HTML conversions sometimes display errors due to content that did not convert correctly from the source. This paper uses the following packages that are not yet supported by the HTML conversion tool. Feedback on these issues are not necessary; they are known and are being worked on.

  • failed: tasks

Authors: achieve the best HTML results from your LaTeX submissions by following these best practices.

License: arXiv.org perpetual non-exclusive license
arXiv:2403.05230v1 [quant-ph] 08 Mar 2024

Maximal Non-Kochen-Specker Sets and
a Lower Bound on the Size of Kochen-Specker Sets

Tom Williams [email protected] Wadham College, University of Oxford, Parks Rd, Oxford OX1 3PN, UK    Andrei Constantin [email protected] Rudolf Peierls Centre for Theoretical Physics, University of Oxford, Parks Road, Oxford OX1 3PU, UK
Abstract

A Kochen-Specker (KS) set is a finite collection of vectors on the two-sphere containing no antipodal pairs for which it is impossible to assign 0s and 1s such that no two orthogonal vectors are assigned 1 and exactly one vector in every triplet of mutually orthogonal vectors is assigned 1. The existence of KS sets lies at the heart of Kochen and Specker’s argument against non-contextual hidden variable theories and the Conway-Kochen free will theorem. Identifying small KS sets can simplify these arguments and may contribute to the understanding of the role played by contextuality in quantum protocols. In this paper we derive a weak lower bound of 10101010 vectors for the size of any KS set by studying the opposite notion of large non-KS sets and using a probability argument that is independent of the graph structure of KS sets. We also point out an interesting connection with a generalisation of the moving sofa problem around a right-angled hallway on S2superscript𝑆2S^{2}italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

I Introduction

Contextuality is the property of models in which the measurement outcomes depend on the set of measurements being performed together. Put differently, in contextual models the outcome of a measurement is not solely determined by the properties of the system but is influenced by the context of other measurements.

Quantum mechanics is fundamentally a contextual model as first demonstrated in 1967 by Kochen and Specker [1]. Since then, quantum contextuality has been a major topic of research, both as a fundamentally non-classical phenomenon and a key resource for quantum computation [2, 3, 4]. Several mathematical frameworks have been developed to study contextuality, including operational approaches [5], sheaf theory [6], the framework of contextuality-by-default [7], graph and hypergraph theory [8, 9]. All proofs of the Kochen-Specker (KS) theorem, as well as the free will theorem of Conway and Kochen [10] rely on the identification of small sets of measurements that exhibit contextuality, known as KS sets. The current record for the smallest KS set was found by Conway and Kochen and consists of 31 measurement directions [11]. For illustration, a conceptually easy to follow KS set is given in the following section containing 168 directions arranged in several nested Clifton graphs.

The main contribution of this paper is to explore the opposite notion of large, measurable non-contextual sets (non-KS sets). We find that the maximal non-KS set must cover at least a fraction of 0.8978 of the area of the two-sphere. We use this information to derive a lower bound on the number of directions contained in any KS set in three dimensions. Although this lower bound does not improve upon the current lower bound of 22 directions [12, 13], our argument is qualitatively different and does not rely on exhaustively checking the existence of KS graphs.

II Background

Hidden Variables. Quantum mechanics is non-deterministic: rather than predicting explicit outcomes of measurements as it is the case in classical mechanics, it merely predicts probabilities for such outcomes, whose indeterminacy is constrained by the uncertainty principle. Suppose this indeterminacy is not fundamental but is instead due to our incomplete knowledge of the system (akin to classical statistical mechanics); in this case, we may postulate the existence of a valuation map vψ(A^)subscript𝑣𝜓^𝐴{v}_{\psi}(\widehat{A})italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG italic_A end_ARG ), which is to be understood as the value of the observable (self-adjoint operator) A^^𝐴\widehat{A}over^ start_ARG italic_A end_ARG when the quantum state is ψ𝜓\psi\in\mathcal{H}italic_ψ ∈ caligraphic_H, where \mathcal{H}caligraphic_H is the Hilbert space of the system.

It is impossible to make progress beyond this postulate without imposing additional conditions, the most natural one being the following. For any function f::𝑓f:\mathbb{R}\to\mathbb{R}italic_f : roman_ℝ → roman_ℝ,

vψ(f(A^))=f(vψ(A^)).subscript𝑣𝜓𝑓^𝐴𝑓subscript𝑣𝜓^𝐴{v}_{\psi}(f(\widehat{A}))=f({v}_{\psi}(\widehat{A}))~{}.italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( italic_f ( over^ start_ARG italic_A end_ARG ) ) = italic_f ( italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG italic_A end_ARG ) ) . (1)

From this condition it follows (see, e.g., Ref. [14]) that for any commuting self-adjoint operators A^,B^^𝐴^𝐵\widehat{A},\widehat{B}over^ start_ARG italic_A end_ARG , over^ start_ARG italic_B end_ARG and any state ψ𝜓\psi\in\mathcal{H}italic_ψ ∈ caligraphic_H,

vψ(A^+B^)subscript𝑣𝜓^𝐴^𝐵\displaystyle{v}_{\psi}(\widehat{A}+\widehat{B})italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG italic_A end_ARG + over^ start_ARG italic_B end_ARG ) =vψ(A^)+vψ(B^),absentsubscript𝑣𝜓^𝐴subscript𝑣𝜓^𝐵\displaystyle={v}_{\psi}(\widehat{A})+{v}_{\psi}(\widehat{B})~{},= italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG italic_A end_ARG ) + italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG italic_B end_ARG ) , (2)
vψ(A^B^)subscript𝑣𝜓^𝐴^𝐵\displaystyle{v}_{\psi}(\widehat{A}\widehat{B})italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG italic_A end_ARG over^ start_ARG italic_B end_ARG ) =vψ(A^)vψ(B^),absentsubscript𝑣𝜓^𝐴subscript𝑣𝜓^𝐵\displaystyle={v}_{\psi}(\widehat{A}){v}_{\psi}(\widehat{B})~{},= italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG italic_A end_ARG ) italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG italic_B end_ARG ) ,
vψ(𝟙^)subscript𝑣𝜓^double-struck-𝟙\displaystyle{v}_{\psi}(\widehat{\mathbb{1}})italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG blackboard_𝟙 end_ARG ) =1.absent1\displaystyle=1~{}.= 1 .

Consider now an arbitrary orthonormal basis of \mathcal{H}caligraphic_H, {|e1,|e2,,}\{\ket{{e}_{1}},\ket{{e}_{2}},...,\}{ | start_ARG italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ⟩ , | start_ARG italic_e start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ⟩ , … , }. The projectors P^i:=|eiei|assignsubscript^𝑃𝑖ketsubscript𝑒𝑖brasubscript𝑒𝑖{\widehat{P}}_{i}:=\ket{{e}_{i}}\bra{{e}_{i}}over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT := | start_ARG italic_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ⟩ ⟨ start_ARG italic_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG | satisfy

P^i2=P^i,P^i=𝟙.formulae-sequencesuperscriptsubscript^𝑃𝑖2subscript^𝑃𝑖subscript^𝑃𝑖double-struck-𝟙{\widehat{P}}_{i}^{2}={\widehat{P}}_{i}~{},\qquad\sum{\widehat{P}}_{i}=\mathbb% {1}~{}.over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , ∑ over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = blackboard_𝟙 . (3)

Hence vψ(P^i)=vψ(P^i2)=vψ(P^i)2subscript𝑣𝜓subscript^𝑃𝑖subscript𝑣𝜓superscriptsubscript^𝑃𝑖2subscript𝑣𝜓superscriptsubscript^𝑃𝑖2{v}_{\psi}({\widehat{P}}_{i})={v}_{\psi}({\widehat{P}}_{i}^{2})={{v}_{\psi}({% \widehat{P}}_{i})}^{2}italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) = italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) = italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and vψ(P^i)=0 or 1subscript𝑣𝜓subscript^𝑃𝑖0 or 1{v}_{\psi}({\widehat{P}}_{i})=0\text{ or }1italic_v start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ( over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) = 0 or 1. Consequently, for any orthonormal basis {|e1,|e2,,}\{\ket{{e}_{1}},\ket{{e}_{2}},...,\}{ | start_ARG italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ⟩ , | start_ARG italic_e start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ⟩ , … , } exactly one projection operator P^i=|eiei|subscript^𝑃𝑖ketsubscript𝑒𝑖brasubscript𝑒𝑖{\widehat{P}}_{i}=\ket{{e}_{i}}\bra{{e}_{i}}over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = | start_ARG italic_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ⟩ ⟨ start_ARG italic_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG | must have value 1 and all others must have value 0. This is equivalent to the colouring problem in which no two orthogonal directions can both be coloured 1 and exactly one direction in any orthogonal basis is coloured 1, all other directions being coloured 0.

The Kochen-Specker (KS) Theorem. The KS theorem states that no such valuation map exists for Hilbert spaces of dimension greater than 2.

Proofs of the KS theorem often take the form of constructing Kochen-Specker sets, defined as finite sets of rank-one projectors for which it is impossible to consistently assign outcomes such that

  1. (KS.1)

    no two orthogonal directions have outcome 1;

  2. (KS.2)

    exactly one direction in any orthogonal frame has outcome 1.

Kochen and Specker’s original proof [1] consisted of 117 directions in three dimensions built from several smaller sets with 8 directions each and known in the literature as Clifton graphs [15]. Figure 1 shows a Clifton graph, denoted by 𝒢1subscript𝒢1{\mathcal{G}}_{1}caligraphic_G start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, together with its bundle diagram. Note that some edges in the bundle diagram are triangles rather than lines as some measurement contexts consist of three measurements. Clifton graphs are the smallest example of so-called 01-gadgets which are sets of directions such that in any {0,1}01\{0,1\}{ 0 , 1 }-colouring there exist two non-orthogonal directions that cannot both be assigned the colour 1 (see Ref. [16] for details). Clifton graphs enjoy the following useful property.

Refer to caption
Refer to caption
Figure 1: Left: the Clifton graph 𝒢1subscript𝒢1{\mathcal{G}}_{1}caligraphic_G start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT where vertices are directions and edges join orthogonal directions. Right: the bundle diagram of 𝒢1subscript𝒢1{\mathcal{G}}_{1}caligraphic_G start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT.
Proposition 1.

Directions a0subscript𝑎0{a}_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and a7subscript𝑎7{a}_{7}italic_a start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT in Figure 1 cannot both have outcome 1.

Proof.

Suppose v(a0)=v(a7)=1𝑣subscript𝑎0𝑣subscript𝑎71v({a}_{0})=v({a}_{7})=1italic_v ( italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = italic_v ( italic_a start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT ) = 1, then condition (KS.1) implies v(a1)=v(a2)=v(a3)=v(a4)=0𝑣subscript𝑎1𝑣subscript𝑎2𝑣subscript𝑎3𝑣subscript𝑎40v({a}_{1})=v({a}_{2})=v({a}_{3})=v({a}_{4})=0italic_v ( italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) = italic_v ( italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = italic_v ( italic_a start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) = italic_v ( italic_a start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) = 0, then condition (KS.2) implies v(a5)=v(a6)=1𝑣subscript𝑎5𝑣subscript𝑎61v({a}_{5})=v({a}_{6})=1italic_v ( italic_a start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT ) = italic_v ( italic_a start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT ) = 1. However, condition (KS.1) does not allow this, hence the contradiction. ∎

Clifton graphs suffer from the limitation that the angle between the directions a0subscript𝑎0{a}_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and a7subscript𝑎7{a}_{7}italic_a start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT in Figure 1 cannot exceed arccos(13)arccos13\text{arccos}(\frac{1}{3})arccos ( divide start_ARG 1 end_ARG start_ARG 3 end_ARG ) (see, e.g. , Ref. [16] for a proof). As such, several Clifton graphs are required to produce a KS set (or, using the language of Ref. [6], a strongly contextual set). Kochen and Specker’s proof uses 15 Clifton graphs and is shown in Figure 2.

Refer to caption
Figure 2: Kochen and Specker’s 117 direction proof. Each solid line connects two orthogonal directions and each dashed line represents a Clifton graph 𝒢1subscript𝒢1{\mathcal{G}}_{1}caligraphic_G start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. The following identifications hold: a=a𝑎superscript𝑎a=a^{\prime}italic_a = italic_a start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, b=b𝑏superscript𝑏b=b^{\prime}italic_b = italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, and c=c𝑐superscript𝑐c=c^{\prime}italic_c = italic_c start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT.

A New KS Set. A conceptually simple KS set consisting of 168 directions can be built using the idea of nested Clifton graphs [16], defined as follows.

Consider the Clifton graph 𝒢1subscript𝒢1{\mathcal{G}}_{1}caligraphic_G start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT but replace the edge (a5,a6)subscript𝑎5subscript𝑎6({a}_{5},{a}_{6})( italic_a start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT , italic_a start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT ) with a second Clifton graph 𝒢1superscriptsubscript𝒢1{\mathcal{G}}_{1}^{\prime}caligraphic_G start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT with a0=a5superscriptsubscript𝑎0subscript𝑎5{a}_{0}^{\prime}={a}_{5}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_a start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT and a7=a6superscriptsubscript𝑎7subscript𝑎6{a}_{7}^{\prime}={a}_{6}italic_a start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_a start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT. The condition that directions a0subscript𝑎0{a}_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and a7subscript𝑎7{a}_{7}italic_a start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT cannot both have outcome 1 is maintained. This process can be repeated iteratively. Denote an n𝑛nitalic_n-times nested Clifton graph by 𝒢nsubscript𝒢𝑛{\mathcal{G}}_{n}caligraphic_G start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT with vectors a0(1),,a7(1),,a0(n),,a7(n)superscriptsubscript𝑎01superscriptsubscript𝑎71superscriptsubscript𝑎0𝑛superscriptsubscript𝑎7𝑛{a}_{0}^{(1)},...,{a}_{7}^{(1)},...,{a}_{0}^{(n)},...,{a}_{7}^{(n)}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT , … , italic_a start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT , … , italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT , … , italic_a start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT, where a5(k)=a0(k1)superscriptsubscript𝑎5𝑘superscriptsubscript𝑎0𝑘1{a}_{5}^{(k)}={a}_{0}^{(k-1)}italic_a start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT = italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k - 1 ) end_POSTSUPERSCRIPT and a6(k)=a7(k1)superscriptsubscript𝑎6𝑘superscriptsubscript𝑎7𝑘1{a}_{6}^{(k)}={a}_{7}^{(k-1)}italic_a start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT = italic_a start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k - 1 ) end_POSTSUPERSCRIPT (see Figure 3). Then the following result holds (see Ref. [16] for a proof).

Proposition 2.

In a nested Clifton graph 𝒢nsubscript𝒢𝑛{\mathcal{G}}_{n}caligraphic_G start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT the directions a0(n)superscriptsubscript𝑎0𝑛{a}_{0}^{(n)}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT and a7(n)superscriptsubscript𝑎7𝑛{a}_{7}^{(n)}italic_a start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT cannot both have outcome 1. The maximum angle between a0(n)superscriptsubscript𝑎0𝑛{a}_{0}^{(n)}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT and a7(n)superscriptsubscript𝑎7𝑛{a}_{7}^{(n)}italic_a start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT is 𝑎𝑟𝑐𝑐𝑜𝑠(nn+2)𝑎𝑟𝑐𝑐𝑜𝑠𝑛𝑛2\text{arccos}(\frac{n}{n+2})arccos ( divide start_ARG italic_n end_ARG start_ARG italic_n + 2 end_ARG ).

Refer to caption
Figure 3: Nested Clifton graph, where […] indicate further iterations.
Refer to caption
Figure 4: A 168-direction KS set. Solid lines join orthogonal directions. Dashed lines represent nested Clifton graphs.

Using three 𝒢1subscript𝒢1{\mathcal{G}}_{1}caligraphic_G start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT graphs and six 𝒢4subscript𝒢4{\mathcal{G}}_{4}caligraphic_G start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT graphs, we construct the 168 direction KS set shown in Figure 4. The idea is simple: consider two orthogonal frames (triangles) and attach a nested Clifton graph between each direction (vertex) and the three directions of the other triangle. One of the three directions of a given frame must have outcome 1 by (KS.2), which in turn guarantees that the three directions of the other frame have outcome 0. This contradicts (KS.2) for the other frame, hence the set is KS.

III Maximal Non-KS Sets

Much of the literature on Kochen-Specker systems has focused on identifying minimal KS sets in three dimensions111The reason for which dimension three is special is two-fold. Every three-dimensional KS set can also be seen as a higher-dimensional one. Moreover, in dimensions four and higher, the size of the smallest KS set has already been found [17]. In dimension three, the same question remains open, currently with a sizeable gap between Uijlen and Westerbaan’s lower bound of 22 and Conway’s graph of 31 directions. [13, 12, 11]: Peres found a set with 33 directions [18], and the smallest set to date consisting of 31 directions was found by Conway [11]. By exhaustively constructing KS graphs and testing for topological embeddability, Uijlen and Westerbaan found a lower bound of 22 directions in three dimensions [12]. The opposite question of finding maximal non-KS sets has not been explored.222Simmons [19] discusses a measure of contextuality defined to be the smallest fraction of a KS set that cannot consistently be assigned a value 0 or 1. This construction is similar to our notion of maximal non-KS sets but is defined for finite sets. We do this here, starting with the definition of a non-KS set.

III.1 Non-KS Sets

Definition 3.

A non-KS set in d𝑑ditalic_d-dimensions is a measurable subset ASd1𝐴superscript𝑆𝑑1A\subseteq{S}^{d-1}italic_A ⊆ italic_S start_POSTSUPERSCRIPT italic_d - 1 end_POSTSUPERSCRIPT that admits a valuation map v:A{0,1}normal-:𝑣normal-→𝐴01v:A\to\{0,1\}italic_v : italic_A → { 0 , 1 } such that

  1. (NKS.1)

    v(n)=v(n)𝑣𝑛𝑣𝑛v(-n)=v(n)italic_v ( - italic_n ) = italic_v ( italic_n ) for all directions n𝑛nitalic_n in A𝐴Aitalic_A.

  2. (NKS.2)

    iIv(ni)1subscript𝑖𝐼𝑣subscript𝑛𝑖1\sum\limits_{i\in I}v({n}_{i})\leq 1∑ start_POSTSUBSCRIPT italic_i ∈ italic_I end_POSTSUBSCRIPT italic_v ( italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ≤ 1 for all sets of mutually orthogonal directions {ni}iIsubscriptsubscript𝑛𝑖𝑖𝐼{\{{n}_{i}\}}_{i\in I}{ italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_i ∈ italic_I end_POSTSUBSCRIPT in A𝐴Aitalic_A.

  3. (NKS.3)

    iIv(ni)=1subscript𝑖𝐼𝑣subscript𝑛𝑖1\sum\limits_{i\in I}v({n}_{i})=1∑ start_POSTSUBSCRIPT italic_i ∈ italic_I end_POSTSUBSCRIPT italic_v ( italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) = 1 for all sets of d𝑑ditalic_d mutually orthogonal directions {ni}iIsubscriptsubscript𝑛𝑖𝑖𝐼{\{{n}_{i}\}}_{i\in I}{ italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_i ∈ italic_I end_POSTSUBSCRIPT in A𝐴Aitalic_A.

Problem. Find a non-KS set ASd1𝐴superscript𝑆𝑑1A\subseteq{S}^{d-1}italic_A ⊆ italic_S start_POSTSUPERSCRIPT italic_d - 1 end_POSTSUPERSCRIPT with maximal measure.

It is clear that any maximal non-KS set must be antipodally symmetric from the following theorem.

Proposition 4.

The union of any non-KS set with its antipodal set is also a non-KS set.

Proof.

Consider a non-KS set A𝐴Aitalic_A with some valuation map v:A{0,1}:𝑣𝐴01v:A\to\{0,1\}italic_v : italic_A → { 0 , 1 }. Denote by A:={nSd1:nA}assign𝐴conditional-set𝑛superscript𝑆𝑑1𝑛𝐴-A:=\{-n\in{S}^{d-1}:n\in A\}- italic_A := { - italic_n ∈ italic_S start_POSTSUPERSCRIPT italic_d - 1 end_POSTSUPERSCRIPT : italic_n ∈ italic_A } the antipodal set of A𝐴Aitalic_A. Consider the map

v:AA\displaystyle v^{\prime}:A\cup-Aitalic_v start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT : italic_A ∪ - italic_A {0,1}absent01\displaystyle\to\{0,1\}→ { 0 , 1 }
n𝑛\displaystyle nitalic_n {v(n)nAv(n)nA\Amaps-toabsentcases𝑣𝑛𝑛𝐴𝑣𝑛𝑛\𝐴𝐴\displaystyle\mapsto\begin{cases}v(n)&n\in A\\ v(-n)&n\in-A\backslash A\end{cases}↦ { start_ROW start_CELL italic_v ( italic_n ) end_CELL start_CELL italic_n ∈ italic_A end_CELL end_ROW start_ROW start_CELL italic_v ( - italic_n ) end_CELL start_CELL italic_n ∈ - italic_A \ italic_A end_CELL end_ROW

This map trivially satisfies the three conditions in the definition of non-KS sets. ∎

In dimension d=3𝑑3d=3italic_d = 3 the problem is equivalent to finding the largest subset such that no two orthogonal directions have value 1 and no three mutually orthogonal directions have value 0. It is instructive to first consider the simpler problems of identifying:

  1. i)

    the largest subset containing no two orthogonal directions;

  2. ii)

    the largest subset containing no three mutually orthogonal directions.

The first of these problems was posed by Witsenhausen in 1974 [20]. The problem has not been solved but the following conjecture by Kalai and Wilson [21] has not yet been disproved.

Conjecture 5.

The maximal measure for a subset of S2superscript𝑆2S^{2}italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT containing no two orthogonal directions is attained by two open caps of geodesic radius π/4𝜋4\pi/4italic_π / 4 around the south and north poles.

The conjecture would imply that the measure is at most 11/21121-1/\sqrt{2}1 - 1 / square-root start_ARG 2 end_ARG as a fraction of the measure of the total sphere. DeCorte and Pikhurko [22] have proven that the maximal measure is at most 0.3130.3130.3130.313.

The second problem is an interesting variant of the first and is discussed by A. Perez [23]. In the next subsection we construct a large set with no three mutually orthogonal directions and conjecture that it is the largest such set.

III.2 Constructing a Large Non-KS Set

In order to construct a subset with no three mutually orthogonal directions, we begin with the union of two double caps. Since a double cap cannot contain two orthogonal directions, by the pigeonhole principle the union of two double caps cannot contain three mutually orthogonal directions (as one would have to contain two).

Without loss of generality, fix the poles of the double caps to be the positive and negative x𝑥xitalic_x and y𝑦yitalic_y axes. In terms of Cartesian coordinates, the boundaries of two of the caps on the upper (positive z𝑧zitalic_z) hemisphere have the following parametrizations:

𝐫1(t)=(12,t,12t2),subscript𝐫1𝑡12𝑡12superscript𝑡2{\mathbf{r}}_{1}(t)=\left(\frac{1}{\sqrt{2}},t,\sqrt{\frac{1}{2}-{t}^{2}}% \right),bold_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) = ( divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG , italic_t , square-root start_ARG divide start_ARG 1 end_ARG start_ARG 2 end_ARG - italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (4)
𝐫2(t)=(u(t),12,12u(t)2).subscript𝐫2𝑡𝑢𝑡1212𝑢superscript𝑡2{\mathbf{r}}_{2}(t)=\left(u(t),\frac{1}{\sqrt{2}},\sqrt{\frac{1}{2}-{u(t)}^{2}% }\right).bold_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) = ( italic_u ( italic_t ) , divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG , square-root start_ARG divide start_ARG 1 end_ARG start_ARG 2 end_ARG - italic_u ( italic_t ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) . (5)

Define the function u(t)𝑢𝑡u(t)italic_u ( italic_t ) such that 𝐫1(t)subscript𝐫1𝑡{\mathbf{r}}_{1}(t)bold_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) and 𝐫2(t)subscript𝐫2𝑡{\mathbf{r}}_{2}(t)bold_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) are orthogonal:

u(t)=12+t1+t.𝑢𝑡12𝑡1𝑡u(t)=-\frac{\frac{1}{2}+t}{1+t}~{}.italic_u ( italic_t ) = - divide start_ARG divide start_ARG 1 end_ARG start_ARG 2 end_ARG + italic_t end_ARG start_ARG 1 + italic_t end_ARG . (6)

Substituting into 𝐫2(t)subscript𝐫2𝑡{\mathbf{r}}_{2}(t)bold_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) gives

𝐫2(t)=(12+t1+t,12,1212t21+t).subscript𝐫2𝑡12𝑡1𝑡121212superscript𝑡21𝑡{\mathbf{r}}_{2}(t)=\left(-\frac{\frac{1}{2}+t}{1+t},\frac{1}{\sqrt{2}},\frac{% 1}{\sqrt{2}}\frac{\sqrt{\frac{1}{2}-{t}^{2}}}{1+t}\right).bold_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) = ( - divide start_ARG divide start_ARG 1 end_ARG start_ARG 2 end_ARG + italic_t end_ARG start_ARG 1 + italic_t end_ARG , divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG , divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG divide start_ARG square-root start_ARG divide start_ARG 1 end_ARG start_ARG 2 end_ARG - italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG start_ARG 1 + italic_t end_ARG ) . (7)

The third mutually orthogonal direction on the upper hemisphere is given by

𝐫3(t)=𝐫1(t)×𝐫2(t)=(1212t21+t,12t2,12+t+t21+t)subscript𝐫3𝑡subscript𝐫1𝑡subscript𝐫2𝑡1212superscript𝑡21𝑡12superscript𝑡212𝑡superscript𝑡21𝑡\begin{split}{\mathbf{r}}_{3}(t)&={\mathbf{r}}_{1}(t)\times{\mathbf{r}}_{2}(t)% \\ &=\left(-\frac{1}{\sqrt{2}}\frac{\sqrt{\frac{1}{2}-{t}^{2}}}{1+t},-\sqrt{\frac% {1}{2}-{t}^{2}},\frac{\frac{1}{2}+t+{t}^{2}}{1+t}\right)\end{split}start_ROW start_CELL bold_r start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL = bold_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) × bold_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = ( - divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG divide start_ARG square-root start_ARG divide start_ARG 1 end_ARG start_ARG 2 end_ARG - italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG start_ARG 1 + italic_t end_ARG , - square-root start_ARG divide start_ARG 1 end_ARG start_ARG 2 end_ARG - italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , divide start_ARG divide start_ARG 1 end_ARG start_ARG 2 end_ARG + italic_t + italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 1 + italic_t end_ARG ) end_CELL end_ROW (8)

The three parametrizations 𝐫1(t),𝐫2(t),𝐫3(t)subscript𝐫1𝑡subscript𝐫2𝑡subscript𝐫3𝑡{\mathbf{r}}_{1}(t),{\mathbf{r}}_{2}(t),{\mathbf{r}}_{3}(t)bold_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) , bold_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) , bold_r start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_t ) on the domain t[12,0)𝑡120t\in\left[-\frac{1}{2},0\right)italic_t ∈ [ - divide start_ARG 1 end_ARG start_ARG 2 end_ARG , 0 ) together with their π/2𝜋2\pi/2italic_π / 2 rotations about the z𝑧zitalic_z axis define the boundary of a closed region AS2𝐴superscript𝑆2A\subseteq{S}^{2}italic_A ⊆ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT which contains the north pole. We can parameterize the entire boundary through the function 𝐑(t)𝐑𝑡{\mathbf{R}}(t)bold_R ( italic_t ) defined as follows. Let

Rπ2:=(010100001)assignsubscript𝑅𝜋2matrix010100001R_{\frac{\pi}{2}}:=\begin{pmatrix}0&-1&0\\ 1&0&0\\ 0&0&1\end{pmatrix}italic_R start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT := ( start_ARG start_ROW start_CELL 0 end_CELL start_CELL - 1 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 1 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 1 end_CELL end_ROW end_ARG ) (9)

denote the rotation by π/2𝜋2\pi/2italic_π / 2 about the z𝑧zitalic_z axis. Using this, define

𝐫0:[0,1/4):subscript𝐫0014\displaystyle{\mathbf{r}}_{0}:\left[0,1/4\right)bold_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT : [ 0 , 1 / 4 ) S2absentsuperscript𝑆2\displaystyle\to S^{2}→ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
t𝑡\displaystyle titalic_t {𝐫1(6t12)t[ 0,112)Rπ23𝐫2(6t1)t[112,16)Rπ22𝐫3(6t32)t[16,14)maps-toabsentcasessubscript𝐫16𝑡12𝑡 0112superscriptsubscript𝑅𝜋23subscript𝐫26𝑡1𝑡11216superscriptsubscript𝑅𝜋22subscript𝐫36𝑡32𝑡1614\displaystyle\mapsto\begin{cases}{\mathbf{r}}_{1}(6t-\frac{1}{2})&t\in[\,0,% \frac{1}{12})\\ R_{\frac{\pi}{2}}^{3}{\mathbf{r}}_{2}(6t-1)&t\in[\,\frac{1}{12},\frac{1}{6})\\ R_{\frac{\pi}{2}}^{2}{\mathbf{r}}_{3}(6t-\frac{3}{2})&t\in[\,\frac{1}{6},\frac% {1}{4})\end{cases}↦ { start_ROW start_CELL bold_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( 6 italic_t - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ) end_CELL start_CELL italic_t ∈ [ 0 , divide start_ARG 1 end_ARG start_ARG 12 end_ARG ) end_CELL end_ROW start_ROW start_CELL italic_R start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 6 italic_t - 1 ) end_CELL start_CELL italic_t ∈ [ divide start_ARG 1 end_ARG start_ARG 12 end_ARG , divide start_ARG 1 end_ARG start_ARG 6 end_ARG ) end_CELL end_ROW start_ROW start_CELL italic_R start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_r start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( 6 italic_t - divide start_ARG 3 end_ARG start_ARG 2 end_ARG ) end_CELL start_CELL italic_t ∈ [ divide start_ARG 1 end_ARG start_ARG 6 end_ARG , divide start_ARG 1 end_ARG start_ARG 4 end_ARG ) end_CELL end_ROW (10)

and further,

𝐑:[ 0,1):𝐑 01\displaystyle{\mathbf{R}}:[\,0,1)bold_R : [ 0 , 1 ) S2absentsuperscript𝑆2\displaystyle\to S^{2}→ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
t𝑡\displaystyle titalic_t {𝐫0(t)t[ 0,14)Rπ2𝐫0(t14)t[14,12)Rπ22𝐫0(t12)t[12,34)Rπ23𝐫0(t34)t[34,1).maps-toabsentcasessubscript𝐫0𝑡𝑡 014subscript𝑅𝜋2subscript𝐫0𝑡14𝑡1412superscriptsubscript𝑅𝜋22subscript𝐫0𝑡12𝑡1234superscriptsubscript𝑅𝜋23subscript𝐫0𝑡34𝑡341\displaystyle\mapsto\begin{cases}{\mathbf{r}}_{0}(t)&t\in[\,0,\frac{1}{4})\\ R_{\frac{\pi}{2}}{\mathbf{r}}_{0}(t-\frac{1}{4})&t\in[\,\frac{1}{4},\frac{1}{2% })\\ R_{\frac{\pi}{2}}^{2}{\mathbf{r}}_{0}(t-\frac{1}{2})&t\in[\,\frac{1}{2},\frac{% 3}{4})\\ R_{\frac{\pi}{2}}^{3}{\mathbf{r}}_{0}(t-\frac{3}{4})&t\in[\,\frac{3}{4},1)\end% {cases}~{}.↦ { start_ROW start_CELL bold_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL italic_t ∈ [ 0 , divide start_ARG 1 end_ARG start_ARG 4 end_ARG ) end_CELL end_ROW start_ROW start_CELL italic_R start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT bold_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t - divide start_ARG 1 end_ARG start_ARG 4 end_ARG ) end_CELL start_CELL italic_t ∈ [ divide start_ARG 1 end_ARG start_ARG 4 end_ARG , divide start_ARG 1 end_ARG start_ARG 2 end_ARG ) end_CELL end_ROW start_ROW start_CELL italic_R start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ) end_CELL start_CELL italic_t ∈ [ divide start_ARG 1 end_ARG start_ARG 2 end_ARG , divide start_ARG 3 end_ARG start_ARG 4 end_ARG ) end_CELL end_ROW start_ROW start_CELL italic_R start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t - divide start_ARG 3 end_ARG start_ARG 4 end_ARG ) end_CELL start_CELL italic_t ∈ [ divide start_ARG 3 end_ARG start_ARG 4 end_ARG , 1 ) end_CELL end_ROW . (11)

It is convenient to extend the domain of 𝐑𝐑\mathbf{R}bold_R to the reals as t:𝐑(t)𝐑(tt):for-all𝑡𝐑𝑡𝐑𝑡𝑡\forall t\in\mathbb{R}:\mathbf{R}(t)\equiv\mathbf{R}(t-\lfloor t\rfloor)∀ italic_t ∈ roman_ℝ : bold_R ( italic_t ) ≡ bold_R ( italic_t - ⌊ italic_t ⌋ ), making 𝐑𝐑\mathbf{R}bold_R a periodic function of period 1111.

Lemma 6.

The following result hold:

t:s[t+13,t+23):𝑹(t)𝑹(s)0:for-all𝑡for-all𝑠𝑡13𝑡23:𝑹𝑡𝑹𝑠0\forall t\in\mathbb{R}:\forall s\in\left[\,t+\frac{1}{3},t+\frac{2}{3}\right):% \textbf{R}(t)\cdot\textbf{R}(s)\leq 0∀ italic_t ∈ roman_ℝ : ∀ italic_s ∈ [ italic_t + divide start_ARG 1 end_ARG start_ARG 3 end_ARG , italic_t + divide start_ARG 2 end_ARG start_ARG 3 end_ARG ) : R ( italic_t ) ⋅ R ( italic_s ) ≤ 0 (12)
Proof.

The inequality follows from standard maximization of 𝐑(t)𝐑(s)𝐑𝑡𝐑𝑠\textbf{R}(t)\cdot\textbf{R}(s)R ( italic_t ) ⋅ R ( italic_s ). ∎

Lemma 7.

Consider the open octant O(t)𝑂𝑡O(t)italic_O ( italic_t ) defined by the mutually orthogonal points 𝐑(t)𝐑𝑡\mathbf{R}(t)bold_R ( italic_t ), 𝐑(t+13)𝐑𝑡13\mathbf{R}(t+\frac{1}{3})bold_R ( italic_t + divide start_ARG 1 end_ARG start_ARG 3 end_ARG ) and 𝐑(t+23)𝐑𝑡23\mathbf{R}(t+\frac{2}{3})bold_R ( italic_t + divide start_ARG 2 end_ARG start_ARG 3 end_ARG ) for some t𝑡normal-ℝt\in\mathbb{R}italic_t ∈ roman_ℝ:

O(t):={𝐫S2|s{0,13,23}:𝐫𝐑(t+s)>0}.assign𝑂𝑡conditional-set𝐫superscript𝑆2:for-all𝑠01323𝐫𝐑𝑡𝑠0O(t):=\{\mathbf{r}\in S^{2}|\forall s\in\{0,\frac{1}{3},\frac{2}{3}\}:\mathbf{% r}\cdot\mathbf{R}(t+s)>0\}.italic_O ( italic_t ) := { bold_r ∈ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | ∀ italic_s ∈ { 0 , divide start_ARG 1 end_ARG start_ARG 3 end_ARG , divide start_ARG 2 end_ARG start_ARG 3 end_ARG } : bold_r ⋅ bold_R ( italic_t + italic_s ) > 0 } . (13)

Then O(t)¯Anormal-¯𝑂𝑡𝐴\overline{O(t)}\subset Aover¯ start_ARG italic_O ( italic_t ) end_ARG ⊂ italic_A.

Proof.

We begin by showing O(t)(A)=𝑂𝑡𝐴O(t)\cap\partial(A)=\emptysetitalic_O ( italic_t ) ∩ ∂ ( italic_A ) = ∅. For a contradiction, assume that u[ 0,1):𝐑(u)O(t):𝑢 01𝐑𝑢𝑂𝑡\exists u\in[\,0,1):\mathbf{R}(u)\in O(t)∃ italic_u ∈ [ 0 , 1 ) : bold_R ( italic_u ) ∈ italic_O ( italic_t ):

u[t+13,t+23)𝐑(u)𝐑(t)0𝐑(u)O(t)𝑢𝑡13𝑡23𝐑𝑢𝐑𝑡0𝐑𝑢𝑂𝑡\displaystyle u\in[\,t+\frac{1}{3},t+\frac{2}{3})\Rightarrow\mathbf{R}(u)\cdot% \mathbf{R}(t)\leq 0\Rightarrow\mathbf{R}(u)\notin O(t)italic_u ∈ [ italic_t + divide start_ARG 1 end_ARG start_ARG 3 end_ARG , italic_t + divide start_ARG 2 end_ARG start_ARG 3 end_ARG ) ⇒ bold_R ( italic_u ) ⋅ bold_R ( italic_t ) ≤ 0 ⇒ bold_R ( italic_u ) ∉ italic_O ( italic_t ) (14)
u[t+23,t+1)𝐑(u)𝐑(t+13)0𝐑(u)O(t)𝑢𝑡23𝑡1𝐑𝑢𝐑𝑡130𝐑𝑢𝑂𝑡\displaystyle u\in[\,t+\frac{2}{3},t+1)\Rightarrow\mathbf{R}(u)\cdot\mathbf{R}% (t+\frac{1}{3})\leq 0\Rightarrow\mathbf{R}(u)\notin O(t)italic_u ∈ [ italic_t + divide start_ARG 2 end_ARG start_ARG 3 end_ARG , italic_t + 1 ) ⇒ bold_R ( italic_u ) ⋅ bold_R ( italic_t + divide start_ARG 1 end_ARG start_ARG 3 end_ARG ) ≤ 0 ⇒ bold_R ( italic_u ) ∉ italic_O ( italic_t )
u[t+1,t+43)𝐑(u)𝐑(t+23)0𝐑(u)O(t)𝑢𝑡1𝑡43𝐑𝑢𝐑𝑡230𝐑𝑢𝑂𝑡\displaystyle u\in[\,t+1,t+\frac{4}{3})\Rightarrow\mathbf{R}(u)\cdot\mathbf{R}% (t+\frac{2}{3})\leq 0\Rightarrow\mathbf{R}(u)\notin O(t)italic_u ∈ [ italic_t + 1 , italic_t + divide start_ARG 4 end_ARG start_ARG 3 end_ARG ) ⇒ bold_R ( italic_u ) ⋅ bold_R ( italic_t + divide start_ARG 2 end_ARG start_ARG 3 end_ARG ) ≤ 0 ⇒ bold_R ( italic_u ) ∉ italic_O ( italic_t )

We have shown that u[t+13,t+43):𝐑(u)O(t):for-all𝑢𝑡13𝑡43𝐑𝑢𝑂𝑡\forall u\in[\,t+\frac{1}{3},t+\frac{4}{3}):\mathbf{R}(u)\notin O(t)∀ italic_u ∈ [ italic_t + divide start_ARG 1 end_ARG start_ARG 3 end_ARG , italic_t + divide start_ARG 4 end_ARG start_ARG 3 end_ARG ) : bold_R ( italic_u ) ∉ italic_O ( italic_t ), which is sufficient since 𝐑𝐑\mathbf{R}bold_R is periodic with period 1111. It follows trivially that:

O(t)int(A)ext(A).𝑂𝑡int𝐴ext𝐴O(t)\subset\text{int}(A)\cup\text{ext}(A).italic_O ( italic_t ) ⊂ int ( italic_A ) ∪ ext ( italic_A ) . (15)

Since O(t)𝑂𝑡O(t)italic_O ( italic_t ) is connected and (0,0,1)O(t)int(A)001𝑂𝑡int𝐴(0,0,1)\in O(t)\cap\text{int}(A)( 0 , 0 , 1 ) ∈ italic_O ( italic_t ) ∩ int ( italic_A ) it follows that

O(t)int(A)A𝑂𝑡int𝐴𝐴O(t)\subset\text{int}(A)\subset Aitalic_O ( italic_t ) ⊂ int ( italic_A ) ⊂ italic_A (16)

and finally

O(t)¯A¯=A.¯𝑂𝑡¯𝐴𝐴\overline{O(t)}\subset\overline{A}=A.over¯ start_ARG italic_O ( italic_t ) end_ARG ⊂ over¯ start_ARG italic_A end_ARG = italic_A . (17)

Lemma 8.

If two antipodal closed octants can perform complete rotations in the exterior of a region US2𝑈superscript𝑆2U\subset S^{2}italic_U ⊂ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, then U𝑈Uitalic_U contains no three mutually orthogonal directions.

Proof.

Rather than the octants rotating, we take the dual perspective of fixing the octants and rotating the region U𝑈Uitalic_U. Define the fixed antipodal closed octants of S2superscript𝑆2S^{2}italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, shown in Figure 5, to be

𝒪±:={(sin(θ)cos(ϕ),sin(θ)sin(ϕ),cos(θ))S2|(θ,ϕ)[ 0,π/2]×[ 0,π/2][π/2,π]×[π,3π/2]}.\begin{split}\mathcal{O}_{\pm}:=\{(\text{sin}(\theta)\text{cos}(\phi),\text{% sin}(\theta)\text{sin}(\phi),\text{cos}(\theta))\in S^{2}|~{}~{}~{}~{}~{}~{}~{% }~{}\\ (\theta,\phi)\in[\,0,\pi/2]\times[\,0,\pi/2]\ \cup\,[\,\pi/2,\pi]\times[\,\pi,% 3\pi/2]\,\}~{}.\end{split}start_ROW start_CELL caligraphic_O start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT := { ( sin ( italic_θ ) cos ( italic_ϕ ) , sin ( italic_θ ) sin ( italic_ϕ ) , cos ( italic_θ ) ) ∈ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | end_CELL end_ROW start_ROW start_CELL ( italic_θ , italic_ϕ ) ∈ [ 0 , italic_π / 2 ] × [ 0 , italic_π / 2 ] ∪ [ italic_π / 2 , italic_π ] × [ italic_π , 3 italic_π / 2 ] } . end_CELL end_ROW (18)

For any region U𝑈Uitalic_U to perform a complete rotation in ext(𝒪±)extsubscript𝒪plus-or-minus\text{ext}(\mathcal{O}_{\pm})ext ( caligraphic_O start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ), each point in U𝑈Uitalic_U must cross the line

L:={(sin(γ),cos(γ),0)S2|γ(0,π/2)}.assign𝐿conditional-setsin𝛾cos𝛾0superscript𝑆2𝛾0𝜋2L:=\{(-\text{sin}(\gamma),\text{cos}(\gamma),0)\in S^{2}|\gamma\in(0,\pi/2)\}~% {}.italic_L := { ( - sin ( italic_γ ) , cos ( italic_γ ) , 0 ) ∈ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_γ ∈ ( 0 , italic_π / 2 ) } . (19)

The set of points orthogonal to p=(sin(γ),cos(γ),0)L𝑝sin𝛾cos𝛾0𝐿p{=}(-\text{sin}(\gamma),\text{cos}(\gamma),0){\in}Litalic_p = ( - sin ( italic_γ ) , cos ( italic_γ ) , 0 ) ∈ italic_L on S2superscript𝑆2S^{2}italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT are

p:={(sin(θ)cos(γ),sin(θ)sin(γ),cos(θ))S2|ϕ[0,2π)}assignsuperscript𝑝perpendicular-toconditional-setsin𝜃cos𝛾sin𝜃sin𝛾cos𝜃superscript𝑆2italic-ϕ02𝜋\begin{split}p^{\perp}:=\{(\text{sin}(\theta)\text{cos}(\gamma),\text{sin}(% \theta)\text{sin}(\gamma),\text{cos}(\theta))\in S^{2}|&\\ \phi\in[0,2\pi)\}&\end{split}start_ROW start_CELL italic_p start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT := { ( sin ( italic_θ ) cos ( italic_γ ) , sin ( italic_θ ) sin ( italic_γ ) , cos ( italic_θ ) ) ∈ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_ϕ ∈ [ 0 , 2 italic_π ) } end_CELL start_CELL end_CELL end_ROW (20)

and the set of such points contained in ext(𝒪±)extsubscript𝒪plus-or-minus\text{ext}(\mathcal{O}_{\pm})ext ( caligraphic_O start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ) are

pext(𝒪±)={(sin(θ)cos(γ),sin(θ)sin(γ),cos(θ))S2|θ(π2,π)(3π2,2π)},superscript𝑝perpendicular-toextsubscript𝒪plus-or-minusconditional-setsin𝜃cos𝛾sin𝜃sin𝛾cos𝜃superscript𝑆2𝜃𝜋2𝜋3𝜋22𝜋\displaystyle\begin{split}p^{\perp}\cap\text{ext}(\mathcal{O}_{\pm})=\{(\text{% sin}(\theta)\text{cos}(\gamma),\text{sin}(\theta)\text{sin}(\gamma),\text{cos}% (\theta))\in S^{2}|&\\ \theta\in(\frac{\pi}{2},\pi)\cup(\frac{3\pi}{2},2\pi)\}~{},~{}~{}~{}~{}~{}~{}~% {}~{}~{}&\end{split}start_ROW start_CELL italic_p start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ∩ ext ( caligraphic_O start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ) = { ( sin ( italic_θ ) cos ( italic_γ ) , sin ( italic_θ ) sin ( italic_γ ) , cos ( italic_θ ) ) ∈ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_θ ∈ ( divide start_ARG italic_π end_ARG start_ARG 2 end_ARG , italic_π ) ∪ ( divide start_ARG 3 italic_π end_ARG start_ARG 2 end_ARG , 2 italic_π ) } , end_CELL start_CELL end_CELL end_ROW (21)

which contains no two orthogonal points. Hence, no point which crosses L𝐿Litalic_L can be part of a mutually orthogonal set of three points in U𝑈Uitalic_U. ∎

Refer to caption
Figure 5: The grey shaded region depicts 𝒪±subscript𝒪plus-or-minus\mathcal{O}_{\pm}caligraphic_O start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT. The green line and point depict L𝐿Litalic_L and p𝑝pitalic_p respectively. The purple line depicts pext(𝒪±)superscript𝑝perpendicular-toextsubscript𝒪plus-or-minusp^{\perp}\cap\text{ext}(\mathcal{O}_{\pm})italic_p start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ∩ ext ( caligraphic_O start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ) and the purple and grey points depict two orthogonal points which cannot lie within pext(𝒪±)superscript𝑝perpendicular-toextsubscript𝒪plus-or-minusp^{\perp}\cap\text{ext}(\mathcal{O}_{\pm})italic_p start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ∩ ext ( caligraphic_O start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ).
Proposition 9.

The region N0:=S2\(AA){{N}_{0}}:={S}^{2}\backslash(A\cup-A)italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT := italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT \ ( italic_A ∪ - italic_A ) contains no three mutually orthogonal directions.

Proof.

It follows from Lemma 7 and the definition of N0subscript𝑁0N_{0}italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT that (O(t)¯O(t)¯)N0=(\overline{O(t)}\cup-\overline{O(t)})\cap N_{0}=\emptyset( over¯ start_ARG italic_O ( italic_t ) end_ARG ∪ - over¯ start_ARG italic_O ( italic_t ) end_ARG ) ∩ italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ∅. Hence, two antipodal closed octants can perform complete rotations in ext(N0)extsubscript𝑁0\text{ext}(N_{0})ext ( italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ). Hence, it follows from Lemma 8 that N0subscript𝑁0N_{0}italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT contains no three mutually orthogonal directions. ∎

Remark 10.

A simpler proof for Proposition 9 is the following. Let u,v,wN0𝑢𝑣𝑤subscript𝑁0u,v,w\in N_{0}italic_u , italic_v , italic_w ∈ italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT be a set of mutually orthogonal vectors. Rotate around u𝑢uitalic_u until one or both v𝑣vitalic_v, w𝑤witalic_w land on the boundary of N0subscript𝑁0N_{0}italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. For concreteness, say v𝑣vitalic_v lands on the boundary, while w𝑤witalic_w may be on the boundary or not after rotation. Then u𝑢uitalic_u and w𝑤witalic_w must belong to the great circle rv=0𝑟𝑣0r\cdot v=0italic_r ⋅ italic_v = 0. However, by construction, the circle rv=0𝑟𝑣0r\cdot v=0italic_r ⋅ italic_v = 0 contains two quadrants which do not belong to N0subscript𝑁0N_{0}italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Since u𝑢uitalic_u is in N0subscript𝑁0N_{0}italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and w𝑤witalic_w at most on the boundary, it follows that u𝑢uitalic_u and w𝑤witalic_w cannot be perpendicular.

If we define N1subscript𝑁1{{N}_{1}}italic_N start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT to be the double cap with poles at the positive and negative z𝑧zitalic_z axis then N1N0=subscript𝑁1subscript𝑁0{{N}_{1}}\cap{{N}_{0}}=\emptysetitalic_N start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∩ italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ∅ and the union N:=N1N0assign𝑁subscript𝑁1subscript𝑁0{N}:={{N}_{1}}\cup{{N}_{0}}italic_N := italic_N start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∪ italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT forms a non-KS set with measure |N|=|N1|+|N0|𝑁subscript𝑁1subscript𝑁0|{N}|=|{{N}_{1}}|+|{{N}_{0}}|| italic_N | = | italic_N start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | + | italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | (see Figure 6). The measures can be calculated by standard integration:

|N1|subscript𝑁1\displaystyle|{{N}_{1}}|| italic_N start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | =14π(202πdϕ0π4sin(θ)dθ)=112,absent14𝜋2superscriptsubscript02𝜋italic-ϕsuperscriptsubscript0𝜋4sin𝜃𝜃112\displaystyle=\frac{1}{4\pi}\left(2\int_{0}^{2\pi}\differential{\phi}\int_{0}^% {\frac{\pi}{4}}\text{sin}(\theta)\ \differential{\theta}\right)=1-\frac{1}{% \sqrt{2}}~{},= divide start_ARG 1 end_ARG start_ARG 4 italic_π end_ARG ( 2 ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_π end_POSTSUPERSCRIPT roman_d start_ARG italic_ϕ end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT divide start_ARG italic_π end_ARG start_ARG 4 end_ARG end_POSTSUPERSCRIPT sin ( italic_θ ) roman_d start_ARG italic_θ end_ARG ) = 1 - divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG , (22)
|N0|subscript𝑁0\displaystyle|{{N}_{0}}|| italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | =12π[812012t2ddt(arctan(2t)))dt\displaystyle=\frac{1}{2\pi}\left[8\int_{-\frac{1}{2}}^{0}\sqrt{\frac{1}{2}-{t% }^{2}}\frac{\differential{}}{\differential{t}}\left(\text{arctan}(\sqrt{2}t))% \right)\differential{t}\right.= divide start_ARG 1 end_ARG start_ARG 2 italic_π end_ARG [ 8 ∫ start_POSTSUBSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT square-root start_ARG divide start_ARG 1 end_ARG start_ARG 2 end_ARG - italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG roman_d start_ARG end_ARG end_ARG start_ARG roman_d start_ARG italic_t end_ARG end_ARG ( arctan ( square-root start_ARG 2 end_ARG italic_t ) ) ) roman_d start_ARG italic_t end_ARG
+412012+t+t21+tddt(arctan(2(t+1)))dt]\displaystyle\left.+4\int_{-\frac{1}{2}}^{0}\frac{\frac{1}{2}+t+{t}^{2}}{1+t}% \frac{\differential{}}{\differential{t}}\left(\text{arctan}(\sqrt{2}(t+1))% \right)\differential{t}\right]+ 4 ∫ start_POSTSUBSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT divide start_ARG divide start_ARG 1 end_ARG start_ARG 2 end_ARG + italic_t + italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 1 + italic_t end_ARG divide start_ARG roman_d start_ARG end_ARG end_ARG start_ARG roman_d start_ARG italic_t end_ARG end_ARG ( arctan ( square-root start_ARG 2 end_ARG ( italic_t + 1 ) ) ) roman_d start_ARG italic_t end_ARG ]
=112+2πln(2)absent1122𝜋ln2\displaystyle=1-\frac{1}{\sqrt{2}}+\frac{\sqrt{2}}{\pi}\text{ln}(2)= 1 - divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG + divide start_ARG square-root start_ARG 2 end_ARG end_ARG start_ARG italic_π end_ARG ln ( 2 )

where we used the relations ϕ=arctan(y/z)italic-ϕarctan𝑦𝑧\phi=\text{arctan}(y/z)italic_ϕ = arctan ( italic_y / italic_z ), cosθ=z𝜃𝑧\cos\theta=zroman_cos italic_θ = italic_z for evaluating |N0|subscript𝑁0|N_{0}|| italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT |. It follows that

|N|=22+2ln(2)π0.8978,𝑁2222𝜋0.8978|{N}|=2-\sqrt{2}+\frac{\sqrt{2}\ln{2}}{\pi}\approx 0.8978,| italic_N | = 2 - square-root start_ARG 2 end_ARG + divide start_ARG square-root start_ARG 2 end_ARG roman_ln ( start_ARG 2 end_ARG ) end_ARG start_ARG italic_π end_ARG ≈ 0.8978 , (23)

where the measures are normalized such that |S2|=1superscript𝑆21|{S}^{2}|=1| italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | = 1.

This gives a lower bound for the maximal non-KS set Nmaxsubscript𝑁max{{N}_{\text{max}}}italic_N start_POSTSUBSCRIPT max end_POSTSUBSCRIPT:

|Nmax|22+2ln(2)π0.8978subscript𝑁max2222𝜋0.8978|{{N}_{\text{max}}}|\geq 2-\sqrt{2}+\frac{\sqrt{2}\ln{2}}{\pi}\approx 0.8978| italic_N start_POSTSUBSCRIPT max end_POSTSUBSCRIPT | ≥ 2 - square-root start_ARG 2 end_ARG + divide start_ARG square-root start_ARG 2 end_ARG roman_ln ( start_ARG 2 end_ARG ) end_ARG start_ARG italic_π end_ARG ≈ 0.8978 (24)
Conjecture 11.

N𝑁Nitalic_N is the non-KS set of largest measure.

Refer to caption
(a) 3D Plot
Refer to caption
(b) 2D Projection onto xy𝑥𝑦xyitalic_x italic_y-plane
Figure 6: Plots of the non-KS set N=N1N0𝑁subscript𝑁1subscript𝑁0{N}={{N}_{1}}\cup{{N}_{0}}italic_N = italic_N start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∪ italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT where N1subscript𝑁1{{N}_{1}}italic_N start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is plotted in red and N0subscript𝑁0{{N}_{0}}italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in blue. The grey regions depict an octant rotating within ext(N0)extsubscript𝑁0\text{ext}(N_{0})ext ( italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ).

III.3 Maximal Non-KS Sets and Minimal KS Sets

Say that a set USd1𝑈superscript𝑆𝑑1U\subseteq{S}^{d-1}italic_U ⊆ italic_S start_POSTSUPERSCRIPT italic_d - 1 end_POSTSUPERSCRIPT avoids a set PSd1𝑃superscript𝑆𝑑1P\subseteq{S}^{d-1}italic_P ⊆ italic_S start_POSTSUPERSCRIPT italic_d - 1 end_POSTSUPERSCRIPT if there exists no orientation of P𝑃Pitalic_P that lies entirely inside U𝑈Uitalic_U. By definition a non-KS set must avoid all KS sets. Hence a lower bound for the size of a KS set can be established by finding the smallest set of directions that any given non-KS set avoids. The following result by Alexandru Damian [24] provides a lower bound for the minimal number of points that do not fit in an arbitrary subset XSd1𝑋superscript𝑆𝑑1X\subseteq{S}^{d-1}italic_X ⊆ italic_S start_POSTSUPERSCRIPT italic_d - 1 end_POSTSUPERSCRIPT of measure |U|𝑈|U|| italic_U |.

Proposition 12.

If a subset USd1𝑈superscript𝑆𝑑1U\subseteq{S}^{d-1}italic_U ⊆ italic_S start_POSTSUPERSCRIPT italic_d - 1 end_POSTSUPERSCRIPT with measure |U|𝑈|U|| italic_U | avoids an n𝑛nitalic_n-point set P𝑃Pitalic_P then n11|U|𝑛11𝑈n\geq\frac{1}{1-|U|}italic_n ≥ divide start_ARG 1 end_ARG start_ARG 1 - | italic_U | end_ARG.

Proof.

Assume to the contrary that there is an n𝑛nitalic_n-point set P𝑃Pitalic_P with n<11|U|𝑛11𝑈n<\frac{1}{1-|U|}italic_n < divide start_ARG 1 end_ARG start_ARG 1 - | italic_U | end_ARG that U𝑈Uitalic_U avoids. Consider a random rotation of the set P={p1,,pn}𝑃subscript𝑝1subscript𝑝𝑛P=\{{p}_{1},...,{p}_{n}\}italic_P = { italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT }. Define a random indicator variable Xisubscript𝑋𝑖{X}_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for each point pisubscript𝑝𝑖{p}_{i}italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT.

Xi={1piU0otherwisesubscript𝑋𝑖cases1subscript𝑝𝑖𝑈0otherwise{X}_{i}=\begin{cases}1&{p}_{i}\in U\\ 0&\text{otherwise}\end{cases}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = { start_ROW start_CELL 1 end_CELL start_CELL italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∈ italic_U end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL otherwise end_CELL end_ROW (25)

The expectation value for each Xisubscript𝑋𝑖{X}_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is then

E[Xi]=P(Xi=1)=|U|>n1n𝐸delimited-[]subscript𝑋𝑖𝑃subscript𝑋𝑖1𝑈𝑛1𝑛E\left[{X}_{i}\right]=P({X}_{i}=1)=|U|>\frac{n-1}{n}italic_E [ italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] = italic_P ( italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 1 ) = | italic_U | > divide start_ARG italic_n - 1 end_ARG start_ARG italic_n end_ARG (26)

Defining by X𝑋Xitalic_X the total number of points that lie inside U𝑈Uitalic_U and computing its expectation using the linearity of expectation

E[X]=E[i=1nXi]=i=1nE[Xi]=n|U|>n1.𝐸delimited-[]𝑋𝐸delimited-[]superscriptsubscript𝑖1𝑛subscript𝑋𝑖superscriptsubscript𝑖1𝑛𝐸delimited-[]subscript𝑋𝑖𝑛𝑈𝑛1E\left[X\right]=E\left[\sum\limits_{i=1}^{n}{X}_{i}\right]=\sum\limits_{i=1}^{% n}E\left[{X}_{i}\right]=n|U|>n-1.italic_E [ italic_X ] = italic_E [ ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_E [ italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] = italic_n | italic_U | > italic_n - 1 . (27)

Since E[X]>n1𝐸delimited-[]𝑋𝑛1E\left[X\right]>n-1italic_E [ italic_X ] > italic_n - 1 there must be at least one rotation such that all n𝑛nitalic_n points lie inside U𝑈Uitalic_U. So U𝑈Uitalic_U does not avoid P𝑃Pitalic_P which is a contradiction. ∎

Theorem 13.

Consider a minimal KS set K𝑚𝑖𝑛subscript𝐾𝑚𝑖𝑛{K}_{\text{min}}italic_K start_POSTSUBSCRIPT min end_POSTSUBSCRIPT and a maximal non-KS set N𝑚𝑎𝑥subscript𝑁𝑚𝑎𝑥{{N}_{\text{max}}}italic_N start_POSTSUBSCRIPT max end_POSTSUBSCRIPT. The following inequality must be satisfied

|K𝑚𝑖𝑛|11|N𝑚𝑎𝑥|.subscript𝐾𝑚𝑖𝑛11subscript𝑁𝑚𝑎𝑥|{K}_{\text{min}}|\geq\frac{1}{1-|{{N}_{\text{max}}}|}.| italic_K start_POSTSUBSCRIPT min end_POSTSUBSCRIPT | ≥ divide start_ARG 1 end_ARG start_ARG 1 - | italic_N start_POSTSUBSCRIPT max end_POSTSUBSCRIPT | end_ARG . (28)

In particular, in three dimensions it follows that

|K𝑚𝑖𝑛|11|N𝑚𝑎𝑥|1212ln(2)π9.79.subscript𝐾𝑚𝑖𝑛11subscript𝑁𝑚𝑎𝑥12122𝜋9.79|{K}_{\text{min}}|\geq\frac{1}{1-|{{N}_{\text{max}}}|}\geq\frac{1}{\sqrt{2}-1-% \frac{\sqrt{2}\ln{2}}{\pi}}\approx 9.79.| italic_K start_POSTSUBSCRIPT min end_POSTSUBSCRIPT | ≥ divide start_ARG 1 end_ARG start_ARG 1 - | italic_N start_POSTSUBSCRIPT max end_POSTSUBSCRIPT | end_ARG ≥ divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG - 1 - divide start_ARG square-root start_ARG 2 end_ARG roman_ln ( start_ARG 2 end_ARG ) end_ARG start_ARG italic_π end_ARG end_ARG ≈ 9.79 . (29)

Hence a minimal KS set must contain at least 10 directions in three dimensions.

Proof.

The result follows directly from the above discussion. ∎

It is interesting to observe the overlap of some small KS sets and our non-KS set N𝑁Nitalic_N. Of course, the KS sets contain points that do not lie inside N𝑁Nitalic_N (in any orientation), but it is interesting to note, for example, that the Peres set is only just not covered by N𝑁Nitalic_N in the sense that N¯¯𝑁\overline{N}over¯ start_ARG italic_N end_ARG covers the Peres set (see Figure 7).

Refer to caption
(a) Conway’s 31 direction KS set
Refer to caption
(b) Peres’ 33 direction KS set
Figure 7: Plots of Conway’s 31 direction KS set and Peres’ 33 direction KS set overlaid on the non-KS set N𝑁Nitalic_N.

IV Moving Sofa

The original moving sofa problem, first formally discussed by Leo Moser [25] in 1966, is stated as follows:

Problem. What is the region of largest area which can be moved around a right-angled corner in a corridor of width one?

The area obtained is referred to as the sofa constant. The exact value of the sofa constant is an open problem, however there exist results on lower and upper bounds (see, e.g. Refs. [26, 27]).

Several variations of the moving sofa problem are said to have been considered by Conway, Shepard and other mathematicians [28] in the 1960’s. One such variation is the ambidextrous moving sofa problem requiring the region to be able to make both right and left right-angled turns, which has since been considered by Romik [29].

There is an interesting connection between the set N0subscript𝑁0{{N}_{0}}italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and a generalization of the moving sofa problem to a right-angled hallway defined on S2superscript𝑆2S^{2}italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT rather than 2superscript2\mathbb{R}^{2}roman_ℝ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

Proposition 14.

A lower bound for the area of the largest shape that can be moved around a right-angled corner on the unit two-sphere is given by

|N0|=112+2πln(2)subscript𝑁01122𝜋2|N_{0}|=1-\frac{1}{\sqrt{2}}+\frac{\sqrt{2}}{\pi}\ln(2)| italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | = 1 - divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG + divide start_ARG square-root start_ARG 2 end_ARG end_ARG start_ARG italic_π end_ARG roman_ln ( start_ARG 2 end_ARG ) (30)
Proof.

Consider ext(𝒪±)extsubscript𝒪plus-or-minus\text{ext}(\mathcal{O}_{\pm})ext ( caligraphic_O start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ) as defined in Lemma 8 to be the hallway, then it follows from Lemma 7 and the definition of N0subscript𝑁0N_{0}italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT that the region N0subscript𝑁0N_{0}italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT can be moved around ext(𝒪±)extsubscript𝒪plus-or-minus\text{ext}(\mathcal{O}_{\pm})ext ( caligraphic_O start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ). Hence, |N0|=112+2πln(2)subscript𝑁01122𝜋2|N_{0}|=1-\frac{1}{\sqrt{2}}+\frac{\sqrt{2}}{\pi}\ln(2)| italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | = 1 - divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG + divide start_ARG square-root start_ARG 2 end_ARG end_ARG start_ARG italic_π end_ARG roman_ln ( start_ARG 2 end_ARG ) is a lower bound for the sofa constant in this variation of the problem. ∎

V Conclusion

In this paper we have constructed a conceptually simple KS set of 168 directions. We have also developed the notion of large measurable non-KS sets that do no exhibit contextuality. We constructed a large non-KS set and calculated its measure to deduce a lower bound for the maximal measure of such a set. Relying on this lower bound, we used a probability argument to obtain a lower bound of 10 directions for the size of any KS set. Although this lower bound does not improve upon the current lower bound of 22 directions [12], it does not rely on exhaustively checking the existence of KS graphs. The probability argument of Proposition 12 is qualitatively different and does not take into account the geometry of the non-KS set. It is conceivable that including such information would lead to a stronger bound while avoiding the burden of exhaustive searches. As a byproduct, we obtained a lower bound for the constant in the moving sofa problem on S2superscript𝑆2S^{2}italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

Acknowledgements. TW would like to express his gratitude to the Oxford Physics Department for the award of the Gibbs Prize for the Best MPhys Research Project in 2023, on which this work is based. AC’s research is supported by a Royal Society Dorothy Hodgkin Fellowship.

References

  • [1] S. Kochen and E. P. Specker, “The Problem of Hidden Variables in Quantum Mechanics,” Journal of Mathematics and Mechanics 17 (1967) 59–87.
  • [2] M. Howard, J. Wallman, V. Veitch, and J. Emerson, “Contextuality supplies the ‘magic’ for quantum computation,” Nature 510 (June, 2014) 351–355, 1401.4174.
  • [3] R. Raussendorf, “Contextuality in measurement-based quantum computation,” Physical Review A 88 (Aug., 2013) 022322, 0907.5449.
  • [4] S. Abramsky, R. S. Barbosa, and S. Mansfield, “Contextual Fraction as a Measure of Contextuality,” Physical Review Letters 119 (Aug., 2017) 050504, 1705.07918.
  • [5] R. W. Spekkens, “Contextuality for Preparations, Transformations, and Unsharp Measurements,” Physical Review A 71 (2005), no. 5, 052108, 0406166.
  • [6] S. Abramsky and A. Brandenburger, “The sheaf-theoretic structure of non-locality and contextuality,” New Journal of Physics 13 (Nov., 2011) 113036, 1102.0264.
  • [7] E. N. Dzhafarov, J. V. Kujala, and V. H. Cervantes, “Contextuality-by-Default: A Brief Overview of Ideas, Concepts, and Terminology,” in Quantum Interaction, H. Atmanspacher, T. Filk, and E. Pothos, eds., pp. 12–23. Springer International Publishing, Cham, 2016. 1504.00530.
  • [8] A. Cabello, S. Severini, and A. Winter, “Graph-Theoretic Approach to Quantum Correlations,” Physical Review Letters 112 (Jan., 2014) 040401, 1401.7081.
  • [9] A. Acin, T. Fritz, A. Leverrier, and A. B. Sainz, “A Combinatorial Approach to Nonlocality and Contextuality,” Commun. Math. Phys. 334 (2015) 533–628, 1212.4084.
  • [10] J. Conway and S. Kochen, “The Strong Free Will Theorem,” Notices of the AMS 56 (2009), no. 2, 226–232, 0807.3286.
  • [11] J. Conway and S. Kochen, “The free will theorem,” Foundations of Physics 36 (2006) 1441–1473, 0604079.
  • [12] S. Uijlen and B. Westerbaan, “A Kochen-Specker system has at least 22 vectors,” New Generation Computing 34 (2016), no. 1-2, 3–23, 1412.8544.
  • [13] F. Arends, J. Ouaknine, and C. W. Wampler, “On searching for small Kochen-Specker vector systems,” in Graph-Theoretic Concepts in Computer Science: 37th International Workshop, WG 2011, Teplá Monastery, Czech Republic, June 21-24, 2011. Revised Papers 37, pp. 23–34, Springer. 2011. 1111.3301.
  • [14] C. Isham, Lectures On Quantum Theory Mathematical And Structural Foundations. Allied Publ., 2001.
  • [15] R. Clifton, “Getting contextual and nonlocal elements-of-reality the easy way,” American Journal of Physics 61 (May, 1993) 443–447.
  • [16] R. Ramanathan, M. Rosicka, K. Horodecki, S. Pironio, M. Horodecki, and P. Horodecki, “Gadget structures in proofs of the Kochen-Specker theorem,” Quantum 4 (2020) 308, 1807.00113.
  • [17] M. Pavičić, J.-P. Merlet, B. McKay, and N. D. Megill, “Kochen–Specker vectors,” Journal of Physics A: Mathematical and General 38 (2005), no. 7, 1577, 0409014.
  • [18] A. Peres, “Two simple proofs of the Kochen-Specker theorem,” Journal of Physics A: Mathematical and General 24 (feb, 1991) L175.
  • [19] A. W. Simmons, “How (maximally) contextual is quantum mechanics?,” Quantum, Probability, Logic: The Work and Influence of Itamar Pitowsky (2020) 505–519.
  • [20] H. S. Witsenhausen, “Spherical sets without orthogonal point pairs,” The American Mathematical Monthly 81 (1974), no. 10, 1101–1102.
  • [21] G. Kalai and R. Wilson, “How large can a spherical set without two orthogonal vectors be,” Combinatorics and more (weblog) (2009).
  • [22] E. DeCorte and O. Pikhurko, “Spherical sets avoiding a prescribed set of angles,” International Mathematics Research Notices 2016 (2016), no. 20, 6095–6117, 1502.05030.
  • [23] A. Perez, “On Generalizations of the Double Cap Conjecture,” 2019. Available at https://math.mit.edu/research/highschool/rsi/documents/2019Perez.pdf.
  • [24] J. O’Rourke, “Open Problems from CCCG 2015.,” in CCCG. 2015.
  • [25] L. Moser, “Moving furniture through a hallway,” SIAM Review 8 (1966), no. 3, 381.
  • [26] J. L. Gerver, “On moving a sofa around a corner,” Geometriae Dedicata 42 (1992), no. 3, 267–283.
  • [27] Y. Kallus and D. Romik, “Improved upper bounds in the moving sofa problem,” Advances in Mathematics 340 (2018) 960–982, 1706.06630.
  • [28] I. Stewart, Another fine math you’ve got me into… Courier Corporation, 2004.
  • [29] D. Romik, “Differential equations and exact solutions in the moving sofa problem,” Experimental Mathematics 27 (2018), no. 3, 316–330, 1606.08111.