. \refciteChang:2012eh and \refciteChang:2012gg, here we only present the basic outline.

Consider the vector space 𝔽qN\mathbb{F}_{q}^{N}. Vectors in 𝔽qN\mathbb{F}_{q}^{N} represent states of the model system, while dual vectors in the dual vector space 𝔽qN⁣∗\mathbb{F}_{q}^{N*} represent possible outcomes of measurements. The probability of obtaining the outcome represented by the dual-vector ⟨x|∈𝔽qN⁣∗\langle x|\in\mathbb{F}_{q}^{N*} when a measurement is performed on the state represented by the vector |ψ⟩∈𝔽qN|\psi\rangle\in\mathbb{F}_{q}^{N} is given by

P​(x|ψ)=|⟨x|ψ⟩|2∑y|⟨y|ψ⟩|2,P(x|\psi)\;=\;\dfrac{\bigl{|}\langle x|\psi\rangle\bigr{|}^{2}}{\sum_{y}\bigl{|}\langle y|\psi\rangle\bigr{|}^{2}}\;, (2)

where the sum in the denominator runs over all the dual vectors ⟨y|\langle y| in a basis of 𝔽qN⁣∗\mathbb{F}_{q}^{N*} which includes ⟨x|\langle x|. The choice of basis of 𝔽qN⁣∗\mathbb{F}_{q}^{N*} corresponds to an observable, each dual vector in the basis representing a different outcome.

The absolute value function in the above expression converts elements of 𝔽q\mathbb{F}_{q} into either zero or one in ℝ\mathbb{R}:

|k¯|={0if k¯=0¯,1if k¯≠0¯.|\,\underline{k}\,|\;=\;\left\{\begin{array}[]{ll}0&\mbox{if $\underline{k}=\underline{0}$}\;,\\ 1&\mbox{if $\underline{k}\neq\underline{0}$}\;.\end{array}\right. (3)

Here, underlined numbers and symbols represent elements of 𝔽q\mathbb{F}_{q}, to distinguish them from elements of ℝ\mathbb{R} or ℂ\mathbb{C}. Since 𝔽q\{0¯}\mathbb{F}_{q}\backslash\{\underline{0}\} is a cyclic multiplicative group, this assignment of ‘absolute values’ is the only one consistent with the requirement that the map from 𝔽q\mathbb{F}_{q} to non-negative ℝ\mathbb{R} be product preserving, that is: |k¯​l¯|=|k¯|​|l¯||\underline{k}\underline{l}|=|\underline{k}||\underline{l}|.101010The product preserving nature of the absolute value function guarantees that the probabilities of product observables on product states factorize in multi-particle systems. This property is crucial if we want to have isolated particle states. Since the same absolute value is assigned to all non-zero brackets, all outcomes ⟨x|\langle x| for which the bracket with the state |ψ⟩|\psi\rangle is non-zero are given equal probabilistic weight.

Note also that the multiplication of |ψ⟩|\psi\rangle with a non-zero element of 𝔽q\mathbb{F}_{q} will not affect the probability. Thus, vectors that differ by non-zero multiplicative constants are identified as representing the same physical state, and the state space is endowed with the finite projective geometry [89, 90, 91, 92]

P​G​(N−1,q)=(𝔽qN\{𝟎¯})/(𝔽q\{0¯}),PG(N-1,q)\;=\;(\,\mathbb{F}_{q}^{N}\backslash\{\mathbf{\underline{0}}\}\,)\,\big{/}\,(\,\mathbb{F}_{q}\backslash\{\underline{0}\}\,)\;, (4)

where each ‘line’ going through the origin of 𝔽qN\mathbb{F}_{q}^{N} is identified as a ‘point,’ in close analogy to the complex projective geometry ℂ​PN−1\mathbb{C}P^{N-1} of canonical NN-level QM defined on ℂN\mathbb{C}^{N}.

2.2 An Example

To give a concrete example of our proposal, let us construct a 2-level system, analogous to spin, on 𝔽q2\mathbb{F}_{q}^{2} for which the state space is P​G​(1,q)PG(1,q). This geometry consists of q+1q+1 ‘points,’ which can be represented by the vectors

| 0⟩=[1¯0¯],| 1⟩=[0¯1¯],|r⟩=[a¯r−11¯],|\,0\,\rangle=\left[\begin{array}[]{c}\underline{1}\\ \underline{0}\end{array}\right],\quad|\,1\,\rangle=\left[\begin{array}[]{c}\underline{0}\\ \underline{1}\end{array}\right],\quad|\,r\,\rangle=\left[\begin{array}[]{l}\underline{a}^{r-1}\\ \underline{1}\end{array}\right], (5)

r=2,3,⋯,qr=2,3,\cdots,q, where a¯\underline{a} is the generator of the multiplicative group 𝔽q\{0¯}\mathbb{F}_{q}\backslash\{\underline{0}\} with a¯q−1=1\underline{a}^{q-1}=1. The number q+1q+1 results from the fact that of the q2−1q^{2}-1 non-zero vectors in 𝔽q2\mathbb{F}_{q}^{2}, every q−1q-1 are equivalent, thus the number of inequivalent vectors is (q2−1)/(q−1)=(q+1)(q^{2}-1)/(q-1)=(q+1). Similarly, the q+1q+1 inequivalent dual-vectors can be represented as:

⟨0¯|\displaystyle\langle\,\overline{0}\,| =\displaystyle= [0¯−1¯],\displaystyle\bigl{[}\,\underline{0}\;-\!\underline{1}\,\bigr{]}\;, (6)
⟨1¯|\displaystyle\langle\,\overline{1}\,| =\displaystyle= [1¯​0¯],\displaystyle\bigl{[}\,\underline{1}\;\;\phantom{-}\underline{0}\,\bigr{]}\;, (7)
⟨r¯|\displaystyle\langle\,\overline{r}\,| =\displaystyle= [1¯−a¯r−1],r=2,3,⋯,q,\displaystyle\bigl{[}\,\underline{1}\;-\!\underline{a}^{r-1}\,\bigr{]}\;,\qquad r=2,3,\cdots,q\;, (8)

where the minus signs are dropped when the characteristic of 𝔽q\mathbb{F}_{q} is two.111111The ‘characteristic’ of a field is the smallest non-negative integer mm such that 1¯+1¯+⋯+1¯⏟m​times=0¯\underbrace{\underline{1}+\underline{1}+\cdots+\underline{1}}_{m\;\mathrm{times}}=\underline{0}, where 1¯\underline{1} is the multiplicative unit and 0¯\underline{0} is the additive unit. For example, the characteristic of 𝔽q\mathbb{F}_{q} with q=pnq=p^{n} is the prime pp. The characteristics of ℚ\mathbb{Q}, ℝ\mathbb{R}, and ℂ\mathbb{C} are defined to be zero. From these definitions, we find:

⟨r¯|s⟩\displaystyle\langle\bar{r}|s\rangle =\displaystyle= 0¯if r=s,\displaystyle\underline{0}\quad\mbox{if $r=s$}\;, (9)
≠\displaystyle\neq 0¯if r≠s,\displaystyle\underline{0}\quad\mbox{if $r\neq s$}\;, (10)

and

|⟨r¯|s⟩|={0if r=s,1if r≠s.\bigl{|}\langle\bar{r}|s\rangle\bigr{|}\;=\;\begin{cases}0&\mbox{if $r=s$}\;,\\ 1&\mbox{if $r\neq s$}\;.\end{cases} (11)

Observables are associated with a choice of basis of 𝔽q2⁣∗\mathbb{F}_{q}^{2*}:

Ar​s≡{⟨r¯|,⟨s¯|},r≠s.A_{rs}\;\equiv\;\{\;\langle\bar{r}|,\;\langle\bar{s}|\;\}\;,\qquad r\neq s\;. (12)

We assign the outcome +1+1 to the first dual-vector of the pair, and the outcome −1-1 to the second to make these observables spin-like. This assignment implies As​r=−Ar​sA_{sr}=-A_{rs}. The indices r​srs can be considered as indicating the direction of the ‘spin,’ and the interchange of the indices as indicating a reversal of this direction.

Applying Eq. (2) to this system, it is straightforward to show that

P​(Ar​s=+1|r)\displaystyle P(A_{rs}=+1\,|\,r) =\displaystyle= 0,P​(Ar​s=−1|r)= 1,\displaystyle 0\;,\qquad P(A_{rs}=-1\,|\,r)\;=\;1\;, (13)
P​(Ar​s=+1|s)\displaystyle P(A_{rs}=+1\,|\,s) =\displaystyle= 1,P​(Ar​s=−1|s)= 0,\displaystyle 1\;,\qquad P(A_{rs}=-1\,|\,s)\;=\;0\;, (14)
P​(Ar​s=±1|t)\displaystyle P(A_{rs}=\pm 1\,|\,t) =\displaystyle= 12,for t≠r,s,\displaystyle\frac{1}{2}\;,\qquad\mbox{for $t\neq r,s$}\;, (15)

and thus,

⟨Ar​s⟩r\displaystyle\langle A_{rs}\rangle_{r} =\displaystyle= −1,\displaystyle-1\;, (16)
⟨Ar​s⟩s\displaystyle\langle A_{rs}\rangle_{s} =\displaystyle= +1,\displaystyle+1\;, (17)
⟨Ar​s⟩t\displaystyle\langle A_{rs}\rangle_{t} =\displaystyle= 0,for t≠r,s.\displaystyle\phantom{-}0\;,\quad\mbox{for $t\neq r,s$.} (18)

So for each ‘spin’ Ar​sA_{rs} there exist two ‘eigenstates’: |s⟩|s\rangle for +1+1 (‘spin’ up) and |r⟩|r\rangle for −1-1 (‘spin’ down). For all other states the two outcomes ±1\pm 1 are equally probable.

2.3 Spin Correlations

A two-‘spin’ system can be constructed on the tensor product space 𝔽q2⊗𝔽q2=𝔽q4\mathbb{F}_{q}^{2}\otimes\mathbb{F}_{q}^{2}=\mathbb{F}_{q}^{4}. The number of non-zero vectors in this space is q4−1q^{4}-1, of which every q−1q-1 are equivalent, so the number of inequivalent states is (q4−1)/(q−1)=q3+q2+q+1(q^{4}-1)/(q-1)=q^{3}+q^{2}+q+1. Of these, (q+1)2(q+1)^{2} are product states, leaving (q3+q2+q+1)−(q+1)2=q​(q2−1)(q^{3}+q^{2}+q+1)-(q+1)^{2}=q(q^{2}-1) that are entangled. Of these, there exists an analog of the spin-singlet state given by

|S⟩=|r⟩⊗|s⟩−|s⟩⊗|r⟩,r≠s,|S\rangle\;=\;|r\rangle\otimes|s\rangle-|s\rangle\otimes|r\rangle\;,\qquad r\neq s\;, (19)

for any two states |r⟩|r\rangle and |s⟩|s\rangle up to a multiplicative constant. If the characteristic of 𝔽q\mathbb{F}_{q} is two, the minus sign is replaced by a plus sign.

Products of the ‘spin’ observables are defined as

Ar​s​At​u={⟨r¯|⊗⟨t¯|,⟨r¯|⊗⟨u¯|,⟨s¯|⊗⟨t¯|,⟨s¯|⊗⟨u¯|},A_{rs}A_{tu}\,=\,\{\,\langle\bar{r}|\otimes\langle\bar{t}|\,,\,\langle\bar{r}|\otimes\langle\bar{u}|\,,\,\langle\bar{s}|\otimes\langle\bar{t}|\,,\,\langle\bar{s}|\otimes\langle\bar{u}|\,\}\;, (20)

the four tensor products representing the outcomes +⁣+++, +⁣−+-, −⁣+-+, and −⁣−--, and the expectation value giving the correlation between the two ‘spins.’ The probabilities of the four outcomes are particularly easy to calculate for the state |S⟩|S\rangle since [42]

(⟨r¯|⊗⟨s¯|)​|S⟩\displaystyle\bigl{(}\langle\bar{r}|\otimes\langle\bar{s}|\,\bigr{)}|S\rangle =\displaystyle= 0¯if r=s,\displaystyle\underline{0}\quad\mbox{if $r=s$}\;, (21)
≠\displaystyle\neq 0¯if r≠s,\displaystyle\underline{0}\quad\mbox{if $r\neq s$}\;, (22)

thus

|(⟨r¯|⊗⟨s¯|)​|S⟩|= 1−δr​s,\Bigl{|}\bigl{(}\langle\bar{r}|\otimes\langle\bar{s}|\,\bigr{)}|S\rangle\Bigr{|}\;=\;1-\delta_{rs}\;, (23)

and we obtain the probabilities and correlations listed in Table. 1. It is straightforward to show that these probabilities cannot be reproduced by any classical hidden variable theory.[40, 41] Thus, GQM is ‘quantum’ in this sense.

 Observable  +⁣+++  +⁣−+-  −⁣+-+  −⁣−--  E.V.
Ar​s​Ar​sA_{rs}A_{rs} 0 12\dfrac{1}{2} 12\dfrac{1}{2} 0 −1-1 —
Ar​s​Ar​tA_{rs}A_{rt} 0 13\dfrac{1}{3} 13\dfrac{1}{3} 13\dfrac{1}{3} −13-\dfrac{1}{3} —
Ar​s​As​tA_{rs}A_{st} 13\dfrac{1}{3} 13\dfrac{1}{3} 0 13\dfrac{1}{3} +13+\dfrac{1}{3} —
Ar​s​At​uA_{rs}A_{tu} 14\dfrac{1}{4} 14\dfrac{1}{4} 14\dfrac{1}{4} 14\dfrac{1}{4} 0\phantom{-}0 —
Table 1: Probabilities and expectation values of product observables in the state |S⟩|S\rangle. The indices rr, ss, tt, and uu are distinct. Cases that can be obtained by flipping signs using Ar​s=−As​rA_{rs}=-A_{sr} are not shown.

It should be noted, though, that GQM also has a common feature with CM when we look at the Clauser-Horne-Shimony-Holt (CHSH) version of Bell’s inequality.[93, 94] The CHSH bound[95] is the upper bound of the absolute value of the following combination of correlators:

⟨A,a;B,b⟩≡⟨A​B⟩+⟨A​b⟩+⟨a​B⟩−⟨a​b⟩,\langle A,a\,;B,b\rangle\;\equiv\;\langle AB\rangle+\langle Ab\rangle+\langle aB\rangle-\langle ab\rangle\;, (24)

where AA and aa are two observables of particle 1, and BB and bb are two observables of particle 2. All four observables are assumed to take on only the values ±1\pm 1 upon measurement. For classical hidden variable theory the bound on |⟨A,a;B,b⟩|\left|\langle A,a;B,b\rangle\right| is 2,[95] while for canonical QM it is 2​22\sqrt{2}.[96, 97] 121212The largest possible value of the CHSH bound is 4.[98] See Ref. \refciteChang:2012we for a model which saturates this bound. Despite GQM not allowing a hidden variable minic, we can nevertheless show, again using Table 1, that its CHSH bound is the ‘classical’ value of 2. This bound is independent of the value of qq chosen for our Galois field 𝔽q\mathbb{F}_{q} and is thus also independent of the size of the vector space 𝔽q2⊗𝔽q2=𝔽q4\mathbb{F}_{q}^{2}\otimes\mathbb{F}_{q}^{2}=\mathbb{F}_{q}^{4}.[40, 41] So the limitation on the correlations is not due to the limited number of ‘spin’ direction available in the model. This GQM example shows that the absence of hidden variable mimics does not necessarily guarantee the violation of the classical CHSH bound.

2.4 The q=1q=1 Limit

The Galois fields 𝔽q\mathbb{F}_{q} are only defined for q=pnq=p^{n}, that is, the order qq must be a power of a prime pp. Thus, setting q=1q=1 is illegitimate from the Galois field point of view. Indeed, 𝔽q\mathbb{F}_{q} consists of qq elements so taking the naive q=1q=1 limit, one expects 𝔽1\mathbb{F}_{1} to consist of only one element (which could be denoted by either 0¯\underline{0} or 1¯\underline{1}) with no distinction between addition and multiplication. Such an object is obviously not a “field.” In other words, a “field with one element,” in the usual sense of the term, does not and cannot exist. However, a different and more interesting picture emerges if instead of trying to define 𝔽1\mathbb{F}_{1} first, we set q=1q=1 directly in the N=2N=2 ‘spin’ model we constructed above.

First, the number of states given in Eq. (5) will be reduced to q+1=2q+1=2,

|↑⟩≡|0⟩=[1¯0¯],|↓⟩≡|1⟩=[0¯1¯],|\uparrow\;\rangle\;\equiv\;|0\rangle\;=\;\left[\begin{array}[]{c}\underline{1}\\ \underline{0}\end{array}\right]\;,\qquad|\downarrow\;\rangle\;\equiv\;|1\rangle\;=\;\left[\begin{array}[]{c}\underline{0}\\ \underline{1}\end{array}\right]\;, (25)

as are the possible outcomes listed in Eq. (8):

⟨↓|≡⟨0¯|=[0¯1¯],⟨↑|≡⟨1¯|=[1¯0¯].\langle\;\downarrow|\;\equiv\;\langle\overline{0}|\;=\;\left[\,\underline{0}\;\;\underline{1}\,\right]\;,\qquad\langle\;\uparrow|\;\equiv\;\langle\overline{1}|\;=\;\left[\,\underline{1}\;\;\underline{0}\,\right]\;. (26)

The sole observable of the system, Eq. (12), will be

A≡A10=−A01={⟨1¯|,⟨0¯|}={⟨↑|,⟨↓|},A\;\equiv\;A_{10}\;=\;-A_{01}\;=\;\{\;\langle\overline{1}|\,,\;\langle\overline{0}|\;\}\;=\;\{\;\langle\;\uparrow|\,,\;\langle\;\downarrow|\;\}\;, (27)

for which |↑⟩|\uparrow\;\rangle and |↓⟩|\downarrow\;\rangle are ‘eigenstates’:

⟨A⟩↑= 1,⟨A⟩↓=−1.\langle A\rangle_{\uparrow}\;=\;1\;,\qquad\langle A\rangle_{\downarrow}\;=\;-1\;. (28)

Thus, a measurement of AA on |↑⟩|\uparrow\;\rangle will always yield +1+1, while that on |↓⟩|\downarrow\;\rangle will always yield −1-1. No superpositions of these states exist, so the system reduces to a ‘classical’ one where each states has a definite outcome upon measurement.

The ‘two’-spin system will reduce to q3+q2+q+1=4q^{3}+q^{2}+q+1=4 states, of which q​(q2−1)=0q(q^{2}-1)=0 are entangled. These four are the product states

|↑↑⟩≡|↑⟩⊗|↑⟩=[1¯0¯0¯0¯],|↑↓⟩≡|↑⟩⊗|↓⟩=[0¯1¯0¯0¯],|↓↑⟩≡|↓⟩⊗|↑⟩=[0¯0¯1¯0¯],|↓↓⟩≡|↓⟩⊗|↓⟩=[0¯0¯0¯1¯].\begin{array}[]{lll}|\uparrow\uparrow\;\rangle&\equiv\;|\uparrow\;\rangle\otimes|\uparrow\;\rangle\;=\;\left[\begin{array}[]{c}\underline{1}\\ \underline{0}\\ \underline{0}\\ \underline{0}\end{array}\right]\;,\quad|\uparrow\downarrow\;\rangle&\equiv\;|\uparrow\;\rangle\otimes|\downarrow\;\rangle\;=\;\left[\begin{array}[]{c}\underline{0}\\ \underline{1}\\ \underline{0}\\ \underline{0}\end{array}\right]\;,\vphantom{\Big{|}}\\ |\downarrow\uparrow\;\rangle&\equiv\;|\downarrow\;\rangle\otimes|\uparrow\;\rangle\;=\;\left[\begin{array}[]{c}\underline{0}\\ \underline{0}\\ \underline{1}\\ \underline{0}\end{array}\right]\;,\quad|\downarrow\downarrow\;\rangle&\equiv\;|\downarrow\;\rangle\otimes|\downarrow\;\rangle\;=\;\left[\begin{array}[]{c}\underline{0}\\ \underline{0}\\ \underline{0}\\ \underline{1}\end{array}\right]\;.\vphantom{\Big{|}}\end{array} (29)

The sole product observable, Eq. (20), is

AA={⟨↑↑|,⟨↑↓|,⟨↓↑|,⟨↓↓|},AA\,=\,\{\,\langle\;\uparrow\uparrow|\,,\,\langle\;\uparrow\downarrow|\,,\,\langle\;\downarrow\uparrow|\,,\,\langle\;\downarrow\downarrow|\,\}\;, (30)

where

⟨↑↑|\displaystyle\langle\;\uparrow\uparrow| =\displaystyle= ⟨↑|⊗⟨↑|=[1¯0¯0¯0¯],\displaystyle\langle\;\uparrow|\otimes\langle\;\uparrow|\;=\;\left[\,\underline{1}\;\;\underline{0}\;\;\underline{0}\;\;\underline{0}\,\right]\,, (31)
⟨↑↓|\displaystyle\langle\;\uparrow\downarrow| =\displaystyle= ⟨↑|⊗⟨↓|=[0¯1¯0¯0¯],\displaystyle\langle\;\uparrow|\otimes\langle\;\downarrow|\;=\;\left[\,\underline{0}\;\;\underline{1}\;\;\underline{0}\;\;\underline{0}\,\right]\,, (32)
⟨↓↑|\displaystyle\langle\;\downarrow\uparrow| =\displaystyle= ⟨↓|⊗⟨↑|=[0¯0¯1¯0¯],\displaystyle\langle\;\downarrow|\otimes\langle\;\uparrow|\;=\;\left[\,\underline{0}\;\;\underline{0}\;\;\underline{1}\;\;\underline{0}\,\right]\,, (33)
⟨↓↓|\displaystyle\langle\;\downarrow\downarrow| =\displaystyle= ⟨↓|⊗⟨↓|=[0¯0¯0¯1¯].\displaystyle\langle\;\downarrow|\otimes\langle\;\downarrow|\;=\;\left[\,\underline{0}\;\;\underline{0}\;\;\underline{0}\;\;\underline{1}\,\right]\,. (34)

The four states are all ‘eigenstates’ of A​AAA with definite outcomes when A​AAA is measured. Thus, the system is ‘classical’ and its CHSH bound is trivially 2.

This discussion can be easily generalized to generic NN-level GQM, in which the q=1q=1 case reduces to a situation where all surviving states are simultaneous ‘eigenstates’ of all surviving observables.131313Note that the vector space 𝔽qN\mathbb{F}_{q}^{N} has qN−1q^{N}-1 non-zero vectors, of which every q−1q-1 of them are physically equivalent in GQM. Thus, the number of physically distinct states is (qN−1)/(q−1)=[N]q(q^{N}-1)/(q-1)=[N]_{q}, which reduces to NN in the q=1q=1 limit. What this exercise shows is that the q=1q=1 limit of GQM, though admittedly fairly trivial, does make sense as a ‘classical’ theory. Note, however, that in letting q=1q=1 we have retained the formalism of GQM as is from the ‘quantum’ q≠1q\neq 1 cases. The only thing that has changed is the value of qq, which changes the number field over which the state space is constructed, formally, from 𝔽q\mathbb{F}_{q} to 𝔽1\mathbb{F}_{1}. But what is 𝔽1\mathbb{F}_{1}?

3 The Field with One Element 𝔽𝟏\bm{\mathbb{F}_{1}}

3.1 The Observation of Tits

The concept of 𝔽1\mathbb{F}_{1} first appeared in a 1957 paper by Tits[58] as the “corps de caractéristique 1.” The main objective of the Tits paper was to use projective geometries to define the then recently discovered Chevalley groups geometrically.141414Chevalley originally used purely algebraic methods in his definition.[99] To each Dynkin diagram and choice of finite field 𝔽q\mathbb{F}_{q} a projective geometry was associated, and the Chevalley group identified with the group of projective transformations on the geometry. For instance, the projective geometry P​G​(N−1,q)PG(N-1,q) is associated with the Dynkin diagram for AN−1A_{N-1}, and the corresponding Chevalley group is the projective linear group P​G​L​(N,q)PGL(N,q). Towards the end of the paper, Tits argues that the q=1q=1 limit of P​G​(N−1,q)PG(N-1,q) actually makes sense as a projective geometry,151515With a minor modification to the axioms. See Ref. \refciteCohn:2004. and that the corresponding group of projective transformations is SNS_{N}.

To see this, let us first list a few properties of P​G​(N−1,q)PG(N-1,q) for q≠1q\neq 1:

  • •

    The number of points in the geometry P​G​(N−1,q)PG(N-1,q) is

    [N1]q=[N]q=qN−1q−1= 1+q+q2+⋯+qN−1.\left[\begin{array}[]{c}N\\ 1\end{array}\right]_{q}\;=\;[N]_{q}\;=\;\dfrac{q^{N}-1}{q-1}\;=\;1+q+q^{2}+\cdots+q^{N-1}\;. (35)

    (See chapter 9 of Ref. \refciteCameron:1994.) Here, the notation is:

    [NM]q=[N]q![M]q!​[N−M]q!,\left[\begin{array}[]{c}N\\ M\end{array}\right]_{q}\;=\;\dfrac{[N]_{q}!}{[M]_{q}![N-M]_{q}!}\;, (36)

    with

    [N]q\displaystyle[N]_{q} ≡\displaystyle\equiv 1+q+q2+⋯+qN−1=qN−1q−1,\displaystyle 1+q+q^{2}+\cdots+q^{N-1}\;=\;\dfrac{q^{N}-1}{q-1}\;, (37)
    [N]q!\displaystyle[N]_{q}! ≡\displaystyle\equiv [N]q​[N−1]q​⋯​[2]q​[1]q.\displaystyle[N]_{q}[N-1]_{q}\cdots[2]_{q}[1]_{q}\;. (38)

    Eq. (36) is known as the Gaussian binomial coefficient, [N]q[N]_{q} the qq-analogue of the natural number NN, and [N]q![N]_{q}! the qq-factorial.


  • •

    The number of kk-dimensional subspaces (points, lines, planes, etc.) in P​G​(N−1,q)PG(N-1,q) are

    [Nk+1]q\displaystyle\left[\begin{array}[]{c}N\\ k+1\end{array}\right]_{q} =\displaystyle= [N]q​[N−1]q​⋯​[N−k]q[k+1]q​[k]q​⋯​[1]q\displaystyle\dfrac{[N]_{q}[N-1]_{q}\cdots[N-k]_{q}}{[k+1]_{q}[k]_{q}\cdots[1]_{q}} (41)
    =\displaystyle= (qN−1)​(qN−1−1)​⋯​(qN−k−1)(qk+1−1)​(qk−1)​⋯​(q−1).\displaystyle\dfrac{(q^{N}-1)(q^{N-1}-1)\cdots(q^{N-k}-1)}{(q^{k+1}-1)(q^{k}-1)\cdots(q-1)}\;. (42)

    Each subspace contains

    [k+1]q=qk+1−1q−1= 1+q+q2+⋯+qk[k+1]_{q}\;=\;\dfrac{q^{k+1}-1}{q-1}\;=\;1+q+q^{2}+\cdots+q^{k} (43)

    points.


In the limit q→1q\rightarrow 1, the qq-analog [N]q[N]_{q} reduces to NN, the qq-factorial to the usual factorial, and the Gaussian binomial coefficients to the usual binomial coefficients. So the ‘projective’ space P​G​(N−1,1)PG(N-1,1) should be a space consisting of NN ‘points,’ and the number of kk-dimensional subspaces should be Ck+1N{}_{N}C_{k+1}, each consisting of k+1k+1 points. This would simply be a set consisting of NN elements, and all the subsets consisting of k+1k+1 elements are the kk-dimensional subspaces. The group of all possible transformations, projective or not, for a set with NN elements is obviously SNS_{N}.

For instance, say N=3N=3: P​G​(2,1)PG(2,1) would consist of three ‘points.’ Let us call them aa, bb, and cc. The ‘lines’ in this geometry would be {a,b}\{a,b\}, {b,c}\{b,c\}, and {c,a}\{c,a\}, and the ‘plane’ would be the entire space {a,b,c}\{a,b,c\}. There is a nice discussion in Ref. \refciteCohn:2004 on how this definition of ‘geometry’ satisfies all the required axioms. All possible ways to map the three points onto themselves form the group S3S_{3}.

Now, recalling that P​G​(N−1,q)PG(N-1,q) was obtained by identifying the lines through the origin of 𝔽qN\mathbb{F}_{q}^{N} as points, Tits argues that the ‘projective geometry’ P​G​(N−1,1)PG(N-1,1) should be interpretable as

P​G​(N−1,1)=(𝔽1N\{𝟎¯})/(𝔽1\{0¯}),PG(N-1,1)\;=\;\bigl{(}\,\mathbb{F}_{1}^{N}\backslash\{\mathbf{\underline{0}}\}\,\bigr{)}\Big{/}\bigl{(}\,\mathbb{F}_{1}\backslash\{\underline{0}\}\,\bigr{)}\;, (44)

where 𝔽1\mathbb{F}_{1} is the “corps de caractéristique 1,” aka the field with one element.

3.2 𝔽1\mathbb{F}_{1} according to Kurokawa and Koyama

Taken literally, “corps de caractéristique 1” implies161616See footnote 11.

1ÂŻ=0ÂŻ,\underline{1}\;=\;\underline{0}\;, (45)

that is, the multiplicative unit and the additive unit are the same number, so it may seem that 𝔽1={1¯}={0¯}\mathbb{F}_{1}=\{\underline{1}\}=\{\underline{0}\}, where

1¯×1¯=1¯,1¯+1¯=1¯.\underline{1}\times\underline{1}=\underline{1}\;,\qquad\underline{1}+\underline{1}\;=\;\underline{1}\;. (46)

Such an object would not have any structure to speak of and it is difficult to envision how one could construct any geometry, projective or otherwise, on it. Thus, a more sophisticated definition of 𝔽1\mathbb{F}_{1} is called for.

Various definitions of 𝔽1\mathbb{F}_{1} and 𝔽1\mathbb{F}_{1}-geometries exist in the literature, which endeavor to make sense of these concepts, a survey of which can be found in Ref. \refciteLopezPena:2009 by Lorscheid and Lopez-Pena. Here, we adopt the definition of 𝔽1\mathbb{F}_{1} by Kurokawa and Koyama in Ref. \refciteKurokawa:2010 who argue that 𝔽1\mathbb{F}_{1} should be interpreted as the set 𝔽1={1¯}+{0¯}\mathbb{F}_{1}=\{\underline{1}\}+\{\underline{0}\} on which only multiplication is defined. Note that 1¯\underline{1} is the multiplicative unit, and 0¯\underline{0} is the multiplicative zero, that is:

1¯​k¯=k¯​1¯=k¯,0¯​k¯=k¯​0¯=0¯,k¯=0¯ or 1¯.\underline{1}\,\underline{k}\,=\,\underline{k}\,\underline{1}\,=\,\underline{k}\,,\qquad\underline{0}\,\underline{k}\,=\,\underline{k}\,\underline{0}\,=\,\underline{0}\,,\qquad\underline{k}\,=\,\mbox{$\underline{0}$ or $\underline{1}$}\;. (47)

This is different from 𝔽2={0¯,1¯}\mathbb{F}_{2}=\{\underline{0},\underline{1}\} in that addition is not allowed.171717It may be more precise to say that only the addition of 0¯\underline{0} is allowed. 1¯+1¯\underline{1}+\underline{1} is prohibited. Another definition referred to in the literature[62, 82, 83, 84] defines 1¯+1¯=1¯\underline{1}+\underline{1}=\underline{1} with all other additions and multiplications among 0¯\underline{0} and 1¯\underline{1} defined in the usual way, cf. Eq. (46). This is the reason for the unfamiliar notation {1¯}+{0¯}\{\underline{1}\}+\{\underline{0}\}: {1¯}\{\underline{1}\} is the multiplicative group by itself, and {0¯}\{\underline{0}\} is the multiplicative zero (but not the additive unit since there is no addition) which is tacked on. Thus, 𝔽1\mathbb{F}_{1} is not really a field but what is known as a ‘monoid.’

Since addition is not allowed, ‘vector’ spaces over 𝔽1N\mathbb{F}_{1}^{N} are not truly ‘vector’ spaces in the usual sense of the term. However, lacking a better alternative, let us use vector space terminology anyway. Since 𝔽1N\mathbb{F}_{1}^{N} is an NN-dimensional space, it will have NN basis vectors, which we denote

|1⟩,|2⟩,⋯,|N⟩.|1\rangle,\;|2\rangle,\;\cdots,\;|N\rangle\;. (48)

The linear combinations of these basis vectors are of the form

∑i=1Nci​|i⟩,ci∈𝔽1,\sum_{i=1}^{N}c_{i}|i\rangle\;,\qquad c_{i}\in\mathbb{F}_{1}\;, (49)

but since we should have no addition, only one of the cic_{i}’s is allowed to be non-zero at a time. So the NN basis vectors are the only (non-zero) vectors in this space. They will each be represented by column vectors with only one 1¯\underline{1}, the remaining elements being all 0¯\underline{0}. Note that these are precisely the states we obtained in taking the limit q=1q=1 in GQM.

Linear transformations on 𝔽1N\mathbb{F}_{1}^{N} must not lead to addition of the basis vectors. So the only allowed transformations are those for which the N×NN\times N matrix representation has at most one 1¯\underline{1} in each row. The non-singular ones, i.e. the automorphisms, are the ones in which there is one 1¯\underline{1} in each row and each column. These are simply the matrices that permute the NN basis vectors. Since 𝔽1\{0¯}={1¯}\mathbb{F}_{1}\backslash\{\underline{0}\}=\{\underline{1}\}, the projective space P​G​(N−1,1)PG(N-1,1) will simply consist of the non-zero vectors in 𝔽1N\mathbb{F}_{1}^{N}. There are NN of these, as required for P​G​(N−1,1)PG(N-1,1). The projective linear group P​G​L​(N,1)PGL(N,1) would be SNS_{N}.

Linear maps from 𝔽1N\mathbb{F}_{1}^{N} to 𝔽1\mathbb{F}_{1}, i.e. the dual vectors, must also satisfy the no-addition requirement, so they would be N×1N\times 1 row-vectors, each with only one 1¯\underline{1}, the remaining elements all 0¯\underline{0}. These are precisely the dual vectors representing outcomes in the q=1q=1 limit of GQM.

Thus, the Kurokawa-Koyama approach matches well with the GQM construction, and the q=1q=1 ‘classical’ limit can be constructed directly on 𝔽1N\mathbb{F}_{1}^{N} defined in this fashion. The resulting state space has the projective geometry P​G​(N−1,1)PG(N-1,1). Note, particularly, that the prohibition of addition in Kurokawa-Koyama leads to the lack of superposition of states in the ‘classical’ limit of GQM. Therefore, the q=1q=1 limit of GQM can be considered to be ‘classical’ due to this very special feature of 𝔽1\mathbb{F}_{1}.

4 Summary and Discussion

Illuminating the precise relation between CM and QM remains one of the outstanding problems in the foundations of physics. In this note we point out that a simple discrete toy model sheds new light on this important problem.

In order to compare the two seemingly disparate formalisms of physics, first, one needs a unifying mathematical language for such a comparison. This idea, that classical theory should be rewritten using the mathematics of QM goes back at least to Koopman[16] and von Neumann.[17, 18] In our proposal this unifying mathematical language is provided by the projective spaces of GQM.[40, 41]

We have demonstrated that GQM defined on 𝔽qN\mathbb{F}_{q}^{N} becomes ‘classical’ in the limit q=1q=1. The field with one element 𝔽1\mathbb{F}_{1}, over which this limit can be constructed, have all the ‘classical’ properties of consisting only of zero and one, and addition being prohibited. In other words, the ‘classical’-ness of q=1q=1 GQM can be traced to the ‘classical’-ness of 𝔽1\mathbb{F}_{1}. This way of achieving ‘classicality’ from ‘quantum’-ness is quite different from the usual methodology of taking ℏ→0\hbar\rightarrow 0, in the limit of which operators become commuting. In GQM, there is no ℏ\hbar and there are no operators.

Our result suggests that the pathway from a ‘quantum’ theory to a ‘classical’ one may not be unique. Looking at canonical QM, taking the ℏ→0\hbar\rightarrow 0 limit (in some yet unknown pathway) may not be the only way for the theory to become classical. Perhaps replacing ℂ\mathbb{C} with some other mathematical structure, perhaps a continuous analog of 𝔽1\mathbb{F}_{1}, though there is no telling such an entity exists, could lead to a theory which is ‘classical’ in some new way.

The basic premise of this paper is that the relation between CM and QM lies in the differences in the structure and geometry of the underlying number fields upon which the descriptions are predicated. While we have not discussed the limit q→1q\to 1 within the context of the Biorthogonal Quantum Mechanics we defined in Ref. \refciteChang:2012we, the premise is quite apparent there as well. The approach there was an alternative method of constructing a quantum-like theory on 𝔽qN\mathbb{F}_{q}^{N}. The CHSH bound on ‘spin’ systems defined on 𝔽q4=𝔽q2×𝔽q2\mathbb{F}_{q}^{4}=\mathbb{F}_{q}^{2}\times\mathbb{F}_{q}^{2} in this approach is qq-dependent. We were able to demonstrate that this bound takes on the classical value of 2 for q=3q=3, and the super-quantum value of 4 for q=9q=9. Once again, it is the differences in the structure of the underlying number fields 𝔽3\mathbb{F}_{3} and 𝔽9=𝔽3​[i¯]\mathbb{F}_{9}=\mathbb{F}_{3}[\underline{i}], the latter of which accommodates the analog of the imaginary unit ii, and their associated geometries, that presage classical, quantum, and super-quantum correlations.181818See also Refs. \refciteHanson:2011,Hanson:2012,Hanson:2013.

QM is our current fundamental framework of physics. As such it has passed all available observational tests. And yet our understanding of QM remains limited. In particular, one would still like to understand if QM follows from few elementary axioms, in analogy what happens in the special theory of relativity. Also, one would like to understand in more detail the precise relation between QM and its classical limit, the problem addressed in this paper. Finally, could QM be part of a larger framework, the way CM appears as a formal limit of QM (or in another context, similar to the way the special theory of relativity can be naturally generalized to the general theory of relativity)? It is very difficult to answer all three questions within the framework of canonical QM of the real world as we understand it today, so a simpler model was called for. This was the main motivation of our effort to build a toy model of QM using Galois fields. Surprisingly, all three questions can be addressed in this context. In our previous publications on this subject[40, 41] we have constructed the precise mathematical rules of the game in this toy world, and we have shed light on the nature of quantum correlations, on the nature of super-quantum correlations[51] (which generalize the usual quantum framework) as well as on the difference between classical and quantum correlations. In this paper, we complete this picture by addressing the subtle question of the ‘classical’ limit in the context of GQM. Understandably enough, our toy world is not the real world. However, for the first time, the two known theories of the real world (the classical and quantum theory) as well as the conjectured third theory (based on stronger than quantum, or superquantum, correlations) can be addressed in one coherent theoretical framework, that is simple yet illuminating. And this is where we find the most useful lessons of our work.

Acknowledgments

We would like to thank Chia Hsiung Tze and Cosmas Zachos for helpful discussions. DM was supported in part by the U.S. Department of Energy, grant DE-FG02-13ER41917, task A. TT is grateful for the hospitality of the Kavli-IPMU during his sabbatical year from fall 2012 to summer 2013 where the initial staged of this work was completed, and where TT was supported by the World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan.

References

  • [1] P. A. M. Dirac, The Principles of Quantum Mechanics, (Oxford University Press, 1st edition 1930, currently available edition 1982).
  • [2] M. V. Berry, “Chaos and the semiclassical limit of Quantum mechanics (is the moon there when somebody looks?),” in Quantum Mechanics: Scientific perspectives on divine action, editors: R. J. Russell, P. Clayton, K. Wegter-McNelly and J. Polkinghorne, CTNS-Vatican Observatory book series (2001) pp 41-54, http://www.phy.bris.ac.uk/people/berry_mv/the_papers/berry337.pdf.
  • [3] G. Wentzel, “Eine Verallgemeinerung der Quantenbedingungen f r die Zwecke der Wellenmechanik,” Zeitschrift fĂźr Physik 38, 518–529 (1926).
  • [4] H. A. Kramers, “Wellenmechanik und halbzählige Quantisierung,” Zeitschrift fĂźr Physik 39, 828–840 (1926).
  • [5] L. Brillouin, “La mĂŠcanique ondulatoire de SchrĂśdinger: une mĂŠthode gĂŠnĂŠrale de resolution par approximations successives,” Comptes Rendus de l’Academie des Sciences 183, 24–26 (1926).
  • [6] J. H. Van Vleck, “The Correspondence Principle in the Statistical Interpretation of Quantum Mechanics,” Proceedings of the National Academy of Sciences of the USA 14, 178–188 (1928).
  • [7] R. Schiller, “Quasi-Classical Theory of the Nonspinning Electron,” Phys. Rev. 125, 1100–1108 (1962).
  • [8] R. Schiller, “Quasi-Classical Transformation Theory,” Phys. Rev. 125, 1109–1115 (1962).
  • [9] R. Schiller, “Quasi-Classical Theory of the Spinning Electron,” Phys. Rev. 125, 1116–1123 (1962).
  • [10] G. Bacciagaluppi and A. Valentini, “Quantum Theory at the Crossroads: Reconsidering the 1927 Solvay Conference” (Cambridge University Press, 2013) [arXiv:quant-ph/0609184]
  • [11] D. Bohm, “A Suggested Interpretation of the Quantum Theory in Terms of “Hidden” Variables I,” Phys. Rev. 85, 166–179 (1952).
  • [12] D. Bohm, “A Suggested Interpretation of the Quantum Theory in Terms of “Hidden” Variables II,” Phys. Rev. 85, 180–193 (1952).
  • [13] N. Rosen, “The Relation Between Classical and Quantum Mechanics,” Am. J. Phys. 32 (1964) 597–600.
  • [14] N. Rosen, “Mixed States in Classical Mechanics,” Am. J. Phys. 33 (1964) 146–150.
  • [15] N. Rosen, “Quantum Particles and Classical Particles,” Foundations of Physics 16 (1986) 687–700.
  • [16] B. O. Koopman, “Hamiltonian Systems and Transformations in Hilbert Space,” Proceedings of the National Academy of Sciences of the USA, 17 (1931) 315–318.
  • [17] J. von Neumann, “Zur Operatorenmethode In Der Klassischen Mechanik,” Annals of Mathematics 33 (3) (1932) 587–642.
  • [18] J. von Neumann, “Zusätze Zur Arbeit ‘Zur Operatorenmethode…’,” Annals of Mathematics 33 (4) (1932) 789–791.
  • [19] E. C. G. Sudarshan, “Interaction between classical and quantum systems and the measurement of quantum observables,” Pramana 6 (1976) 117–126.
  • [20] E. Gozzi and D. Mauro, “Minimal coupling in Koopman-von Neumann theory,” Annals Phys.  296, 152–186 (2002) [quant-ph/0105113].
  • [21] D. Mauro, “On Koopman-von Neumann waves,” Int. J. Mod. Phys. A 17, 1301–1325 (2002) [quant-ph/0105112].
  • [22] D. Mauro, “Topics in Koopman-von Neumann Theory,” arXiv:quant-ph/0301172.
  • [23] E. Gozzi and D. Mauro, “On Koopman-von Neumann waves 2,” Int. J. Mod. Phys. A 19, 1475–1494 (2004) [quant-ph/0306029].
  • [24] P. Carta, E. Gozzi and D. Mauro, “Koopman-von Neumann formulation of classical Yang-Mills theories. I.,” Annalen Phys.  15, 177–215 (2006) [hep-th/0508244].
  • [25] W. H. Zurek, “Decoherence and the Transition from Quantum to Classical – Revisited,” in Quantum Decoherence, PoincarĂŠ Seminar 2005, editors: B. Duplantier, J.-M. Raimond, and V. Revasseau, Progress in Mathematical Physics Volume 48 (Springer, 2007) 1–31 [arXiv:quant-ph/0306072].
  • [26] W. H. Zurek, “Decoherence, einselection, and the quantum origins of the classical,” Rev. Mod. Phys.  75, 715 (2003).
  • [27] H. Everett, “Relative state formulation of quantum mechanics,” Rev. Mod. Phys.  29, 454 (1957).
  • [28] T. N. Sherry and E. C. G. Sudarshan, “Interaction Between Classical and Quantum Systems: A New Approach to Quantum Measurement. 1.,” Phys. Rev. D 18, 4580 (1978).
  • [29] T. N. Sherry and E. C. G. Sudarshan, “Interaction Between Classical And Quantum Systems: A New Approach To Quantum Measurement. Ii. Theoretical Considerations,” Phys. Rev. D 20, 857 (1979).
  • [30] S. R. Gautam, T. N. Sherry and E. C. G. Sudarshan, “Interaction Between Classical and Quantum Systems: A New Approach to Quantum Measurement. 3. Illustration,” Phys. Rev. D 20, 3081 (1979).
  • [31] A. Peres and D. R. Terno, “Hybrid classical-quantum dynamics,” Phys. Rev. A 63, 022101 (2001).
  • [32] D. R. Terno, “Inconsistency of quantum classical dynamics and what it implies,” Found. Phys.  36, 102 (2006) [quant-ph/0402092].
  • [33] J. L. Alonso, A. Castro, J. Clemente-Gallardo, J. C. Cuch/’i, P. Echenique, and F. Falceto, “Geometry and Statistics of Ehrenfest dynamics,” arXiv:1010.1494 [quant-ph].
  • [34] H. -T. Elze, “Linear dynamics of quantum-classical hybrids,” Phys. Rev. A 85, 052109 (2012) [arXiv:1111.2276 [quant-ph]].
  • [35] H. -T. Elze, “Four questions for quantum-classical hybrid theory,” J. Phys. Conf. Ser.  361, 012004 (2012) [arXiv:1202.3448 [quant-ph]].
  • [36] H. -T. Elze, “Quantum-classical hybrid dynamics a summary,” J. Phys. Conf. Ser.  442, 012007 (2013).
  • [37] H. R. Jauslin and D. Signy, “Dynamics of mixed classical-quantum systems, geometric quantization and coherent states,” in Mathematical Horizons for Quantum Physics, edited by H. Araki, B.-G. Englert, L.-C. Kwek, and J. Suzuki, Lecture Notes Series, Institute for Mathematical Sciences, National University of Sinpapore, Vol. 20 (World Scientific, Singapore, 2010) [arXiv:1111.5774 [quant-ph]].
  • [38] L. L. Salcedo, “Statistical consistency of quantum-classical hybrids,” Phys. Rev. A 85, 022127 (2012) [arXiv:1201.4237 [quant-ph]].
  • [39] C. Barcelo, R. Carballo-Rubio, L. J. Garay and R. Gomez-Escalante, “Hybrid classical-quantum formulations ask for hybrid notions,” Phys. Rev. A 86, 042120 (2012) [arXiv:1206.7036 [quant-ph]].
  • [40] L. N. Chang, Z. Lewis, D. Minic and T. Takeuchi, “Galois Field Quantum Mechanics,” Mod. Phys. Lett. B 27, 1350064 (2013) [arXiv:1205.4800 [quant-ph]].
  • [41] L. N. Chang, Z. Lewis, D. Minic and T. Takeuchi, “Spin and Rotations in Galois Field Quantum Mechanics,” J. Phys. A: Math. Theor. 46, 065304 (2013) [arXiv:1206.0064 [quant-ph]].
  • [42] B. Schumacher and M. D. Westmoreland, “Modal Quantum Theory,” Foundations of Physics 42 (2012) 910–925 [arXiv:1010.2929 [quant-ph]].
  • [43] B. Schumacher and M. D. Westmoreland, “Non-contextuality and free will in modal quantum theory,” arXiv:1010.5452 [quant-ph].
  • [44] M. D. Westmoreland and B. Schumacher, “Possibility, probability and entanglement: Non-contextuality in modal quantum mechanics,” in the Proceedings of Foundations of Probability and Physics 6 (FPP6), June 13-16, 2011, VäxjĂś, Sweden, AIP Conference Proceedings 1424, pp. 364-368 (2012).
  • [45] B. Schumacher and M. D. Westmoreland, “Almost quantum theory,” arXiv:1204.0701 [quant-ph].
  • [46] R. P. James, G. Ortiz, and A. Sabry, “Quantum Computing over Finite Fields,” arXiv:1101.3764 [quant-ph].
  • [47] J. Willcock and A. Sabry, “Solving UNIQUE-SAT in a Modal Quantum Theory,” arXiv:1102.3587 [quant-ph].
  • [48] A. J. Hanson, G. Ortiz, A. Sabry, and J. Willcock, “The Power of Discrete Quantum Theories,” arXiv:1104.1630 [quant-ph].
  • [49] A. J. Hanson, G. Ortiz, A. Sabry, and Y.-T. Tai, “Geometry of Discrete Quantum Computing,” J. Phys. A: Math. Theor. 46, 185301 (2013) [arXiv:1206.5823 [quant-ph]].
  • [50] A. J. Hanson, G. Ortiz, A. Sabry, and Y.-T. Tai, “Discrete Quantum Theories,” J. Phys. A: Math. Theor. 47, 115305 (2014) [arXiv:1305.3292 [quant-ph]].
  • [51] L. N. Chang, Z. Lewis, D. Minic and T. Takeuchi, “Biorthogonal Quantum Mechanics: Super-Quantum Correlations and Expectation Values without Definite Probabilities,” J. Phys. A: Math. Theor. 46, 485306 (2013) [arXiv:1208.5189 [math-ph]].
  • [52] T. Takeuchi, L. N. Chang, Z. Lewis and D. Minic, “Some Mutant Forms of Quantum Mechanics,” in the Proceedings of Quantum Theory: Reconsideration of Foundations 6 (QTRF6), June 11-14, 2012, VäxjĂś, Sweden, AIP Conference Proceedings 1508 [arXiv:1208.5544 [quant-ph]].
  • [53] L. N. Chang, Z. Lewis, D. Minic and T. Takeuchi, Int. J. Mod. Phys. A 29, 1430006 (2014), and in the Proceedings of the Conference in Honour of the 90th Birthday of Freeman Dyson, 26-29 August 2013, Institute of Advanced Studies at Nanyang Technological University, Singapore [arXiv:1312.0645 [quant-ph]].
  • [54] D. Ellerman, “Quantum mechanics over sets,” arXiv:1310.8221 [quant-ph].
  • [55] D. Ellerman, “Partitions and Objective Indefiniteness in Quantum Mechanics,” arXiv:1401.2421 [quant-ph].
  • [56] T. Curtright and L. Mezincescu, “Biorthogonal quantum systems,” J. Math. Phys.  48, 092106 (2007) [quant-ph/0507015].
  • [57] T. Curtright, L. Mezincescu and D. Schuster, “Supersymmetric biorthogonal quantum systems,” J. Math. Phys.  48, 092108 (2007) [quant-ph/0603170].
  • [58] J. Tits, “Sur les analogues algebriques des groupes semi-simple complexes,” Colloque d’algebre superieure, tenu a Bruxelles du 19 au 22 decembre 1956, Centre Belge de Recherches Math., Gauthier-Villar, Paris 1957, 261–289.
  • [59] E. W. Weisstein, “qq-Analog,” from MathWorld – a Wolfram Web Resource, http://mathworld.wolfram.com/q-Analog.html.
  • [60] Y. I. Manin, “Lectures on zeta functions and motives (according to Deninger and Kurokawa),” AstĂŠrisque 228 (1995) 121-163.
  • [61] Y. I. Manin, “Cyclotomy and analytic geometry over 𝔽1\mathbb{F}_{1},” arXiv:0809.1564 [math.AG].
  • [62] Y. Zhu, “Combinatorics and characteristic one algebra,” preprint, 2000.
  • [63] C. SoulĂŠ, “Les variĂŠtĂŠs sur le corps Ă  un ĂŠlĂŠment,” Moscow Math. J. 4 (2004) 217-244 [arXiv:math/0304444v1 [math.AG]].
  • [64] C. SoulĂŠ, “Lectures on algebraic varieties over 𝔽1,\mathbb{F}_{1},” in Noncommutative Geometry, Arithmetic, and Related Topics, Proceedings of the Twenty-First Meeting of the Japan-U.S. Mathematics Institute, edited by C. Consani and A. Connes (Johns Hopkins University Press, 2012) 267–278.
  • [65] N. Kurokawa, H. Ochiai, and M. Wakayama, “Absolute derivations and zeta functions,” Documenta Math, Extra Volume: Kazuya Kato’s Fiftieth Birthday (2003) 565–584.
  • [66] N. Kurokawa, “Zeta functions over 𝔽1\mathbb{F}_{1},” Proc. Japan Acad. 81, Ser. A (2005) 180-184.
  • [67] S. Koyama and N. Kurokawa, “Absolute zeta functions and absolute tensor products,” in Noncommutative Geometry, Arithmetic, and Related Topics, Proceedings of the Twenty-First Meeting of the Japan-U.S. Mathematics Institute, edited by C. Consani and A. Connes (Johns Hopkins University Press, 2012) 225–240.
  • [68] A. Deitmar, “Schemes over 𝔽1\mathbb{F}_{1},” in Number Fields and Function Fields – Two Parallel Worlds, edited by G. van der Geer, B. J. J. Moonen, and R. Schoof, Prog. Math. 239 (2005) pp. 87-100 [arXiv:math/0404185 [math.NT]].
  • [69] A. Deitmar, “Remarks on zeta functions and KK-theory over 𝔽1\mathbb{F}_{1},” Proc. Japan Acad. 82, Ser. A (2006) 141–146 [arXiv:math/0605429 [math.NT]].
  • [70] A. Deitmar, “𝔽1\mathbb{F}_{1}-schemes and toric varieties,” Contributions to Algebra and Geometry 49 (2008) pp. 517–525 [arXiv:math/0608179v9 [math.NT]].
  • [71] A. Deitmar, S. Koyama, and N. Kurokawa, “Absolute zeta functions,” Proc. Japan Acad. 84, Ser. A (2008) 138-142.
  • [72] J. Borger, “Λ\Lambda-rings and the field with one element,” arXiv:0906.3146v1 [math.NT].
  • [73] O. Lorscheid, “Algebraic groups over the field with one element,” Math. Zeitschrift 271 (2012) 117–138 [arXiv:0907.3824v1 [math.AG]].
  • [74] J. LĂłpez PeĂąa and O. Lorscheid, “Mapping 𝔽1\mathbb{F}_{1}-Land: An overview of geometries over the field with one element,” in Noncommutative Geometry, Arithmetic, and Related Topics, Proceedings of the Twenty-First Meeting of the Japan-U.S. Mathematics Institute, edited by C. Consani and A. Connes (Johns Hopkins University Press, 2012) 241–266 [arXiv:0909.0069v1 [math.AG]].
  • [75] L. Le Bruyn, “(Non)Commutative 𝔽un\mathbb{F}_{\mathrm{un}} Geometry,” arXiv:0909.2522v1 [math.RA].
  • [76] A. Connes, C. Consani, and M. Marcolli, “Fun with 𝔽1\mathbb{F}_{1},” arXiv:0806.2401v1 [math.AG].
  • [77] A. Connes and C. Consani, “On the notion of geometry over 𝔽1\mathbb{F}_{1},” J. Algebraic. Geom. 20 (2011) 525-557 [arXiv:0809.2926v2 [math.AG]].
  • [78] A. Connes and C. Consani, “Schemes over 𝔽1\mathbb{F}_{1} and Zeta Functions,” [arXiv:0903.2024v3 [math.AG]].
  • [79] A. Connes and C. Consani, “Characteristic 1, Entropy and the Absolute Point,” in Noncommutative Geometry, Arithmetic, and Related Topics, Proceedings of the Twenty-First Meeting of the Japan-U.S. Mathematics Institute, edited by C. Consani and A. Connes (Johns Hopkins University Press, 2012) 75–140 [arXiv:0911.3537v1 [math.AG]].
  • [80] A. Connes and C. Consani, “From Monoids to Hyperstructures: In Search of an Absolute Arithmetic,” [arXiv:1006.4810v1 [math.AG]].
  • [81] A. Connes, “The Witt Construction in Characteristic One and Quantization,” [arXiv:1009.1769v1 [math.QA]].
  • [82] P. Lescot, “Algèbre absolue,” Annales des sciences mathĂŠmatiques du QuĂŠbec 33, 1 (2009) 63–82 [arXiv:0911.1989 [math.RA]].
  • [83] P. Lescot, “Absolute Algebra II – Ideals and spectra” J. Pure Applied Algebra 215, 7 (2011) 1782–1790.
  • [84] P. Lescot, “Absolute Algebra III – The saturated spectrum” J. Pure Applied Algebra 216, 5 (2012) 1004–1015 [arXiv:1310.8399 [math.RA]].
  • [85] N. Kurokawa and S. Koyama, Absolute Mathematics (Nippon Hyōron Sha, 2010) (in Japanese).
  • [86] N. Kurokawa and S. Koyama, Intoduction to the ABC Conjecture (PHP Institute, 2013) (in Japanese).
  • [87] S. Mochizuki,
    http://www.kurims.kyoto-u.ac.jp/∟\simmotizuki/papers-english.html.
  • [88] M. V. Berry and J. P. Keating, “H=x​pH=xp and the Riemann zeros,” in Supersymmetry and Trace Formulae: Chaos and Disorder, Editors: I. V. Lerner, J. P. Keating and D. E. Khmelnitskii, NATO Science Series B, Physics Vol. 370 (Springer, 1999).
  • [89] J. W. P. Hirschfeld, Projective Geometries over Finite Fields, 2nd ed. (Oxford University Press, 1998).
  • [90] J. W. P. Hirschfeld, G. KorchmĂĄros, and F. Torres, Algebraic Curves over a Finite Field (Princeton University Press, 2008).
  • [91] V. I. Arnold, Dynamics, Statistics and Projective Geometry of Galois Fields (Cambridge University Press, 2011).
  • [92] S. Ball and Z. Weiner, An Introduction to Finite Geometry,
    http://www-ma4.upc.es/˜simeon/IFG.pdf
  • [93] J. S. Bell, “On the Einstein Podolsky Rosen paradox,” Physics 1 (1964) 195.
  • [94] J. S. Bell, Speakable and Unspeakable in Quantum Mechanics, Cambridge University Press (1987).
  • [95] J. F. Clauser, M. A. Horne, A. Shimony and R. A. Holt, “Proposed experiment to test local hidden variable theories,” Phys. Rev. Lett.  23 (1969) 880.
  • [96] B. S. Cirel’son, “Quantum Generalizations of Bell’s Inequality,” Lett. Math. Phys. 4 (1980) 93.
  • [97] L. J. Landau, “On the violation of Bell’s inequality in quantum theory,” Phys. Lett. A 120 (1987) 52.
  • [98] S. Popescu and D. Rohrlich, “Nonlocality as an Axiom,” Foundations of Physics, 24 (1994) 379.
  • [99] C. Chevalley, “Sur certains groups simples,” Tōhoku Math. Journ. 7 (2) (1955), 14–66.
  • [100] H. Cohn, “Projective Geometry over 𝔽1\mathbb{F}_{1} and the Gaussian Binomial Coefficients,” American Mathematical Monthly 111 (2004) 487–495 [arXiv:math/0407093 [math.CO]].
  • [101] P. J. Cameron, Combinatorics: Topics, Techniques, Algorithms (Cambridge University Press, 1994).
  • [102] R. Penrose, “Angular momentum: an approach to combinatorial space-time,”
    in Quantum Theory and Beyond, Essays and Discussions Arising from a Colloquium,
    edited by T. Bastin (Cambridge University Press, 1971)
    http://math.ucr.edu/home/baez/penrose/Penrose-AngularMomentum.pdf
  • [103] R. Penrose, “On the nature of quantum geometry,”
    in Magic Without Magic: John Archbald Wheeler,
    edited by J. R. Klauder (Freeman, 1972) p. 333–354,
    http://math.ucr.edu/home/baez/penrose/Penrose-OnTheNatureOfQuantumGeometry.pdf