<<

in quantum cognition

Nicole Yunger Halpern1, 2 and Elizabeth Crosson1 1Institute for Quantum Information and Matter, Caltech, Pasadena, CA 91125, USA 2Kavli Institute for Theoretical Physics, University of California, Santa Barbara, CA 93106, USA (Dated: November 15, 2017) Matthew Fisher recently postulated a mechanism by which quantum phenomena could influence cognition: Phosphorus nuclear spins may resist decoherence for long times. The spins would serve as biological . The qubits may resist decoherence longer when in Posner molecules. We imagine that Fisher postulates correctly. How adroitly could biological systems process quantum information (QI)? We establish a framework for answering. Additionally, we apply biological qubits in , quantum communication, and quantum computation. First, we posit how the QI encoded by the spins transforms as Posner molecules form. The transformation points to a natural computational basis for qubits in Posner molecules. From the basis, we construct a quantum code that detects arbitrary single- errors. Each molecule encodes one qutrit. Shifting from information storage to computation, we define the model of Posner quantum computation. To illustrate the model’s quantum-communication ability, we show how it can teleport information in- coherently: A state’s weights are teleported; the coherences are not. The dephasing results from the entangling operation’s simulation of a coarse-grained Bell measurement. Whether Posner quantum computation is universal remains an open question. However, the model’s operations can efficiently prepare a Posner state usable as a resource in universal measurement-based quantum computation. The state results from deforming the Affleck-Lieb-Kennedy-Tasaki (AKLT) state and is a projected entangled-pair state (PEPS). Finally, we show that entanglement can affect molecular-binding rates (by 0.6% in an example). This work opens the door for the QI-theoretic analysis of biological qubits and Posner molecules.

Fisher recently proposed a mechanism by which quan- This paper is intended for QI scientists, for chemists, tum phenomena might affect cognition [1]. Phosphorus and for biophysicists. Some readers may require back- atoms populate biochemistry. A phosphorus nucleus’s ground about the QI theory behind the results. We refer , he argued, can store quantum information (QI) for these readers to App. A and to [7, 8]. Next, we overview long times. The nucleus has a spin quantum number this paper’s contributions. 1 s = 2 . Hence the nucleus forms a qubit, a quantum two- Computational bases before and after molecule level system. The qubit is the standard unit of QI. formation: Phosphorus nuclear spins originate outside Fisher postulated physical processes that might en- Posners, in Fisher’s narrative. The spins occupy ions tangle phosphorus nuclei. Six phosphorus atoms might, that join together, forming Posners. Molecular formation with other ions, form Posner molecules Ca9(PO4)6 [2– changes how QI is encoded physically. 4].1 The molecules might protect the spins’ states for Outside of molecules, phosphorus nuclear spins couple long times. Fisher also described how the QI stored in little to orbital degrees of freedom (DOFs). Spin states the spins might be read out. This QI, he conjectured, form an obvious choice of computational basis.2 In a could impact neuron firing. The neurons could partici- Posner molecule, the spins are indistinguishable. They pate in quantum cognition. occupy an antisymmetric state [1, 5]: The spins entangle These conjectures require empirical testing. Fisher with orbital DOFs. Which physical states form a useful has proposed experiments [1], including with Radzi- computational basis is not obvious. hovsky [5]. Some experiments have begun [6]. We identify such a basis. Molecule formation, we posit, Suppose that Fisher conjectures correctly. How effec- maps premolecule spin states to antisymmetric molecule

arXiv:1711.04801v1 [quant-ph] 13 Nov 2017 tively could the spins process QI? We provide a frame- states deterministically. The premolecule orbital state work for answering this question, and we begin answer- determines the map. We formalize the map with a ing. We translate Fisher’s physics and chemistry into in- projector-valued measure (PVM). The mapped-to anti- formation theory. The language of molecular tumbling, symmetric states form the computational basis, in terms heat, etc. is replaced with the formalism of Bloch-sphere of which Posners’ QI processing should be expressed. rotations, positive operator-valued measures (POVMs), Quantum error-correcting and -detecting computational bases, etc. Additionally, we identify and codes: The basis elements may decohere quickly: Pos- quantify QI-storage, -communication, and -computation ners’ geometry protects only spins. The basis elements capacities of the phosphorus nuclear spins and Posners.

2 In QI, computations are expressed in terms of a com- putational basis for the system’s [7]. Ba- 1 Ca9(PO4)6 has been called the Posner cluster and Posner sis elements are often represented by bit strings, as in molecule. We call it the Posner, for short. {|00 ... 0i, |00 ... 01i,... |11 ... 1i}. 2 are spin-and-position entangled states. Do the dynamics satisfy τA + τB = 0 can be measured projectively. If protect any states against errors? the equation is satisfied, the twelve qubits can undergo Hamiltonians’ ground spaces may form quantum error- coordinated rotations. correcting and -detecting codes (QECD codes) [8]. One Finally, hextuples can cease to correspond to geome- might hope to relate the Posner Hamiltonian HPos to a tries or to GC ’s (as Posners break down into their con- 3 QECD code. Alas, HPos has not been characterized. stituent ions). Thereafter, qubits can rotate indepen- Yet HPos likely preserves two observables. One, GC , dently again, group together into new hextuples, etc. generates cyclic permutations of the spins. One such per- This model enables us to recast Fisher’s narrative [1] as mutation shuffles the spins counterclockwise about the a . We also identify a criterion necessary molecule’s symmetry axis, through an angle 2π/3. This for constructing, from Posner operations, quantum cir- permutation preserves the Posner’s geometry [1–4]. The cuits of nonconstant depth: Molecules must break down zin other charge, S12...6, is the spins’ total zin-component. and form anew. Only outside molecules can qubits un- (The internal-frame axisz ˆin remains fixed relative to the dergo arbitrary rotations. Only in molecules can qubits atoms’ positions.) undergo entangling operations late in a computation. To The dynamics likely preserve eigenstates shared by GC alternate between rotating and entangling, qubits must zin zin and S1...6. Yet GC shares many eigenbases with S1...6: leave and enter Posners. The charges fail to form a complete set of commuting Entanglement generated by, and quantum- observables (CSCO). We identify a useful operator that communication application of, molecular bind- 2 2 breaks the degeneracy: S123 ⊗ S456 equals a product of ing: Two Posners, Fisher conjectures, can bind to- the spin-squared operators S2 of trios of a Posner’s spins. gether [1]. Quantum-chemistry calculations support the This operator (i) respects the Posner’s geometry and (ii) conjecture [9]. The binding is expected to entangle the facilitates the construction of Posner states that can fuel Posners [1]. How much entanglement does binding gen- universal quantum computation (discussed below). erate, and entanglement of what sort? zin 2 From the eigenbasis shared by GC , S1...6, and S123 ⊗ We characterize the entanglement in two ways. First, 2 S456, we form QECD codes. A state |ψi in one charge we compare Posner binding to a Bell measurement [7]. A zin sector of GC and one sector of S1...6 likely cannot trans- Bell measurement yields one of four possible outcomes— form, under the dynamics, into a state |φi in a second two bits of information. Posner binding transforms a zin sector of GC and a second sector of S1...6. Hence |ψi subspace as a coarse-grained Bell measurement. A Bell and |φi suggest themselves as codewords. Charge preser- measurement is performed, and one bit is discarded, ef- vation would prevent one codeword from evolving into fectively. another. Second, we present a quantum-communication proto- We construct two quantum codes, each partially pro- col reliant on Posner binding. We define a qutrit (three- tected by charge preservation. Via one code, each Posner level) subspace of the Posner Hilbert space. A Posner P encodes one qutrit. The codewords correspond to distinct P2 may occupy a state |ψi = j=0 cj|ji in the subspace. eigenvalues of GC . This code detects arbitrary single- 2 The coefficients |cj| form a probability distribution Q. physical-qubit errors. Via the second code, each Posner This distribution has a probability p of being teleported encodes one qubit. This repetition code corrects two bit to another Posner, P 0. Another distribution, Q˜, consists flips. The codewords correspond to distinct eigenvalues | |2 ˜ − zin of combinations of the cj ’s. Q has a probability 1 p of of S1...6. being teleported. Measuring P 0 in the right basis would Model of Posner quantum computation: Fisher yield an outcome distributed according to Q or according posits chemical processes, such as binding, that Pos- to Q˜. ners may undergo [1]. We abstract away the chemistry, The weights of |ψi (or combinations of the weights) formalizing the computations effected by the processes. are teleported [10]. The coherences are not. We therefore Posner operations Pos- These effected form the model of dub the protocol incoherent teleportation. The dephasing ner quantum computation . comes from the binding’s simulation of a coarse-grained The model includes the preparation of singlets Bell measurement. Bell measurements teleport QI coher- √1 (|01i − |10i). Qubits can rotate arbitrarily when the 2 ently. phosphorus atoms are outside molecules. The qubits Incoherent teleportation effects a variant of superdense 1 evolve trivially, under the identity , when Posners form. coding [11]. A trit (a classical three-level system) is com- But Posner creation associates a hextuple of qubits with municated effectively, while a bit is communicated di- a geometry and with an observable GC . rectly. The trit is encoded superdensely in the bit, with A Posner’s six qubits can undergo identical arbitrary help from Posner binding. rotations. Also, measurements can be performed: GC Posner-molecule state that can serve as a uni- has eigenvalues τ = 0, ±1. Whether Posners A and B versal resource for measurement-based quantum computation: Measurement-based quantum computa- tion (MBQC) [12–14] is a quantum-computation model alternative to the circuit model [15]. MBQC begins with 3 A characterization may be expected in [9]. a many-body entangled state |ψi. Single qubits are mea- 3 sured adaptively. are satisfied, if sufficient control is available. Whether MBQC can efficiently simulate universal quantum the gate set is universal remains an open question. computation if begun with the right |ψi. Most quan- Organization of this paper: Section I reviews tum states cannot serve as universal resources [16]. Clus- Fisher’s proposal. Section II details the physical set- ter states [12, 17, 18] on 2D square lattices can [12, 13, up and models Posner creation. How Posner creation 19, 20]. So can the Affleck-Kennedy-Lieb-Tasaki (AKLT) changes the physical encoding of QI appears in Sec. III. state [21–23] on a honeycomb lattice, |AKLThoni. Local QECD codes are presented in Sec. IV. measurements can transform |AKLThoni into the univer- The model of Posner quantum computation is defined sal . Hence |AKLThoni can fuel universal in Sec. V. Posner binding is analyzed, and applied to in- MBQC [20, 24]. coherent teleportation, in Sec. VI. Section VII showcases 0 0 We define a variation |AKLThoni on |AKLThoni. the universal resource state |AKLThoni. 0 |AKLThoni can be prepared efficiently with Posner oper- Section VIII quantifies entanglement’s effect on ations. Preparing |AKLThoni, one projects onto a spin- molecular-binding probabilities. Quantum cognition is 3 0 2 subspace. Preparing |AKLThoni, one projects onto a compared with DiVincenzo’s criteria in Sec. IX. Oppor- slightly larger subspace. Local measurements (supple- tunities for further study are detailed in Sec. X. mented by Posner hydrolyzation, singlet formation, and 0 Posner creation) can transform |AKLThoni into the uni- 0 versal cluster state. Hence |AKLThoni can fuel universal I. REVIEW: FISHER’S MBQC as |AKLThoni can. QUANTUM-COGNITION PROPOSAL Whether Posner operations can implement the extra local measurements, or the adaptive measurements in Biological systems are warm, wet, and large. Such en- MBQC, remains an open question. Yet the universality of vironments quickly diminish quantum coherences. Fisher a Posner state, efficiently preparable by a (conjectured) catalogued the influences that could decohere nuclear biological system, is remarkable. Most states cannot fuel spins in biofluids. Examples include electric and mag- 0 universal MBQC [16]. The universality of |AKLThoni fol- netic fields generated by other nuclear spins and by elec- lows from (i) Posners’ geometry and (ii) their ability to trons. share singlets. These sources, Fisher estimated, decohere the 0 31 Like |AKLThoni, |AKLThoni is a projected entangled- phosphorus-31 ( P) nuclear spin slowly. Coherence pair state (PEPS) [25]. The state is formed from two times might reach ∼ 1 s or 105 − 106 s, depending on the basic tensors. Each tensor has three physical qubits ion or molecule occupied by the phosphorus. No other and three virtual legs. One virtual leg has bond di- biologically prevalent atom, Fisher conjectures, has such mension six. Each other virtual leg has bond dimen- a long-lived nuclear spin. 0 sion two. |AKLThoni is the unique ground state of some Phosphorus atoms inhabit many biological ions and frustration-free Hamiltonian H 0 [26, 27]. The re- 3− AKLT molecules. Examples include the phosphate ion, PO4 . lationship between HAKLT0 and HPos remains an open Three phosphates feature in the molecule adenosine question. So does whether HAKLT0 has a constant-size triphosphate (ATP). ATP stores energy that powers gap. chemical reactions. Two phosphates can detach from an Entanglement’s influence on binding probabili- ATP molecule, forming a diphosphate ion. A diphos- ties: Entanglement, Fisher conjectures, can affect Pos- phate can break into two phosphates, with help from the ners’ probability of binding together [1]. He imagined a enzyme pyrophosphatase. The two phosphates’ phospho- Posner A entangled with a Posner A0 and a B entangled rus nuclear spins form a singlet, Fisher and Radzihovsky with a B0. Suppose that A has bound to B. A0 more (F&R) conjecture [1, 5]. A singlet is a maximally en- likely binds to B0, Fisher argues, than in the absence of tangled state. Entanglement is a correlation, shareable entanglement. by quantum systems, stronger than any achievable by We test the principle behind Fisher’s conjecture, in a classical systems [7]. two-Posner example. Let A and B denote the Posners. Many biomolecules contain phosphate ions. Occupying First, we suppose that (i) neither Posner contains en- a small molecule, Fisher argues, could shelter the phos- tangled spins and (ii) A shares no entanglement with B. phorus nuclear spin: Small molecules tumble in fluids. Next, we suppose that (i) each Posner contains one sin- The average of an external field B, over uniform tum- glet and (ii) A shares one singlet with B. The entangle- bling, vanishes. A B whose average magnitude vanishes ment boosts the binding probability by ≈ 0.6%. Though cannot decohere spins quickly. small, the boost supports Fisher’s conjecture. Comput- Which small biomolecules could a phosphorus inhabit? ing power limited our test’s size. Yet our technique can An important candidate is Ca9(PO4)6. A Posner con- 3− be scaled up to Fisher’s four-Posner example. sists of six phosphate ions (PO4 ) and nine calcium ions Comparison with DiVincenzo’s criteria: DiVin- (Ca2+) [2–4]. Posners form in some simulated biofluids cenzo codified the criteria required for realizing quantum and possibly in vivo [29–31]. A Posner could conceivably computation and communication [28]. We compare the contain a phosphate that forms a singlet with a phos- criteria with Fisher’s narrative. At least most criteria phate in another Posner. The Posners would share en- 4 tanglement. ture [5]. We therefore ignore the electronic DOFs. We Two Posners can bind together, according to quantum- ignore calcium ions similarly. We focus on the DOFs that chemistry calculations [1, 9]. The binding projects the might store QI for long times. Posners onto a possibly entangled state. Moreover, pre- existing entanglement could affect the probability that Posners bind. II B. Posner-molecule geometry and notation Bindings, influenced by entanglement, could influence neuron firing. Suppose that a Posner A shares entangle- Quantum-chemistry calculations have shed light on the 0 ment with a Posner A and that a B shares entanglement shapes available to Posners [2–4, 9]. A Posner’s shape 0 with a B . Posners A and B could enter one neuron, depends on the environment. Posners in biofluids have 0 0 while A and B enter another. Suppose that A binds begun to be studied [4]. We follow [1, 5], supposing that with B. The binding, with entanglement, could raise the more-detailed studies will support [4]. 0 0 probability that A binds to B . The Posner forms a cube (Fig. 1). At each face’s center Bound-together Posners move slowly, Fisher ar- sits a phosphate. The molecule has one symmetry axis. gues. Compound molecules must displace many water The axis coincides with a diagonal of the cube. The molecules, which slow down the pair. Relatedly, the Pos- Posner remains invariant under 2π/3 rotations about this ner pair has a large moment of inertia. Hence the pair diagonal: The molecule has C3 symmetry. rotates more slowly than separated Posners by the con- This cube diagonal serves as the z-axisz ˆin of a refer- servation of angular momentum. ence frame fixed in the molecule. The atoms’ positions + Hydrogen ions H can attach easily to slow molecules, remain constant relative to this internal frame. The in- + Fisher expects. H hydrolyzes Posners, breaking the ternal frame can move relative to the lab frame, denoted molecules into their constituent ions. Hence entangle- by the subscript “lab.” The spins’ Bloch vectors are de- ment might correlate hydrolyzation of A and B with hy- fined with respect to the lab frame. 0 0 drolyzation of A and B . Hydrolyzation would release Imagine gazing down the diagonal, as in Fig 1a. You 2+ calcium ions Ca into the neurons. Suppose that many would see three phosphates that form a triangle. We la- entangled Posners hydrolyzed in these two neurons. The bel the triangle’s z -coordinate by h . One hidden phos- 2+ in + neurons’ Ca concentrations could rise. The neurons phate sits directly behind each visible phosphate. We la- could fire synchronously due to entanglement. bel by h− the hidden triangle’s zin-coordinate.z ˆin points oppositely the direction in which we imagined gazing. Hence h+ > h−. II. PHYSICAL SET-UP AND φ labels the triangles’ shared orientation, as shown in POSNER-MOLECULE CREATION Fig. 1b. We denote by ϕj the angular orientation of cube face j (the site of a phosphate): Imagine rotating This section concerns (i) the physical set-up and (ii) the xin-axis counterclockwise until it intersects the cube the joining together of phosphates (and calcium ions) in face’s center (a phosphate). The angle swept out is ϕj. Posner molecules. Part of the material appears in [1, 5] One visible and one invisible phosphate sit at the angle φ; and is reviewed. Part of the material has not, according another pair, at φ + 2π/3; and another pair, at φ + 4π/3. to our knowledge, appeared elsewhere. We label the center of cube face j (the site of phosphate The phosphorus nuclei are associated with spin and j) with an angle and a height: (ϕj, hj). spatial Hilbert spaces in Sec. II A. Section II B reviews, and introduces notation for, the Posner’s geometry. Sec- tion II C models the creation of a Posner from close- II C. Qualitative model for the creation of a together ions. The ions can have distinguishable DOFs Posner molecule before, but not after, forming a Posner. Posners form from phosphate and calcium ions. We model the formation process qualitatively in this section. II A. Spin and spatial Hilbert spaces We first review how, according to Fisher, phosphorus nu- clear spins might come to form singlets. We then envi- Each phosphorus nucleus has two relevant DOFs: a sion phosphates falling into a Lennard-Jones potential as spin and a position. We will sometimes call the position a Posner forms. F&R have discussed the indistinguisha- spin orb the orbital or spatial DOF. Let Hnuc and Hnuc denote the bility of phosphorus nuclei in a Posner [1, 5]. We ex- associated Hilbert spaces. The nucleus has a spin quan- pand upon this discussion, considering how distinguish- 1 spin 2 tum number of s = 2 . Hence Hnuc = C . The orbital able ions become indistinguishable. orb 3− Hilbert space is infinite-dimensional: dim(Hnuc) = ∞. Several molecules contain phosphate ions PO4 . Ex- Each phosphorus nucleus’s Hilbert space decomposes as amples include ATP (Sec. I). Each ATP molecule con- spin orb Hnuc = Hnuc ⊗ Hnuc. tains three phosphates (hence the “triphosphate”). Two The electrons’ states transform trivially under all rele- of the phosphates can break off, forming a diphosphate vant operations, Fisher and Radzihovsky (F&R) conjec- ion. The enzyme pyrophosphatase can hydrolyze a 5

zˆin

h+ V(x) Visible triangle Cube’s center h- Hidden Cube triangle diagonal x

FIG. 2: Lennard-Jones potential: The Lennard-Jones (a) a b potential, VLJ(x) = x12 − x6 , models van der Waals forces y between particles. The real parameters a, b > 0. We grossly in Center of approximate, with VLJ(x), the potential experienced by cube face Center of phosphate ions coalescing into a Posner molecule. x denotes cube face the distance from a phosphate to the system’s center of mass. Cube’s ϕ center xin The singlet has the form 1 |Ψ−i := √ (|↑i|↓i − |↓i|↑i) . (1) 2 Center of cube face The singlet is one of the four Bell pairs. The Bell pairs are mutually orthogonal, maximally entangled states of pairs (b) of qubits [7]. Bell pairs serve as units of entanglement in QI. FIG. 1: Posner-molecule geometry and coordinates: Quantum-chemistry calculations have shed light on the Phosphorus nuclei are identical fermions, as F&R em- Posner molecule’s geometry [2–4, 9]. The molecule forms a phasize [1, 5]. But some of the nuclei’s DOFs might be cube. At each cube face’s center sits one phosphate ion distinguishable before Posners form. Consider, for exam- 3− (PO4 ). The molecule appears to have one symmetry axis ple, two ATP molecules on opposite sides of a petri dish. when in biofluids [1, 4]. The axis coincides with a cube Call the molecules A and B. A diphosphate could break diagonal. Imagine gazing down the diagonal, as in Fig. 1a. off from each ATP molecule. Each diphosphate could hy- In the opposite direction points the internal z-axis,z ˆin. (The drolyze into two phosphates, A1 and A2 or B1 and B2. internal reference frame remains fixed relative to the atoms’ Consider the phosphorus nuclear spins of one phosphate positions.) Gazing down the diagonal, you see three pair—say, of A1 and A2. These spins would be indistin- phosphate ions (the large, black dots in Fig. 1b). The guishable: Neither nucleus could be associated with an phosphates form a triangle. φ denotes the least angle swept upward-pointing spin or with a downward-pointing spin. out counterclockwise from the +xin-axis to a phosphate. The triangle remains invariant under rotations, about the But the spatial DOF of A1 and A2 could be distin- symmetry axis, through an angle 2π/3. The long-dash line guished from the spatial DOF of B1 and B2: We can in Fig. 1b illustrates such a rotation. The invariance endows imagine painting phosphate pair A red and phosphate the Posner with C3 symmetry. Directly behind the visible pair B blue. The phosphate pairs could diffuse to the phosphates sit the other three phosphates (Fig. 1a). We dish’s center. The red pair and the blue pair could be denote the triangles’ zin-coordinates by h±. tracked along their trajectories. Consider six phosphates (and nine Ca2+ ions) ap- proaching each other. We model the ions qualitatively as subject to a Lennard-Jones potential. Such poten- tials feature in molecular-dynamics simulations [32]. The diphosphate, cleaving the ion into separated phosphates. model encapsulates interatomic forces’ key features. The separated phosphates contain phosphorus nuclear We temporarily approximate each phosphate as having spins that, Fisher conjectures [1], form a singlet. a classical position. Let x denote some phosphate’s dis- tance from the ions’ center of mass. Figure 2 illustrates Let 1 and 2 label the phosphorus nuclear spins. Let the Lennard-Jones potential, zˆenz denote the z-axis of a reference frame fixed in the ˆ a b enzyme. Let Szenz denote the zenz-component of a phos- − VLJ(x) = 12 6 . (2) phorus nucleus’s spin operator. Let |↑i and |↓i denote the x x ˆ ˆ ~ ˆ ~ Szenz eigenstates: Szenz |↑i = 2 |↑i, and Szenz |↑i = − 2 |↑i. The real parameters a, b > 0. 6

− The potential has two limiting behaviors. The behav- come “Yes, a Posner has formed.” Πno-coll. has projected dVLJ(x) the ions’ joint state. iors split where the derivative vanishes: dx = 0 ⇒ 1/6 Suppose, instead, that the ions have not emitted much x = 2a  =: x . At large distances x  x , the nega- b 0 0 heat. The environment has registered the outcome “No, tive term in Eq. (2) dominates. The derivative is positive, no Posner has formed.” 1 − Π− has projected the so V (x) attracts. Far-apart ions approach each other. no-coll. LJ ions’ joint state.5,6 At short distances x  x0, the positive term dominates. Let Sˆ1...6 denote the six phosphorus nuclei’s total spin The derivative is negative, so VLJ(x) repels. The ions cannot coincide at the same position. operator. We assume that Posner creation can be mod- eled as a two-stage process. First, the independent phos- Consider an ion approaching x = 0 from afar. VLJ(x) drops precipitously when the concavity changes from neg- phates tumble in the fluid. The spins rotate unitarily. 2 d VLJ(x) 26a 1/6 Second, the phosphates combine into a Posner via an ative to positive: 2 = 0 ⇒ x = . This dx 7b evolution that preserves (Sˆzlab )⊗6. The assumption fol- point forms a “lip” of the potential. The ions have more lows from Fisher’s claims that the spins barely deco- energy, separated, than they would have in a molecule. here [1]: The spins do not entangle with anything. At The ions slide down the potential well, releasing bind- worst, therefore, the spins rotate on the Bloch sphere dur- ing energy as heat. The heat disrupts the environment, ˆzlab 4 ing Posner creation. Most rotations fail to preserve S . which effectively measures the ions’ state. But Posner creation that involves rotations is mathemat- At the well’s bottom, the ions constitute a Posner ically equivalent to (i) rotations followed by (ii) (Sˆzlab )⊗6- molecule. The phosphorus nuclei’s quantum states have conserving Posner creation. The initial rotations can position representations (wave functions) that overlap be absorbed into the pre-Posner rotations. We there- significantly. The nuclei are indistinguishable [5]: No fore will say that Posner creation “essentially preserves” nuclear pair can be identified as red-painted or as blue- (Sˆzlab )⊗6. painted. The six phosphorus nuclei occupy a totally an- tisymmetric spin-and-spatial state. We will abbreviate “totally antisymmetric” as “antisymmetric.” III. ENCODED STATES AND THEIR CHANGING PHYSICAL REPRESENTATIONS II D. Formalizing the model for Posner-molecule creation Phosphorus nuclear spins cleanly encode QI before Posners form. The spins, Fisher conjectures, are decou- Let us model, with mathematical tools of QI, the en- pled from the nuclei’s positions [1]. Posner creation an- vironment’s measuring of the ions, the creation of a Pos- tisymmetrizes the spin-and-orbital state. The spins be- ner, and the antisymmetrization process. Let tPos denote come entangled with the positions, no longer encoding the scale of the time over which the ions slide down the QI cleanly. Lennard-Jones well from the lip, emit heat, jostle about, But, we posit, Posner creation maps each pre-Posner and settle into the Posner geometry. spin state to an antisymmetric Posner state determinis- The environment effectively measures the ions with tically. Posner creation preserves QI but changes how a frequency 1/tPos. We model the measurement with QI is encoded physically. Hence spin configurations can a projector-valued measure (PVM) [7]. Consider the ⊗6 Hilbert space (Hnuc) of the Posner’s six phosphorus − nuclei. An antisymmetric subspace Hno-coll. consists of the states available to the indistinguishable nuclei. (The 5 One might try to model the environment as measuring the ions states are detailed in Sec. III C.) The subscript stands continuously. This model is unfaithful: The environment would for “no-colliding-nuclei”: No two nuclei can inhabit the continuously project the ions onto states inaccessible to a Posner. No Posner could form, due to the quantum Zeno effect [33]. The same Posner-cube face. Posner-creation time tPos sets the measurement’s time scale. − − 6 Let Πno-coll. denote the projector onto Hno-coll.. The F&R suggest that, upon forming, a molecule is entangled with PVM has the form its environment [5, Eq. (7)]. Our PVM is consistent with F&R’s model, by the principle of deferred measurement [7]: Let S de- Π− , 1 − Π− . (3) note a general quantum system. A measurement of S consists of no-coll. no-coll. two steps: First, S is entangled with a memory M. Second, M is measured. Suppose that (i) the entanglement is maximal and (ii) Suppose that one length-(1/tPos) time interval has just the M measurement is projective. The M measurement projects passed. The environment has measured the ions. Sup- the system’s state. Suppose that S evolves after the M measure- pose that, during the interval, the ions have emitted con- ment. This entangling, M measurement, and evolution is equiv- siderable heat. The environment has registered the out- alent to the entangling, followed by the S evolution, followed by the M measurement. The M measurement can be deferred un- til after the evolution. Deferral fails to alter the measurement statistics. Let S denote the nuclei, and let M denote the environ- ment. The M measurement is deferred in F&R’s model, not in 4 That the environment measures the state via heat transfer was ours. The models are equivalent, by the deferred-measurement proposed in [1]. principle. 7 label a computational basis for the Posner Hilbert space, III B. Notation and quick review: Encodings e.g., ↑↑↑↑↑↑≡ 000000. This section is organized as follows. Section III A con- Imagine an agent Alice who wishes to send another cerns pre-molecule phosphorus nuclear spins. Section III agent, Bob, a message. A quantum message is a quantum B introduces notation. A map between (i) physical states state |ψLi ∈ HL. |ψLi is called the logical state. Let L of pre-Posner spins and (ii) logical states is formalized. Bcomp denote a preferred basis for the Hilbert space HL. Logical states are mapped to Posner states in Sec. III C. Operations are expressed in terms of this computational basis for the logical space. Alice must encode |ψLi in the state of a physical sys- tem. The agents would choose a code, a dictionary be- III A. Physical encoding of quantum information L in the phosphorus nuclei that will form a Posner tween the computational basis Bcomp for the logical space molecule and the computational basis Bcomp for the physical space. L Alice would decompose |ψLi in terms of Bcomp elements |jLi; replace each |jLi with a Bcomp element |ji; and pre- Consider six phosphates that approach each other, P pare the resultant physical state: |ψ i = cj|j i = soon to form (with Ca2+ ions) a Posner. We index the L j L P c |ji = |ψi. phosphorus nuclei as a = 1, 2,..., 6. Each nucleus has a j j H cannot be arbitrarily large, if the encoding is faith- spin DOF and an orbital DOF. Nucleus a occupies some L ful. A faithful encoding can be reversed, yielding the ρ ∈ D(H ). D(H) denotes the set of a nuc exact form of |ψ i. The six-qubit state |ψi can faith- density operators (trace-one linear operators) defined on L fully encode a |ψ i of ≤ 6 qubits, called logical qubits. the Hilbert space H. ρ may be pure (unentangled with L a The phosphorus nuclear spins—the physical DOFs that any external DOFs) or mixed (entangled with external encode the logical qubits—are called physical qubits. DOFs, e.g., another phosphorus nucleus’s spin). Suppose that |ψLi is a state of six logical qubits. We Tracing out the orbital DOF yields the reduced spin L spin 2 label the logical space’s computational basis as Bcomp = state: ρa := Trorb(ρa) ∈ D(C ). The magnetic spin 1 {|00 ... 0i, |00 ... 01i,..., |11 ... 1i}. A simple code from quantum number m = ± quantifies the spin’s z - L a 2 lab B to B has the form component. comp comp Shifting focus from chemistry to information theory, |m1, . . . , m6i ≡ |m1 . . . m6i , (4) we adopt QI notation: We usually omit hats from oper- for m , . . . , m = 0, 1. For example, all six physical ators, and we often omit factors of and 1 . We often 1 6 ~ 2 qubits’ pointing upward is equivalent to all six logical replace the spin operator’s α-component with the Pauli | i | i ˆα α ~ α α qubits’ pointing upward: 0,..., 0 = 0 ... 0 . α-operator, for α = x, y, z: S ≡ S = 2 σ ≡ σ . The σz eigenstates are often labeled as |0i := |↑i and |1i := |↓i. III C. Transformation of the encoding during Tracing out the spin DOF from ρa yields the reduced Posner-molecule creation orb orb orbital state: ρa := Trspin(ρa) ∈ D(Hnuc). We pa- Horb |xi rameterize nuc with the eigenstates of the position Consider six phosphates that join together, forming a operator, x. The coordinates are defined with respect to Posner. The phosphorus nuclei might begin with distin- the lab frame. {|xi} forms a continuous set. guishable DOFs (Sec. II C). The spins entangle with each The spin and/or orbital DOFs can store QI. But water other and with orbital DOFs [1, 5]. The QI |ψLi stored and other molecules buffet the phosphates. An indepen- in the spins “spills” into the orbital DOFs. dent phosphate’s position decoheres quickly. The spin, in But, we posit, Posner creation maps each pre-Posner contrast, is expected to remain coherent for long times spin state to an antisymmetric Posner state deterministi- (see Sec. IX and [1]). The spins encode protected QI. cally. The physical qubits change from spins to spin-and- The nuclear spins form six qubits. The qubits cor- orbital DOFs. The physical state’s form changes from spin ⊗6 12 12 0 − 0 respond to the Hilbert space (Hnuc ) = C , which |ψi ∈ C to some |ψ i ∈ H . The Posner state |ψ i 6 spin no-coll. has dimensionality 2 = 64. A useful basis for Hnuc encodes |ψLi faithfully. z consists of tensor products of σ eigenstates: Bcomp := Reparameterizing position will prove useful. We la- {|0, 0,..., 0i, |0, 0,..., 0, 1i,..., |1, 1,..., 1i}. The nota- beled by x a pre-Posner phosphorus nucleus’s position. tion |A, B, . . . , Ki ≡ |Ai⊗|Bi⊗...⊗|Ki. The set Bcomp A Posner’s phosphorus nuclei occupy the centers of cube is called the computational basis for the physical states. faces (Fig. 1). Let r = (r, ϕ, h) label a nucleus’s position Consider N hextuples of phosphates (N sets of six relative to the cube’s center. The cube’s size determines phosphates). The phosphorus nuclei correspond to a spin each nucleus’s distance r from the cube center. Hence space C6N . We suppose, without loss of generality, that we suppress the r: |ri ≡ |ϕ, hi. The angle variable is the 6N spins occupy a pure joint state |ψi. Each hextu- restricted to ϕ = φ, φ + 2π/3, φ + 4π/3 (Fig. 1b). The ple could contain three singlets, for example. Or a spin height variable is restricted to h = h± (Fig. 1a). in some hextuple A could form a singlet with a spin in Which states can one phosphorus nucleus occupy when some hextuple B. in a Posner? One might reason na¨ıvely as follows. The 8

spin basis {|0i, |1i} spans the nuclear-spin space Hnuc . The corresponding, anymore, to any particular nucleus. The orb basis {|ϕ, hi} spans the nuclear-position space Hnuc. nuclei delocalize across the cube-face centers. Hence a product basis spans the nuclear Hilbert space Let us mathematize this physics. The one-nucleus spin orb Hnuc = Hnuc ⊗ Hnuc: states (5) combine into the antisymmetric six-nucleus states

{|0; φ, h+i, |0; φ, h−i, |0; φ + 2π/3, h+i, |0; φ + 2π/3, h−i, 6! 6 1 X O πα √ (−1) |mπ (j), rπ (j)i (7) |0; φ + 4π/3, h+i, |0; φ + 4π/3, h−i, |1; φ, h+i, |1; φ, h−i, α α 6! α=1 j=1 |1; φ + 2π/3, h+i, |1; φ + 2π/3, h−i, |1; φ + 4π/3, h+i, := |(m1, r1)(m2, r2)(m3, r3); (m4, r4)(m5, r5)(m6, r6)i . |1; φ + 4π/3, h−i} . (5) Each term contains a tensor product of six one-nucleus kets. Each ket is labeled by one tuple (m , r ). We have condensed tensor products |mi ⊗ |ϕ, hi into πα(j) πα(j) |m; ϕ, hi. One might expect the phosphorus nucleus to No tuple equals any other tuple in the same term, by be able to occupy any state in (5). The hextuple of nuclei Pauli’s exclusion principle. Permuting one term’s six tu- ples yields another term, to within a minus sign. would be able to occupy a product state th πα denotes the α term’s permutation. The permuta- tion’s sign, (−1)πα = (−1)parity of permutation, equals the 7 |m1; ϕ1, h1i ⊗ ... ⊗ |m6; ϕ6, h6i . (6) term’s sign. The semicolon in Eq. (7) separates the h+ spins from the h− spins. (7) is equivalent to a Slater determinant [34]. The nuclei cannot occupy such a state, due to their If not for the Posner’s geometry, two tuples could con- indistinguishability. The nuclei are fermions. Hence Pos- tain the same position variables. r1 could equal r3, for ner formation antisymmetrizes the nuclei’s joint state. example, if m1 did not equal m3. But each cube face can We have assumed, in the spirit of [1], that Posner house only one phosphate. The phosphorus nuclei’s state − creation essentially preserves each phosphorus nucleus’s occupies the no-colliding-nuclei subspace Hno-coll. of the Szlab (Sec. II D). Hence the pre-Posner nuclei’s set {m} antisymmetric subspace. of spin quantum numbers equals the in-Posner nuclei’s Posner creation, we posit, projects the nuclei’s state − set. But Posner creation prevents any particular m from onto Hno-coll.. The projector has the form

X0 ED Π− := (m1,r1)(m2,r2)(m3,r3); (m1,r1)(m2,r2)(m3,r3); . (8) no-coll. (m4,r4)(m5,r5)(m6,r6) (m4,r4)(m5,r5)(m6,r6)

P0 The sum runs over values of (m1, . . . , m6). The value pre-Posner physical qubits. The right-hand side (RHS) Pos of (r1,... r6) = ((h+, φ),..., (h−, φ + 4π/3)) remains in- represents an element of the computational basis Bcomp variant throughout the terms.8 In every term, the first − for the space Hno-coll. of the in-Posner physical qubits. spin quantum number, m1, would correspond to the posi- Each pre-Posner state consists of a unique assignment tion r1 = (h+, φ). Different terms correspond to different of m-values to nuclei, a unique distribution of six fixed values m1 = 0, 1. m-values across six kets. Similarly, each Posner state − Projection by Πno-coll. applies the map consists of a unique assignment of m-values to positions, a unique distribution of six fixed m-values across six r- |m1i ⊗ ... ⊗ |m6i 7→ (9) values. Sixty-four pre-Posner Bcomp states exist. Hence |(m1, r1)(m2, r2)(m3, r3); (m4, r4)(m5, r5)(m6, r6)i. Pos 64 Bcomp basis elements must exist. A counting argument The left-hand side (LHS) represents an element of the in App. C confirms this conclusion. computational basis B for the space (H )⊗6 of the Let us combine the map (9) with the simple code (4). comp nuc The result is another simple code. This code maps be- Pos tween (i) elements of the computational basis Bcomp for the Posner space and (ii) elements of the computational 7 A permutation’s parity is defined as follows. Let π denote L 0 basis Bcomp for the logical space: the first term’s permutation. Consider beginning with π0 and swapping ket labels pairwise. Some minimal number n` of swaps |(m1, r1)(m2, r2)(m3, r3); (m4, r4)(m5, r5)(m6, r6)i yields permutation π`. The parity of n` is the parity of π`. 8 Each pre-Posner spin variable m pairs with one position r. What = |m1m2 . . . m6i . (10) determines which spin pairs with r1? Two factors: (i) the choice of coordinate system and (ii) the phosphates’ pre-Posner posi- Equation (10) shows how the QI, initially stored in pre- tions and momenta. See App. B for details. Posner spin states, is encoded faithfully in spin-and- 9

Pos orbital states. We will often replace the physical state’s values). C transforms the Bcomp elements (10) as label (the LHS) with the logical state’s label (the RHS), to streamline notation. C : |m1m2m3m4m5m6i ≡ |(m1, r1)(m2, r2)(m3, r3); (m4, r4)(m5, r5)(m6, r6)i

7→ |(m3, r1)(m1, r2)(m2, r3); (m6, r4)(m4, r5)(m5, r6)i ≡ |m m m m m m i . (11) IV. CHARGE-PROTECTED ENCODINGS FOR 3 1 2 6 4 5 QUANTUM INFORMATION STORED IN The BPos elements (10) are not C eigenstates. POSNER MOLECULES comp But C eigenstates can be constructed. We adopt F&R’s notation for the eigenvalues,

The computational-basis elements (10) are states of τ i2π/3 spin-and-orbital DOFs. The Posner’s dynamics conserve ω , wherein ω := e and (12) the spins’ states for long times, Fisher hypothesizes [1]. τ = 0, 1, 2 or, equivalently, τ = 0, ±1 . The dynamics might not conserve the orbital DOFs’ states. Hence the dynamics may not not conserve the F&R call τ a three-level “pseudospin.” We call τ, in- states (10). stead, the eigenvalue of the observable GC that generates C.9 The general form of a C eigenstate appears in [5]. But we posit, guided by [1, 5], that the Posner’s dy- F&R use second quantization. G namics conserve certain charges: (i) the generator C We translate into QI. We also extend [5] by characteriz- of a permutation operator C (Sec. IV A) and (ii) the zin ing the eigenspaces of C and by identifying a useful basis total spin operator’s zin-component, S1...6 (Sec. IV B). for each eigenspace (Sec. IV C). The τ = 0 eigenspace Eigenstates shared by these charges (Sec. IV C) may be has degeneracy 24; the τ = 1 eigenspace, degeneracy 20; conserved. and the τ = −1 eigenspace, degeneracy 20. The τ = 0 The dynamics likely will not map an eigenstate |ψi, eigenspace will play an important role in Posner resource (ψ) associated with different eigenvalues τψ and m1...6 of the states for universal quantum computation (Sec. VII). The zin charges, into an eigenstate |φi associated with eigen- other charge assumed to be conserved, S1...6, shares this (φ) eigenbasis. values τφ and m1...6. These eigenstates may serve as long-lived codewords. Charge preservation help “pro- tect” such codes. IV B. Charge 2: The total spin operator Szin We identify a quantum error-detecting code partially 1...6 protected by C. A repetition code is partially protected by Szin . Section IV D introduces these codes. Fisher conjectures that phosphorus nuclear spins in 1...6 Posners have long coherence times [1]. We interpret this conjecture as meaning that the Posner’s dynamics con- zin L6 zin serves S1...6 = a=1 Sa for long times. The total mag- P6 netic spin quantum number, m ... = mj, remains IV A. Charge 1: The generator GC of the 1 6 j=1 permutation operator C constant. This interpretation is supported by two arguments in App. D. Possible interactions between a Posner’s phos- Imagine transforming a Posner, geometrically, as fol- zin phorus nuclear spins conserve S1...6. So, we argue, do lows. The rotation is clockwise, about the symmetry axis collisions with other molecules. zˆin, through an angle 2π/3. Only the molecule’s archi- tecture rotates: its atoms, its internal coordinate system, its geometry. Imagine that the spins (represented by the 9 ma’s) could remain untouched. The transformed Posner A pseudospin is a physical DOF that transforms according to a would look identical to the original Posner [2–4, 9]. certain rule. τ is, rather, the eigenvalue of an observable. Sup- pose that τ were a three-level quantum pseudospin. τ would oc- This symmetry characterizes the molecule’s geome- cupy a quantum state in some three-dimensional effective Hilbert try, revealed by quantum-chemistry calculations [2–4, 9]. space Hpseudo. No such space can be associated uniquely with a Posner, to our knowledge. Rather, the Posner Hilbert space Hence our introduction of the symmetry in terms of a ro- − − Hno-coll. has dimensionality 64. Hno-coll. equals a direct sum of tation of the architecture. But QI is expressed more nat- − the three GC eigenspaces: Hno-coll. = Hτ=0 ⊕ Hτ=1 ⊕ Hτ=2. urally in terms of spins. We therefore recast the transfor- Each subspace is degenerate. Hence no subspace can serve mation, as a counterclockwise cyclic permutation of the as one element in a basis for any Hpseudo. One could con- spins (Fig. 1b). This spin transformation may be viewed jure up a Hpseudo by choosing one state |τ=0i ∈ Hτ=0, one as active; the earlier geometric transformation, as pas- |τ=1i ∈ Hτ=1, and one |τ=2i ∈ Hτ=2; then constructing sive. Hpseudo = span {|τ=0i, |τ=1i, |τ=2i}. We do so in Sections IV D 1 and VI. But the choice of |τ=0i is nonunique, as is the choice Let the operator C represent this spin permutation. C of |τ=1i, as is the choice of |τ=2i. Hence no unique three-level cyclically permutes the h+ spins [the first three m-values Hilbert space corresponds to a Posner, to our knowledge. Hence τ appears not to label a unique quantum pseudospin. in (10)] separately from the h− spins (the final three m- 10

− We decompose Hno-coll. into composite-spin subspaces The trios function logically as independent units. Such in App. E. That appendix also reviews addition of quan- trios can be used to prepare universal quantum- tum angular momentum. computation resource states (Sec. VII B). Hence the spin- operator trios in Eq. (13). Each qubit trio corresponds to a Hilbert space C6. Let IV C. Eigenbasis shared by the conserved charges zin 2 us focus on qubits 1-3, for concreteness. C, S123, and S123 share the basis in Table I. Each basis element is symmet- We introduced the computational basis Bcomp for ric with respect to cyclic permutations of the three logical − Hno-coll. in Eq. (7). Most Bcomp elements transform non- qubits. trivially under C [Eq. (11)]. The Posner dynamics con- Tensoring together two one-triangle states yields serve C. So, too, would the dynamics ideally conserve a state of a Posner’s phosphorus nuclear spins: quantum codewords. We therefore seek a useful C eigen- |000i|000i, |000i|W i,..., |ω2i|ω2i. Sixty-four such states basis from which to construct QEC codes. exist. We classify them with quantum numbers in The C eigenspaces have degeneracies. Which basis App. F. should we choose for each eigenspace? A basis shared We have pinpointed an eigenbasis shared by the con- zin with S1...6, the other conserved charge. served charges. The Posner dynamics are expected not zin Yet C and S1...6 do not form a complete set of com- to map states in one charge sector to states in another. muting observables (CSCO) [35]. Many eigenbases of C Hence different-sector states suggest themselves as quan- zin are eigenbases of S1...6. Another operator is needed to tum codewords. We present partially charge-protected break the degeneracy, to complete the CSCO. We choose QECD codes next. the spin-squared sum

2 2 2 2 S123 + S456 ≡ (S1 + S2 + S3) + (S4 + S5 + S6) (13) IV D. Quantum error-detecting and -correcting 3 codes accessible to Posner molecules X ⊗(a−1) ⊗(3−a) ≡ 1 ⊗ Sa ⊗ 1 a=1 We exhibit two codes formed from states accessible to 6 Posners. Each codeword is an eigenstate of a conserved X 1⊗(a−4) 1⊗(6−a) zin + ⊗ Sa ⊗ . (14) charge, C or S1...6. Each code’s codewords correspond a=4 to distinct eigenvalues of the charge. Hence the Posner dynamics likely do not map any codeword into any other. Geometry and measurement-based quantum computa- Section IV D 1 introduces a quantum error-detecting tion (Sec. VII B) motivate the choice of S2 + S2 :A 123 456 code. One Posner, we show, can encode one logical qutrit. Posner contains two triangles of spins (Fig. 1a). The po- The code detects one arbitrary physical-qubit error. Sec- sitions in the h triangle are labeled r , r , and r . Hence + 1 2 3 tion IV D 2 shows how to implement a repetition code the first three tuples in Eq. (10) correspond to the h tri- + with Posner states. The code corrects two bit flips. angle. Hence the magnetic spin quantum numbers m , 1 More Posner codes, we expect, await discovery. Op- m , and m may be viewed as occupying the h triangle. 2 3 + portunities are detailed in Sec. X. These spins’ joint state is equivalent to a three-qubit log- We have already discussed an encoding of logical states ical state, |m m m i. An analogous argument concerns 1 2 3 in physical systems (Sec. III B). Earlier, the logical h . Hence the antisymmetric state (10) is equivalent to − Hilbert space H shared the physical Hilbert space’s di- a product of two three-logical-qubit states:10 L mensionality, 64. Section III B concerned a bijective, in- jective map between the spaces. QECD encodes a small |(m1, r1)(m2, r2)(m3, r3); (m4, r4)(m5, r5)(m6, r6)i logical space in a larger physical space. Notation will re- ≡ |m m m i|m m m i . (21) 1 2 3 4 5 6 flect the distinction between Sec. III B and QECD: Script subscripts L (as in HL) will replace the Roman L (as in HL). QECD is reviewed in App. A 3.

10 In greater detail,

|(m1, r1)(m2, r2)(m3, r3); (m4, r4)(m5, r5)(m6, r6)i (15) IV D 1. Qutrit error-detecting code formed from

≡ |m1m2m3m4m5m6i (16) Posner-molecule states

≡ |m1i|m2i|m3i|m4i|m5i|m6i (17)

= (|m1i|m2i|m3i)(|m4i|m5i|m6i) (18) One Posner, we show, can encode one logical qutrit.

≡ |m1, m2, m3i|m4, m5, m6i (19) The code detects arbitrary single-physical-qubit errors.

≡ |m1m2m3i|m4m5m6i . (20) The physical qubits are the spin-and-orbital DOFs of Sec. III C. Equation (16) is equivalent to Eq. (10). Equation (17) is equiva- lent to Eq. (4). Equation (18) follows from the tensor product’s The code has the form associativity. Equation (19) consists of a rewriting with new no- qutrit tation. Equation (20) is analogous to Eq. (4). HL = span {|0Li, |1Li, |2Li} , (22) 11

State Decomposition S123 m123 τ |000i |000i 3/2 3/2 0 √1 (|100i + |010i + |001i) |W i 3 3/2 1/2 0 ¯ √1 (|011i + |101i + |110i) |W i 3 3/2 −1/2 0 |111i |111i 3/2 −3/2 0

√1 (|100i+ω2|010i+ω|001i) |ωi 3 1/2 1/2 1 √1 (|011i+ω2|101i+ω|110i) |ω¯i 3 1/2 −1/2 1 2 √1 (|100i+ω|010i+ω2|001i) |ω i 3 1/2 1/2 2 2 √1 (|011i+ω|101i+ω2|110i) |ω i 3 1/2 −1/2 2

TABLE I: Symmetric basis for a trio of qubits: A Posner molecule consists of two triangles of spins (Fig. 1). The triangles encode quantum information independently, in accordance with Eq. (21). Each triangle therefore functions as a trio 6 of logical qubits. The three physical qubits correspond to an eight-dimensional Hilbert space, C . A useful basis is an zin eigenbasis shared by the conserved charges, C (a permutation operator) and S123 (the z-component, relative to the internal 2 zˆin-axis, of the total spin). These operators share many basis. The eigenbasis shared also by S123 proves useful in the preparation of universal quantum-computation resource states (Sec. VII). τ describes, here, how a triangle transforms under the permutation represented by C.

wherein zeroes, counting the ones, and following majority rule. More physical bits end as 0s than as 1s in our example. 1 |0Li = √ (|W i|W¯ i − |W¯ i|W i) , (23) A logical zero, the receiver infers, was likely sent. 2 The repetition code can be translated into quantum rep 1 2 2 2 2 states. For example, let H = {|0Li, |1Li} , wherein |1Li = √ (|ω i|ω i − |ω i|ω i) , and (24) L 2 |0Li = |000000i and |1Li = |111111i . This code cor- 1 rects two σx errors. But each codeword is unentangled.11 |2Li = √ (|ωi|ω¯i − |ω¯i|ωi) . (25) Hrep 2 Hence L fails to satisfy the off-diagonal error-detection criterion, Each logical state |j i occupies the τ = j subspace. L hj |σασβ|k i = 0 ∀j =6 k , (29) The codewords satisfy the two quantum error- L L detection criteria [8, 36–40]. First, the states are locally whenever α =6 x and/or β =6 x. indistinguishable:

x y hjL|σ |jLi = hjL|σ |jLi = 0 and (26) V. THE MODEL OF POSNER QUANTUM z 1 COMPUTATION hjL|σ |jLi = (27) 12 Fisher has conjectured that certain chemical processes for all j. That is, the codewords satisfy the diagonal occur in biofluids [1]. We abstract away the chemistry, criterion. Second, the codewords satisfy the off-diagonal identifying the computations that the processes effect. criterion, We call the computations Posner operations.12 The op- Posner hj |σα|k i = 0 ∀j =6 k , ∀α = x, y, z , (28) erations form a model of quantum computation, L L quantum computation. by direct calculation. The model’s operations, we will show, can be used (i) to teleport QI incoherently and (ii) to prepare, effi- ciently, universal resource states for measurement-based IV D 2. Repetition code formed from Posner-molecule quantum computation (Sections VII-VII). Whether the states model’s operations can realize universal quantum com- putation remains an open question (Sec. IX). The repetition code originated in classical error correc- tion [41]. Each logical bit is cloned until n copies exist: 0 7→ 00 ... 0, and 1 7→ 11 ... 1. Suppose that errors flip 11 More precisely, the element |000000i of the computational basis | {z } | {z } n n for the logical space (Sec. III C) is unentangled. The spin-and- under half the bits. For example, 000000 may transform orbital state represented by |000000i [by Eq. (10)] is entangled. 12 into 011000. One decodes the bit string by counting the We occasionally call the processes “Posner operations.” 12

FIG. 3: Circuit-diagram element that represents singlet-state preparation (operation 1). FIG. 4: Circuit-diagram element that represents a rotation of a qubit not in any Posner (operation 2): nˆ denotes the axis rotated about. θ denotes the angle Posner quantum computation is defined in Sec. V A. rotated through. The model is analyzed in Sec. V B. We discuss the model’s ability to entangle, the requirements for run- ning arbitrary-depth quantum circuits, and the control required to perform QI-processing tasks. Fisher’s narra- tive [1] is also cast as a quantum circuit.

V A. Definition of Posner quantum computation

Terminological notes are in order. When discussing FIG. 5: Circuit-diagram element that represents physical processes, we discuss phosphorus nuclear spins, hextuple formation (operation 3): Straight lines spin-and-orbital DOFs, and Posners. When discussing represent qubits not in hextuples. Each wavy line represents logical DOFs, we discuss qubits. A circuit-diagram ele- a qubit in a hextuple (that is not paired with any other ment represents each operation (Figures 3-12): hextuple as a result of operation 5). As the lines’ labels show, the circuit element is defined as preserving the qubits’ 1. Singlet-state preparation (Fig. 3): Arbitrarily ordering. many singlets |Ψ−i can be prepared. Singlets are pre- pared when an enzyme hydrolyzes diphosphates into entangled phosphate pairs (Sec. II C).13 Suppose, instead, that the diphosphate leaves the en- These singlets are prepared differently than in conven- zyme uncleaved. The second possible measurement tional quantum circuits. Conventionally, one prepares outcome has obtained. The diphosphate cannot form two qubits in the state |0i⊗2; performs a Hadamard14 a Posner molecule with other ions. Hence the diphos- on the first qubit; and performs a CNOT,15 controlling phate cannot participate in quantum cognition. Hence on the first qubit: CNOT(H ⊗ 1)|00i = |Ψ−i. the diphosphate plays no role in Posner quantum com- putation. Hence the PVM’s 1 − |Ψ−ihΨ−| outcome In contrast, Fisher posits that enzymes prepare plays no role. Any diphosphate that remains un- singlets by projective measurements [1]. We for- cleaved “is discarded,” in QI language. The |Ψ−ihΨ−| malize Fisher’s statement as follows. A diphos- outcome is classically postselected on. Classical post- phate’s phosphorus nuclear spins occupy some state selection provides no superquantum computational ρ . The diphosphate enters a pyrophos- diphos power; see footnote 17. phatase enzyme. The enzyme measures the PVM {|Ψ−ihΨ−|, 1 − |Ψ−ihΨ−|}. In summary, Posner operations include the prepara- | −i Suppose that the diphosphate separates into two dis- tion of Ψ . Quantum-cognition systems prepare | −i connected phosphates. The spins’ state has been pro- Ψ by measuring a PVM nondestructively, then jected with |Ψ−ihΨ−|. postselecting classically on the “yes” outcome. “No”- outcome ions do not participate in later chemical events of interest. 13 Biofluids might prepare phosphorus nuclear spins in nonsinglet 2. Rotations of independent qubits (Fig. 4): Any states. For example, one phosphate might detach from ATP, qubit can rotate through any angle θ about any axis leaving adenosine diphosphate (ADP). Identifying the phos- , via a unitary . phate’s quantum state would require physical modeling outside nˆ Unˆ (θ) nˆ is defined relative to the this paper’s scope. Therefore, we suppose that qubits can be lab frame. Qubits rotate as phosphates tumble in the prepared only in singlets. fluid. 14 The Hadamard gate H transforms one qubit [7]. In terms of 1 Pauli operators, H = √ (σx + σz). The gate has a geometric Hextuple formation Qubits can group to- 2 3. (Fig. 5): interpretation expressed in terms of the Bloch sphere: The state gether in hextuples (groups of six). Hextuple formation rotates through 180◦ about the axis √1 (ˆx +z ˆ). 2 evolves the logical qubits trivially, under the operator 15 The CNOT, or controlled-not, gate transforms two qubits [7]. 1. But hextuple formation associates the qubits with a One qubit is called the control, and one is called the target. If the geometry and with an observable G . control occupies the state |0i, the CNOT preserves the target’s C state. If the first qubit occupies |1i, the target evolves under σx. Logical qubits form hextuples as ions bind together, 1 The CNOT has the form |0ih0| ⊗ + |1ih1| ⊗ σx. forming Posners. A logical qubit can occupy, at most, 13

FIG. 6: Circuit-diagram element that represents hextuple-coordinated single-qubit rotations (operation 4). nˆ denotes the axis rotated about. θ denotes FIG. 7: Circuit-diagram element that represents a the angle rotated through. Posner-binding measurement (operation 5) that yields a positive outcome: Wavy lines represent qubits in hextuples that are not in a dodectuple (a pair of one hextuple. Section III C explains why hextuple hextuples). Coils represent qubits in a dodectuple. formation fails to change logical qubits’ states: The spins’ state changes, suggesting that the logical qubits’ state changes. But the logical information’s physical encoding changes, too. Hextuple creation impacts the logical system in three ways: (i) Each hextuple has an observable GC . (ii) Hextuple creation induces a geometry that influences operation 4. (iii) The six logical qubits’ Hilbert space 12 − transforms from C to the isomorphic Hno-coll.. Let us detail these three effects.

First, creating a hextuple creates an observable GC (Sec. IV A). GC has eigenvalues τ = 0, 1, 2 (equiva- lently, τ = 0, ±1). τ impacts operation 5. Second, hextuple creation induces a geometry. Each logical qubit is assigned to a cube face, in accordance with Sec. III C. The six qubits can be distributed across the six faces in any of 6! ways. Physically, different assignments follow from different pre-Posner FIG. 8: Circuit-diagram element that represents a orbital states (App. B). The six qubits form two tri- Posner-binding measurement (operation 5) that angles, called trios below, in accordance with Fig. 1. yields a negative outcome. This geometry limits the single-qubit unitaries that can evolve the six qubits (operation 4). The geometry also influences our construction of universal quantum- tion 3. Whether the hextuples’ GC eigenvalues sum to computation resource states (Sec. VII). zero can be measured nondestructively: τA + τB = 0. Third, hextuple creation changes the system’s Hilbert 2 ⊗6 − − First, we discuss the measurement’s physical manifes- space from (C ) to Hno-coll. (Sec. III C). Hno-coll. is 2 ⊗6 tation. Then, we mathematize the operation with a isomorphic to (C ) , as the map (10) is injective and PVM. bijective. Hence we will keep referring to the logical space as (C2)⊗6. The measurement manifests in the binding, or fail- ure to bind, of two Posners. Fisher conjectures as 4. Hextuple-coordinated single-qubit rotations follows [1], supported by quantum-chemistry calcula- (Fig. 6): The qubits in a hextuple can undergo identi- tions [9]: Two Posners, A and B, can bind together. cal arbitrary single-qubit rotations simultaneously, via They bind upon approaching each other “head-on”: ⊗6 [Unˆ (θ)] . The qubits rotate as the Posner tumbles Their directed symmetry axes (Fig. 1a) point oppo- about an axisn ˆ relative to the lab frame. sitely each other. 5. Posner-binding measurement (Figures 7 and 8): Such Posners bind, Fisher conjectures [1], when and Let A and B denote two hextuples formed via opera- only when not rotating relative to each other. Relative 14

rotation would hinder A in “grabbing onto” B. Hence if A and B approach head-on, whether they bind de- pends entirely on whether τA +τB = 0. The molecules’ bound-together-or-not status serves as a classical mea- surement record. So does the environment, as in Pos- ner creation (Sec. II): Posner binding releases about 1 eV of heat [1, 9]. Let us formalize the measurement, using the mathe- − ⊗2 matics of QI. We define a projector on Hno-coll. :

ΠAB := (ΠτA=0 ⊗ ΠτB =0) + (ΠτA=±1 ⊗ ΠτB =∓1) . (30) The PVM

{ΠAB, 1 − ΠAB} (31) can be measured. Suppose that the first outcome ob- tains (that the Posners bind). The two-Posner state ρ updates as Π ρ Π ρ 7→ AB AB . (32) FIG. 9: Circuit-diagram element that represents Tr (ΠAB ρ ΠAB) dodectuple-coordinated single-qubit unitaries (operation 6a). nˆ denotes the axis rotated about. θ The twelve qubits for a dodectuple. Suppose, instead, denotes the angle rotated through. that the second outcome obtains (that the Posners fail to bind). The joint state updates as ρ − {Π , ρ} + Π ρ Π ρ 7→ AB AB AB . (33) 1 − Tr (ΠAB ρ) The anticommutator of operators O and O0 is denoted by {O,O0}. 6. Dodectuple operations: Suppose that hextuples A and B have been measured with the PVM (31). Sup- pose that outcome ΠAB has obtained. The twelve log- ical qubits can undergo operation 6a; or 6b; or 6a, followed by 6b. (a) Dodectuple-coordinated single-qubit uni- taries (Fig. 9): The two hextuple’s qubits can undergo identical arbitrary single-qubit rotations: ⊗12 [Unˆ (θ)] . The qubits rotate as the Posners tum- ble in the fluid. FIG. 10: Circuit-diagram element that represents (b) Dodectuple → 2 hextuples (Fig. 10): A dodec- the separation of a dodectuple into two hextuples tuple can separate into independent hextuples, the (operation 6b): Each coil represents a qubit in a hextuples that joined together. The Posners can dodectuple. Each wavy line represents a qubit in a hextuple drift apart. The hextuples can return to undergo- that is not in a dodectuple. ing the operation 4 (and can, we posit, undergo operation 6). (c) Dodectuple break-up (Fig. 11): The hextuples Fisher’s narrative allows for, though does not require, can break down into their constituents: The qubits another operation: can cease to correspond to meaningful geometries Hextuple break-up One hextuple can or to observables GC . The qubits thereafter be- 6. (Fig. 12): have independently. They can, again, undergo op- break down into its constituents: The qubits can cease erations 2-3. to correspond to geometries or to observables GC . The hextuples break down as the Posners hy- Different regions of the body have different pH’s and drolyze. Fisher conjectures that bound-together different magnesium-ion (Mg2+) concentrations. A Posners hydrolyze more often than separated Pos- Posner can migrate to a region packed with H+ and/or 2+ 3− 2+ ners [1], as discussed in the introduction. with Mg . These ions can bind to PO4 , as Ca 15

is compared with a Bell measurement, and applied in QI-processing tasks, in Sec. VI.

V B 1. Entanglement generation

Entanglement enables quantum computers to solve cer- tain problems quickly.16 Two Posner operations create entanglement: Bell-pair creation (operation 1) and the Posner-binding measurement (operation 5). Bell pairs serve as units of entanglement in QI [7]. We present two implications of Bell-pair creation for Posners. First, Bell-pair creation (operation 1), with the Posners’ geometry, can efficiently prepare a state that fuels univer- FIG. 11: Circuit-diagram element that represents sal measurement-based quantum computation (Sec. VII). dodectuple break-up (operation 6c): Each coil Second, distributing Bell pairs across Posners can affect represents a qubit in a dodectuple. Each straight line their binding probabilities (Sec. VIII). represents a qubit not grouped with any other qubits. The role played by Bell pairs in QI processing is well- known. Less obvious is how much, and which kinds of, entanglement ΠAB creates and destroys. We charac- terize this entanglement in two ways (Sec. VI A). The PVM (31), we show, transforms a subspace as a coarse- grained Bell measurement. Bell measurements facilitate [10]. The PVM (31) facilitates in- coherent teleportation: A state’s weights are teleported; FIG. 12: Circuit-diagram element that represents the coherences are not. The dephasing comes from the hextuple break-up (operation 6): Each coil represents a PVM’s simulation of a coarse-grained Bell measurement. qubit in a dodectuple. Each straight line represents a qubit not grouped with any other qubits. One might expect Posner binding to render Posner quantum computation universal: Conventional wisdom says, nearly any entangling gate, plus all single-qubit uni- can. The higher the H+ and Mg2+ concentrations, taries, form a universal gate set [44–48]. Posner opera- the more H+ and Mg2+ ions dislodge Posners’ Ca2+ tions include entangling gates and all single-qubit uni- ions [6]. The dislodging hydrolyzes the molecules. taries. But the conventional wisdom appears inapplicable to Operation 6 can be used to prepare Posners, efficiently, Posner operations, for three reasons. First, conventional- in resource states that can power universal measurement- wisdom gates evolve the system unitarily. The Posner- based quantum computation (Sec. VII). binding measurement (31) does not. (Hence our shift to measurement-based quantum computation in Sec. VII.) Second, many universality proofs decompose a desired V B. Analysis of Posner quantum computation entangling gate into implementable gates. The Posner- binding measurement seems unlikely to decompose. We have dissected Fisher’s narrative into physical pro- Third, conventional-wisdom entangling gates are de- cesses, and we have abstracted out the computations that fined in terms of qubits’ states. The Posner-binding mea- the processes effect. Fisher’s narrative [1] can now be cast surement is defined in terms of τ. τ is an eigenvalue of an as a quantum circuit. The circuit appears in Fig. 13. observable GC of a hextuple of qubits. One must deduce Four features of Posner operations merit analysis. Two how the measurement transforms any given qubit. Does operations entangle logical qubits. The entanglement this indirect entangler of qubit states, with all single- generated is discussed in Sec. V B 1. Section V B 2 con- qubit rotations (operation 2), form a universal set? The tains a criterion for realizing nonconstant-depth circuits: answer merits further study. Posners must bind, hydrolyze, and rebind. Section V B 3 concerns control: To perform the QI-processing tasks in- troduced in Sections VI-VII, one might need fine control over Posners. Biofluids might not exert such control. But assuming control facilitates first-step QI analyses. One 16 More precisely, contextuality underlies quantum speedups [42, operation merits its own section: The measurement (31) 43]. 16

FIG. 13: Circuit representation of Fisher’s quantum-cognition narrative: Fisher conjectures that certain chemical processes occur, in a certain sequence, in the body [1]. The sequence is reviewed in this paper’s introduction. We abstracted out the computations effected by the chemical processes, in Sec. V A. The abstraction enables us to recast Fisher’s narrative as a quantum circuit. Time progresses from left to right in the figure (from the bottom to the top of the page). The sets of six qubits are labeled a, b, a0, and b0, as in [1, p. 5, Fig. 3]. The circuit elements are defined in Sec. V A. 17

V B 2. Circuit depth measurements.17 But assuming perfect control can facilitate QI- Consider attempting to implement universal quantum theoretic analyses. Many QI protocols are phrased in the computation with Posner operations. One must be able language of “agents.” One imagines intelligent agents, to run finite-depth circuits, to perform many computa- Alice and Bob, who wish to process QI. One specifies tions sequentially. One must be able to entangle qubits, and analyzes protocols in terms of the agents’ intents and to rotate qubits independently, late in the computa- and actions. Alice and Bob are often assumed to per- tion. form certain operations with perfect control. Examples of such “allowed operations” include local operations and Let us focus on the biological-circuit components that classical communications [51]. follow state preparation (on the operations beyond 1). Control partitions (i) what can be achieved in princi- Qubits can undergo entangling operations only when ple from (ii) what can be achieved easily with today’s qubits are in Posners (via operation 5). Qubits can rotate knowledge and techniques. Item (ii) shifts with our un- independently only when phosphates are outside of Pos- derstanding and technology. Item (i) is permanent and ners (via operation 2). Hence universal quantum com- is the focus of much QI theory. putation would require (i) Posner binding and hydrolysis A few decades ago, for example, experimentalists had (operations 5 and 6c) and/or (ii) one-Posner hydrolysis trouble performing CNOT gates. Many groups have mas- (operation 6). tered the gate by now. These groups implement protocols Suppose that all the Posners fail to hydrolyze. Pos- devised before CNOTs appeared practical. ner operations can realize only depth-4 circuits. Realiz- Similarly, precise phosphate rotations appear imprac- ing constant-depth circuits does not suffice for realizing tical. But some precise-rotation mechanism could be dis- universal quantum computation [49]. Hence Posner dis- covered. Also, by assuming perfect control, we derive a integration is necessary for realizing universal quantum limit on what Posners can achieve without perfect con- computation. Whether Posner operations suffice remains trol. We ascertain what QI processing is possible in prin- an open question. ciple.

V B 3. Control required to perform VI. THE POSNER-BINDING MEASUREMENT quantum-information-processing tasks with Posner molecules AND APPLICATIONS THEREOF TO QUANTUM INFORMATION PROCESSING In Sec. VI–VIII, we concatenate Posner operations to The measurement (31) entangles two Posners’ states. form QI-processing protocols. Implementing the proto- 1 cols may require fine control over the chemical processes Yet the measurement projectors, ΠAB and − ΠAB, en- that effect the computations. The body might seem un- tangle states in different ways. How much either pro- likely to realize fine control. We illustrate with two ex- jector entangles is not obvious. Neither is the PVM’s amples. Then, we justify the assumption of fine control. potential for processing QI. This section sheds light on these unknowns. We com- Consider, as a first example, running an arbitrary pare the PVM to a Bell measurement, a standard QI quantum circuit. Arbitrary qubits must rotate through operation (Sec. VI A). The next two sections detail ap- arbitrary angles θ, about arbitrary axesn ˆ, arbitrarily pre- plications of the PVM: The PVM facilitates incoherent cisely. In the quantum-cognition setting, logical qubits teleportation (Sec. VI B). Also, the PVM can be used to rotate as phosphates tumble (via operation 2). Phos- project Posners onto their τ = 0 eigenspaces (Sec. VI C). phates tumbles upon colliding with other particles. Fluid particles collide randomly, trading angular momentum randomly. Random collisions appear unlikely to gener- 17 ate the precise rotations required for a given circuit. If two Posners bind, then τA + τB = 0; the inference is justified. But binding does not constitute merely a measurement. Binding The τA +τB = 0 measurement (operation 5) provides a constitutes a measurement followed by classical postprocessing. second example. Consider Posners A and B approaching By classical postprocessing, we mean the following. Consider per- each other with momenta pA and pB. Consider observ- forming some protocol in each of several trials. Let the protocol ing whether the Posners bind. One might wish to infer, involve a measurement. Consider the data collected throughout trials. Consider discarding some of the data, keeping only the from the binding or lack thereof, whether τA + τB = 0. data collected during the trials in which the measurement yielded But the inference is justified only if pA and pB were such some outcome x. One has classically postselected on x. If two that whether the Posners would bind depended only on Posners bind, then (i) whether τA + τB = 0 is measured and (ii) whether τA + τB vanished. the “yes” outcome is classically postselected on. If two Posners Suppose that the Posners approached not head-on, but bind, step (i) alone is not implemented; a measurement alone is not performed. at a slight angle:p ˆA =6 −pˆB. The Posners would fail Classical postprocessing differs from the postselection in, to bind. But one could not infer that τA + τB =6 0. e.g., [50]. The latter postselection affords computational power Only finely tuned two-Posner encounters reflect whether unlikely to grace quantum systems. In contrast, classical post- processing happens in today’s laboratories. τA + τB = 0. Only finely tuned encounters constitute 18

VI A. Comparison of the Posner-binding Relabeling and direct substitution show that measurement with a Bell measurement + + − − 1 − ΠAB = |Φ ihΦ | + |Φ ihΦ | . (44) First, we review Bell states and measurements [7]. A Bell measurement prepares an entangled state of two Let us quantify the coarse-graining in Proposition 1. qubits. Four maximally entangled states span the two- Let |χi denote an arbitrary two-qubit state. Consider qubit Hilbert space, 4. The orthonormal Bell basis is C measuring |χi in the Bell basis. One of four possi- ble outcomes obtains. The outcome can be encoded in + 1 {|Φ i := √ (|00i + |11i) (34) log (4) = 2 bits. You could encode, in one bit, whether a 2 2 Φ outcome or a Ψ outcome obtained. You could encode, 1 |Φ−i := √ (|00i − |11i) (35) in the second bit, whether a + outcome or a − outcome 2 obtained. 1 |Ψ+i := √ (|01i + |10i) (36) Imagine knowing the first bit’s value and forgetting 2 the second bit’s. The state most reasonably attributable + + − − − 1 to the system would be (|Φ ihΦ | + |Φ ihΦ |)|χi or |Ψ i := √ (|01i − |10i)} . (37) | +ih +| | −ih −| | i 2 ( Ψ Ψ + Ψ Ψ ) χ , depending on the first bit. This state would be the state most reasonably at- A Bell measurement is represented by the PVM tributable to the system if, instead, (40) were measured and the outcome were known. + + − − + + − − {|Φ ihΦ |, |Φ ihΦ |, |Ψ ihΨ |, |Ψ ihΨ |} . (38) The information in the measurement outcome can be quantified differently. Appendix G contains details. Many QI protocols involve Bell measurements. Exam- ples include quantum teleportation [10], superdense cod- ing (the effective transmission of two bits via the direct VI B. Application 1 of binding Posner molecules: transmission of just one bit, with help from entangle- Incoherent teleportation ment) [11], and teleportation-based quantum computa- tion [52–55]. Quantum teleportation transmits a state |ψi from one Posner binding simulates a coarse-grained Bell mea- system to another [10]. Consider agents Alice and Bob surement. The Bell-state projectors (34) are defined on 2 − who live in the same town. Suppose that Bob moves C . In contrast, the Posner Hilbert space Hno-coll. is iso- 6 away, across the world. morphic to C . We therefore define an effective qubit Let Alice hold a qubit A that occupies a state |ψi = subspace. Let |1τ i denote an arbitrary τ = 1 eigenstate c0|0i + c1|1i. Alice may wish to send Bob |ψi. Mailing of C; and |2τ i, an arbitrary τ = 2 eigenstate. |1τ i and A would damage the state. Alice should not measure A, |2τ i serve analogously to |0i and |1i in span {|1τ i, |2τ i}. call Bob on the telephone, and tell him the outcome. Bob Proposition 1. Let A and B denote two Posners. would receive too little information to reconstruct |ψi in The measurement (31) transforms the effective two-qubit his lab. space Suppose that, before Bob moved away, he and Alice created a Bell state, e.g., |Ψ−i. Let B and C denote the span {|1τ , 1τ i, |1τ , 2τ i, |2τ , 1τ i, |2τ , 2τ i} (39) entangled qubits. Suppose that Bob takes C across the world. Alice should perform a Bell measurement (38) of identically to the coarse-grained Bell measurement AB. One of four possible outcomes will obtain. Alice should tell Bob which, via telephone. Her call communi- |Φ+ihΦ+| + |Φ−ihΦ−|, |Ψ+ihΨ+| + |Ψ−ihΨ−| . (40) cates log2(4) = 2 bits. Bob should transform C with a unitary whose form depends on the news. C will come Proof. The projector (30) transforms the two-qubit space to occupy the state |ψi. A will occupy a different state. as Alice will have teleported |ψi to Bob. in- Π = |1 , 2 ih1 , 2 | + |2 , 1 ih2 , 1 | . (41) We introduce a variation on quantum teleportation, AB τ τ τ τ τ τ τ τ coherent teleportation. The protocol illustrates the power

Let us relabel 1τ as 0 and 2τ as 1. The projector becomes of Posner binding. The protocol relies on entanglement, classical information, and Posner binding. + + − − ΠAB = |Ψ ihΨ | + |Ψ ihΨ | . (42) Posner binding resembles a coarse-grained Bell mea- surement, as shown in Sec. VI A. Hence Posner binding Direct substitution into the RHS yields the LHS. fails to teleport all the information teleportable with a Consider the complementary projector in the measure- Bell measurement. The coherences in |ψi are not sent. ment (31). 1 − ΠAB transforms the effective two-qubit A classical random variable, which results from decoher- space as ing |ψi, is. The set-up is introduced in Sec. VI B 1. The protocol 1 − ΠAB = |1τ , 1τ ih1τ , 1τ | + |2τ , 2τ ih2τ , 2τ | . (43) is introduced in Sec. VI B 2 and analyzed in Sec. VI B 3. 19

VI B 1. Set-up Posner C occupies (Bob holds) the reduced state 0 0 2 2 ρC := TrAB(|χ ihχ |) = |c0| |0τ ih0τ | + |c1| |1τ ih1τ | Let |jτ i denote an arbitrary τ = j eigenstate of C, for | |2| ih | j = 0, 1, 2. The |jτ i’s form the computational basis for + c2 2τ 2τ . (52) the qutrit space span{|0τ i, |1τ i, |2τ i}. This basis serves, Posner C’s state encodes information about |ψi, the in incoherent teleportation, similarly to the σz eigenbasis square moduli of the coefficients in Eq. (47). Yet C has in conventional teleportation. never interacted with A directly. Information has tele- Consider restricting the projector (30) to the space of ported from A to C, with help from |+τ , +τ i and from two qutrits: Posner binding. 0 ΠAB := |0τ , 0τ ih0τ , 0τ | + |1τ , 2τ ih1τ , 2τ | Posners A and B had a probability

+ |2τ , 1τ ih2τ , 1τ | . (45) pΠ = Tr(ΠAB TrC (|χihχ|)) (53) Let |+ i := √1 (|0 i + |1 i + |2 i). τ 3 τ τ τ of binding together. (An analogous probability can be introduced into quantum teleportation: Let Alice have a nonzero probability of failing to perform her Bell mea- VI B 2. Incoherent-teleportation protocol surement.) Suppose, instead, that A and B fail to bind together. Let A, B, and C denote three Posners. Suppose that The projector B and C begin in |+ , + i, then bind together.18 The τ τ 1 − Π = (Π ⊗ Π ) + (Π ⊗ Π ) joint state becomes AB τA=0 τB =1 τA=0 τB =2 + (Πτ ⊗ Πτ ) + (Πτ ⊗ Πτ ) 1 A=1 B =0 A=1 B =1 ΠBC |+τ , +τ i = √ (|0τ , 0τ i + |1τ , 2τ i + |2τ , 1τ i) . + (Πτ =2 ⊗ Πτ =0) + (Πτ =2 ⊗ Πτ =2) (54) 3 A B A B (46) projects the state of AB. The three-Posner state |χi [Eq. (48)] updates to In the first term’s absence, (46) would be a triplet. A triplet is a Bell pair, a maximally entangled state that 1 [(1 − ΠAB) ⊗ 1] |χi = [c (|0τ , 1τ , 2τ i + |0τ , 2τ , 1τ i) can fuel quantum teleportation. (46), we will show, fuels 2 0 incoherent teleportation. + c1(|1τ , 0τ , 0τ i + |1τ , 1τ , 2τ i) Suppose that, after (46) is prepared, Posners B and + c2(|2τ , 0τ , 0τ i + |2τ , 2τ , 1τ i)] C drift apart. (In quantum-computation language, Alice 00 and Bob share a Bell pair.) Let B approach A. Let A =: |χ i . (55) occupy an arbitrary state Posner C occupies (Bob holds) the reduced state

|ψi = c0|0τ i + c1|1τ i + c2|2τ i . (47) 00 00 1 2 2 TrAB(|χ ihχ |) = [(|c1| + |c2| )|0τ ih0τ | (56) The complex coefficients satisfy the normalization condi- 2 P2 2 2 2 2 2 tion j=0 |cj| = 1. (In quantum-computation language, + (|c2| + |c0| )|1τ ih1τ | + (|c0| + |c1| )|2τ ih2τ |] . |ψi is the unknown state that contains information that Again, C contains information about |ψi, despite never Alice will teleport to Bob.) The three Posners occupy having interacted directly with A. the joint state Suppose that Bob measures GC , the observable that |χi := |ψi (ΠBC |+τ , +τ i) . (48) generates the unitary C. Bob samples from a random variable whose values 0, 1, and 2 are distributed ac- Suppose that Posners A and B bind together. (Dur- cording to (p0 = |c |2 + |c |2, p0 = |c |2 + |c |2, p0 = ing the analogous quantum-teleportation step, Alice per- 0 1 2 1 2 0 2 |c |2 + |c |2). forms a Bell measurement of her qubits.) The three- 0 1 Posner state becomes

ΠAB|χi/hχ|ΠAB|χi (49) VI B 3. Analysis of incoherent teleportation

= c0|0τ , 0τ , 0τ i + c1|1τ , 2τ , 1τ i + c2|2τ , 1τ , 2τ i (50) =: |χ0i . (51) Five points merit analysis. First, we quantify the clas- sical information teleported. Second, we characterize the QI not teleported. Third, we compare the resources re- quired for incoherent teleportation to the resources re- 18 One might worry that the spin state would decohere before quired for quantum teleportation. Incoherent telepor- the Posners bound. But chemical binding consists of electronic tation, we show fourth, implements — dynamics. |+τ , +τ i is a state of nuclear spins. Nuclear dynam- ics tend to unfold much more slowly than electronic dynamics. the effective sending of much classical information via the The Born-Oppenheimer approximation reflects this separation of direct sending of little classical information, with help time scales. Hence the nuclear state appears unlikely to decohere from entanglement. Fifth, we explain how to prepare before the Posners bind. |+τ , +τ i and |ψi with Posner operations. 20

a. Quantification of the information teleported: A fails to bind to B.A GC measurement of Posner C Posners A and B teleport a trit to C.A trit is clas- simulates a measurement of the GC of |ψi. sical random variable that can assume one of three pos- d. Incoherent teleportation as superdense coding: sible values. Imagine preparing a Posner in the state Incoherent teleportation offers less power, we have seen, |ψi [Eq. (47)] and measuring GC . The measurement than quantum teleportation. Yet incoherent teleporta- 2 has a probability p0 = |c0| of yielding 0, a probabil- tion offers more power than classical communication. 2 2 ity p1 = |c1| of yielding 1, and a probability p2 = |c2| Suppose that Alice has incoherently teleported |ψi. Bob of yielding 2. So does a GC measurement of C, if A binds may wish to know which probability distribution he 0 0 0 to B [Eq. (52)]. The distribution has been teleported holds, {p0, p1, p2} or {p0, p1, p2}. Alice should send Bob from A to C. a bit directly: a zero if A bound to B and a one other- Suppose that A fails to bind to B. Measuring Posner wise.19 0 1 2 2 C has a probability p0 = 2 (|c1| + |c2| ) of yielding 0, Alice would directly send Bob a bit, while effectively 0 1 2 2 sending a trit, with help from entanglement and Pos- a probability p1 = 2 (|c2| + |c0| ) of yielding 1, and a 0 1 2 2 ner binding. A trit is equivalent to log2(3) > 1 bits. probability p2 = 2 (|c0| + |c1| ) of yielding 2 [Eq. (56)]. The measurement of Posner C is equivalent to an encoded Hence Alice packs much classical information (a trit) into generalized measurement of |ψi. a small classical system (a bit). A positive-operator-valued measure (POVM) Much classical information packs into a small classi- cal system, with help from a Bell pair and a Bell mea- {M1,M2,...,M`} represents a generalized quan- tum measurement [7]. The measurement elements are surement, in superdense coding [11]. Conventional super- positive operators Mk > 0. They satisfy the complete- dense coding packs two bits into one. Our protocol packs P † 1 information less densely. ness condition k Mk Mk = . The Mk’s need not be projectors, unlike PVM elements. e. Preparing |ψi and |+τ , +τ i: Incoherent telepor- Consider the POVM tation involves two coherent quantum states, |ψi and |+τ , +τ i. Instances of these states can be prepared with 1 Posner operations. We illustrate with an example in {|0τ ih0τ | = √ (|1τ ih1τ | + |2τ ih2τ |), (57) 2 App. H. To construct each state, one arranges singlets 1 in each Posner. Then, one rotates spins about the ylab- |1τ ih1τ | = √ (|2τ ih2τ | + |0τ ih0τ |), (58) 2 axis. Alternative preparation protocols might exist. 1 |2τ ih2τ | = √ (|0τ ih0τ | + |1τ ih1τ |)} . (59) 2 VI C. Application 2 of binding Posner molecules: Projecting Posner molecules onto their τ = 0 Measuring this POVM is equivalent to measuring the en- P subspaces coded observable GC := j jτ |jτ ihjτ |. Measuring the GC of |ψi has a probability p0 of yielding the encoded out- j The AKLT state can be prepared via projections onto come jτ . subspaces associated with the spin quantum number s = Suppose that Posners A and B fail to bind. A measure- 3 0 2 . Posners can occupy a variation AKLT on the AKLT ment of the GC of C simulates an encoded measurement state. The Posners must be projected onto their τ = 0 of the GC of |ψi. subspaces (Sec. VII). Posner-binding measurements can b. Classicality of the teleported information: Only 2 effect these projections. the square moduli |cj| are teleported. The coefficients’ phases are not. Hence incoherent teleportation achieves Proposition 2. Let A, B, C, . . . , M label m Posner less than quantum teleportation does. molecules. The following sequence of events projects each Section VI A clarifies why: Quantum teleportation in- Posner’s state onto the τ = 0 eigenspace: volves Bell measurements. Incoherent teleportation in- 1. A and B bind together, then drift apart. volves measurements of whether τA + τB = 0. The τA +τB = 0 measurement simulates a coarse-grained Bell 2. B and C bind together, then drift apart. measurement. 3. and bind together, then drift apart. , , and c. Comparison of resources required for incoherent C A A B have been projected onto their subspaces. teleportation with resources required for quantum telepor- C τ = 0 tation: In quantum teleportation, qubit C undergoes a local unitary conditioned on the Bell measurement’s out- come. Our Posner C needs no such conditional correct- 19 How could such classical communication manifest in biologi- ing. cal systems? In ordinary QI protocols, classical communication Yet part of our story depends on the Posner-binding manifests as telephone calls. Today’s phones do not fit in human brains. But one can envision classical channels in a biofluid. For measurement’s outcome: the interpretation of the out- example, if A and B bind, they shove water molecules away to- come of a GC measurement of Posner C. Suppose that gether. If A and B fail to bind, water propagates away from A binds to B.A GC measurement of Posner C simulates them differently. The patterns in the fluid’s motion may be dis- tinguished. The fluid-motion pattern would encode the bit. a measurement of the GC of |ψi. Suppose, instead, that 21

4. Each remaining Posner (D,...,M) binds to a pro- jected Posner, then drifts away. Proof. First, we prove that steps 1-3 project A, B, and C onto their τ = 0 subspaces. Then, we address step 4. A projector of the form (30) represents each binding. A product Π123 of projectors represents the sequence 1-3 of bindings: h   1⊗(m−2) Π123 = ΠτA=0 ⊗ ΠτB =0 ⊗ (60)   i 1⊗(m−2) + ΠτA=±1 ⊗ ΠτB =∓1 ⊗ h   1 1⊗(m−3) × ⊗ ΠτB =0 ⊗ ΠτC =0 ⊗   i 1 1⊗(m−3) + ⊗ ΠτB =±1 ⊗ ΠτC =∓1 ⊗ h   1 1⊗(m−3) × ΠτA=0 ⊗ ⊗ ΠτC =0 ⊗   i 1 1⊗(m−3) + ΠτA=±1 ⊗ ⊗ ΠτC =∓1 ⊗ 1⊗(m−3) = ΠτA=0 ⊗ ΠτB =0 ⊗ ΠτC =0 ⊗ . (61) FIG. 14: AKLT0 state: Spins on a honeycomb lattice can Equation (60) can be understood in terms of a frustrated occupy the Affleck-Lieb-Kennedy-Taski (AKLT) state lattice, as explained in App. I. |AKLThoni [21]. |AKLThoni serves as a universal resource in Step 4 of Proposition 2 is proved as follows. Suppose measurement-based quantum computation (MBQC) [20, 24]. 0 that Posners A and D bind together, then drift apart. So does the similar state |AKLThoni, which Posner The joint state of AD is acted on by operations can prepare efficiently. Each dashed oval encloses the spins in a Posner molecule. Each molecule consists of two trios of phosphorus nuclear spins. Each large black dot (Πτ =0 ⊗ Πτ =0) + (Πτ =±1 ⊗ Πτ =∓1) . (62) A D A D represents a trio. Each small white dot represents a spin. The state of A was projected onto the τA = 0 subspace Each thin black line connects the two spins in a singlet. 0 during steps 1-3. Hence the final term in Eq. (62) anni- This figure resembles Fig. 3a of [24], as |AKLThoni resembles |AKLThoni. This figure does not illustrate the spatial hilates the AD state. Hence ΠτD =0 projects the state of 0 arrangement of Posners in |AKLThoni. Rather, the figure D. 0 illustrates the entanglement in |AKLThoni. Proposition 2 will provide a subroutine in the following section’s protocol. the first state recognized as a matrix product state (MPS) [59–61]. MPSs can efficiently be represented VII. EFFICIENT PREPARATION OF POSNER approximately by classical computers. Also, using MOLECULES IN UNIVERSAL |AKLT1Di, one can simulate arbitrary single-qubit rota- QUANTUM-COMPUTATION RESOURCE tions. One performs local operations, including adaptive STATES single-qubit measurements,20 on the state [56, 57, 63]. Two-dimensional (2D) analogs of |AKLT1Di have been How complicated an entangled state can Posner opera- 3 defined. Spin- 2 particles on a honeycomb lattice can tions (Sec. V) prepare efficiently? Many measures quan- occupy the state |AKLThoni [21]. Local operations on tify multipartite entanglement. We study computational |AKLThoni can efficiently simulate universal quantum resourcefulness. Posners operations, we show, can effi- computation [20, 24]. We will draw on the proof by Wei ciently prepare a state that fuels universal MBQC: By et al. [24]. Similar results appear in [20]. operating on the state locally, one can efficiently simu- Wei et al. prove the universality of |AKLThoni as fol- late a universal quantum computer. lows. Local POVMs, they show, reduce |AKLThoni to an The Posner state is a variation on an Affleck-Lieb- encoded 2D graph state |G(A)i.21 The graph G is ran- Kennedy-Tasaki (AKLT) state. AKLT first studied a dom, depending on the set A of measurement outcomes. one-dimensional (1D) chain of spin-1 particles. They con- Also the encoding depends on A. The overline in |G(A)i structed a nearest-neighbor antiferromagnetic Hamilto- nian [21–23]. The ground state, |AKLT1Di, has a known form. A constant gap, independent of the system size, separates the lowest two energies. 20 Measurements are adaptive if earlier measurements’ outcomes |AKLT1Di has many applications in quantum com- dictate which measurements are performed later. 21 putation [20, 24, 56–65]. For example, |AKLT1Di was A graph state is defined in terms of a graph G. Each vertex 22

represents the encoding. Wei et al. prescribe local mea- |AKLThoni is trivalent: Each site links, via singlets, to surements of a few qubits. The measurements convert three other sites. |G(A)i into a cluster state on a 2D square lattice (if A is a typical set).22 Such cluster states serve as resources 0 in universal MBQC [12, 13, 19, 20]: By measuring single VII B. Preparing Posner molecules in |AKLThoni qubits adaptively, one can efficiently simulate a universal quantum computer. Posner operations (Sec. V) can nearly prepare We introduce a variation on |AKLThoni. We call |AKLThoni. Whether Posner operations can project trios 0 3 the variation the AKLT state and denote the state by onto their s123 = subspaces remains unknown. But 0 0 2 |AKLThoni. Figure 14 illustrates the state. |AKLThoni is Posner operations can project onto a molecule’s τ = 0 prepared similarly to |AKLThoni, resembles |AKLThoni subspace. locally, and fuels universal MBQC similarly. The τ = 0 subspace decomposes into a direct sum This section is organized as follows. Section VII A re- of tensor products of two three-qubit subspaces. The views the set-up and the state construction of Wei et al. first three-qubit subspace is labeled by s123, the total 0 |AKLThoni is defined in Sec. VII B. How to construct spin quantum number of the qubit triangle at zin = h+. 0 |AKLThoni efficiently from phosphorus nuclear spins, us- The second three-qubit subspace is labeled by s456. The ing Posner operations, is detailed. Section VII C de- 3 3  1 1 ⊕2 0 τ = 0 subspace has the form 2 ⊗ 2 ⊕ 2 ⊗ 2 . (See scribes the reduction of |AKLThoni to a 2D cluster state, Appendices E and F for a derivation. See [66] for back- known to fuel universal MBQC. The protocol is analyzed ground and notation.) The first term represents the space in Sec. VII D. 3 3 3 0 s123 ⊗ s456 = 2 ⊗ 2 of two spin- 2 particles. Wei et al. |AKLThoni holds interest not only as a computational project onto this space, in step 2. resource, but also in its own right. The state is analyzed Projecting onto the larger τ = 0 space yields in Sec. VII E. For instance, AKLT0 is shown to be a 0 |AKLThoni. We now detail how Posners can come to PEPS. 0 occupy |AKLThoni. The steps are explained in physical terms (of molecules, binding, etc.). Figure 15 recasts the protocol in operational terms, as a quantum circuit: VII A. Set-up by Wei et al. 1. Pyrophosphatase enzymes cleave some number N of − Wei et al. consider a 2D honeycomb lattice, illustrated diphosphates. N singlets |Ψ i are prepared (via op- in Fig. 3a of [24]. (Figure 14 has nearly the same form.) eration 1). A black dot represents each site. At each site sit three 2. The phosphates group together in trios. Singlets con- spin- 1 DOFs. White dots represent these DOFs, called 2 nect the trios as thin black lines connect the white virtual spins . dots in Fig. 14. Let s123 and m123 denote a site’s total spin quan- tum number and total magnetic spin quantum number. 3. Each trio, with a nearest-neighbor trio, forms a Posner These numbers can assume the values (s123, m123) = molecule (via operation 3). 1 1 3 3 3 1 ( 2 , ± 2 ), ( 2 , ± 2 ), and ( 2 , ± 2 ). The qubit trio can behave 1 3 4. Posners approach each other head-on, with opposite as one physical spin of s123 = or . 2 2 momenta, then drift apart, ideally as described in |AKLThoni may be prepared as follows [21]: Proposition 2. For simplicity, we focus on the case 1. Consider two nearest-neighbor sites. Choose a virtual in which each Posner P approaches only Posners P 0 spin in each site. Form a singlet |Ψ−i between these that are nearest neighbors of P in the hexagonal lat- spins. Perform this process on every pair of nearest- tice (Fig. 14). But this assumption is unnecessary. neighbor sites. Suppose that each approach leads to binding. Propo- sition 2 is realized. Every Posner’s state is projected 2. Project each physical spin (each site) onto its s = 3 123 2 onto the τ = 0 subspace. subspace. But two approaching Posners might fail to bind. The 23 7 success probability ≈ 18 . Suppose that Posners P and P 0 fail to bind. Suppose that P has already been corresponds to a spin. Consider the Hamiltonian   X x O z HG = σ σ . (63)  i k 23 i∈G k∈NB(i) This probability is calculated as follows. The N Posners oc- cupy some pure state |ψi. Consider the two Posners’ joint re- i indexes the vertices in G. The nearest neighbors of i are indexed duced state, ρPP 0 . The Posners share one singlet. The Pos- by k ∈ NB(i). HG has a unique ground state. This ground state ner pair contains ten other phosphorus nuclear spins. Let a de- is called a graph state [12, 18]. note an arbitrary one of these spins. a forms a singlet with 22 A cluster state is a graph state associated with a regular lattice a spin in some other Posner, P 00. P 00 is traced out from G [12, 17, 18]. |ψi in this calculation of ρPP 0 . Hence ρPP 0 equals a tensor 23

0 0 FIG. 15: Part of a circuit that efficiently prepares the AKLT state |AKLThoni: The circuit elements represent Posner operations, as explained in Sec. V A. Some thin black lines extend off the diagram. These lines represent singlets that terminate in Posners not drawn here.

0 projected onto its τP = 0 subspace. P can be “re- 2 explains the need to deviate from this reduction. Sec- freshed”: Let P 0 drift into a region of high pH and/or tion VII C 3 details the deviation. high Mg2+ concentration. P 0 likely hydrolyzes (un- dergoes operation 6). Two of the phosphorus nuclear spins used to form a singlet. They form a singlet no VII C 1. Model: Reduction of |AKLThoni to a cluster state longer, due to the binding failure. These two spins can drift away; a fresh singlet can replace them. Four Wei et al. prescribe two steps. First, each site is mea- other phosphorus nuclear spins remain. They continue sured with a POVM. The measurements yield a state to form singlets with spins in other Posners.24 The two equivalent to a random graph state. Second, a few qubits new, and four old, phosphates can form a Posner P˜0, are measured with local POVMs. via operation 3. Let us detail the initial measurements. Site v is mea- P˜0 occupies the state that P occupied before the bind- sured with the POVM ing failure. P˜0 can approach P head-on. If the binding ( r 0 2 fails, P˜ can be refreshed again. Fv,x = (|± ± ±ih±± ±|) , (64) 3 r 0 2 VII C. Reduction of |AKLThoni to a cluster state Fv,y = (|i, i, iihi, i, i| + |−i, −i, −iih−i, −i, −i|) , known to fuel universal MBQC 3 r ) 2 Fv,z = (|000ih000| + |111ih111|) . Local operations can reduce |AKLThoni to a cluster 3 state on a 2D square lattice [24]. Such cluster states zin serve as universal resources in MBQC [12, 13, 19]. Sec- The Fv,z projects onto the subspace spanned by the S123 eigenstates associated with the magnetic spin quantum tion VII C 1 reviews the reduction in [24]. Section VII C 3 numbers m123 = ± 2 . Fv,α projects onto the subspace αin spanned by the analogous S123 eigenstates, for α = x, y. Each subspace has dimensionality two. Hence the 1 product of ten maximally mixed qubit states 2 and |Ψ−i: POVM reduces each site’s Hilbert space to a qubit 2 q  1 ⊗5  1 ⊗5 2 2 − − 2 0 space. The leads to the the completeness relation ρPP 0 = 2 ⊗ |Ψ ihΨ | ⊗ 2 . Posners P and P 3 7 † have a probability ≈ Tr (Π 0 ρ 0 ) = of binding. [Π 0 is P PP PP 18 PP α=x,y,z Fv,αFv,α. defined as in Eq. (30).] This approximation does not account for A denotes the set of POVM outcomes. The POVMs many correlations amongst sites [24]. The exact calculations are yield a state |G(A)i. G(A) denotes a random graph left as an opportunity for future study. 24 This claim can be checked via direct calculation. Computational whose form depends on A. |G(A)i denotes the graph resources limited our calculation to the reduced state of 13 spins. state associated with G(A). The overline denotes an en- Whether longer-range correlations affect the results is left for coding dependent on A. The system occupies a state future study. See footnote 23. equivalent, via the encoding, to |G(A)i. 24

A random graph G(A) defines |G(A)i. In contrast, a regular graph defines a cluster state. The cluster state on a 2D square lattice serves as a universal resource in MBQC. This cluster state can be distilled from |G(A)i, if A is typical. The distillation consists of a few single-qubit Pauli measurements [24].

0 VII C 2. Toward a reduction of |AKLThoni to a cluster state

The Wei et al. system differs from the Posner system in two ways. First, Posner operations cannot necessarily simulate (i) the POVM (64) or (ii) the Pauli measure- ments in Sec. VII C 1. Whether Posner operations can remains an open question. Second, Wei et al. invoke 0 |AKLThoni. Posner operations can prepare |AKLThoni. Posners therefore require a step absent from [24]. To facilitate the explanation, we invoke the agent framework of QI (Sec. V B 3). Different experimentalists can per- form different operations easily. An agent Alice might run a biochemistry lab. She might be able to effect Pos- ner operations. An agent Bob might be able to perform FIG. 16: Coarse-graining the hexagonal lattice into local POVMs but not to create and arrange singlets. a square lattice: The black lines form a hexagonal lattice Together, Alice and Bob could produce cluster states. (see Fig. 14). Three qubits (small white dots) occupy each Alice would create |AKLT0 i and pass the state to Bob. site (large black dot). Two neighboring sites form a Posner hon molecule (encircled with a dashed hoop). The lattice can be Bob would perform local POVMs. (He might ask Alice coarse-grained: Each Posner’s two sites can be lumped to refresh a few Posners.) Together, the agents would together (into a red dot). The coarse-grained lattice is form cluster states that fuel universal MBQC. square (as shown by the long, red lines). This coarse-graining might facilitate a universality protocol simpler than the one in Sec. VII C 3. 0 VII C 3. Reduction of |AKLThoni to a cluster state

0 Bob will perform the protocol in Sec. VII C 1. But |AKLThoni to a cluster state on a 2D square lattice. This 2 2 first, he measures each Posner’s S123 ⊗ S456. Suppose cluster state can be used directly in universal MBQC. 3 that Posner P yields the outcome labeled by 2 [yields Posners’ universality is remarkable. Most quantum states 2 3 2 9 2 cannot power universal MBQC [16]. The Posners’ sin- the outcome ~ 2 × 2 = 2 ~ ]. The measurement has succeeded. glets and their geometry (the decomposition of Posners into triangles, and the triangles’ trivalence), underlie the Now, suppose that Posner P yields the outcome la- 26 1 state’s universality. beled by 2 . The measurement has failed. Bob returns P to Alice. Alice hydrolyzes P via operation 6. She re- Opportunities for enhancing and simplifying our pro- freshes the internal singlet, as in step 4 in Sec. VII B. tocol exist: ˜ 25 Let P denote the refreshed Posner. Bob measures the 1. A precise structure—a honeycomb lattice—underlies 2 2 ˜ 0 S123 ⊗ S456 of P . He and Alice “repeat until success” |AKLT i. In contrast, biomolecules drift randomly. 3 hon (until obtaining the 2 outcome). Biological singlets likely will not form honeycombs of Bob holds an |AKLThoni state. He now follows the prescription of Wei et al. (Sec. VII C 1).

26 0 |AKLThoni is not the simplest universal Posner resource state. 0 Singlets on a trivalent lattice would suffice. The τ = 0 pro- VII D. Analysis of |AKLThoni preparation jections are unnecessary. The unnecessariness follows from the 0 second equality in Eq. (31) of [24]. Yet |AKLThoni merits Posner operations, we have shown, can prepare defining, for three reasons. First, preparing |AKLT0 i is nat- 0 hon |AKLT i efficiently. A few local measurements reduce ural: In |AKLThoni, pairs of sites are projected onto their hon 3 3 0 s123 ⊗ s456 = 2 ⊗ 2 subspaces. In |AKLThoni, pairs of sites 0 are projected onto slightly larger subspaces. Hence |AKLThoni resembles |AKLThoni locally. Second, suppose that Alice did 25 P contained four spins apart from the internal singlet. Each of not project the Posners onto their τ = 0 subspaces. Bob would 1 these spins remains in a singlet with a spin in another Posner. obtain more “error” outcomes, labeled by 2 ’s, in Sec. VII C 3. 0 See the calculational comments in footnote 24. Third, |AKLThoni holds interest in its own right (Sec. VII E). 25

their own accord. Random graphs will more likely arise. Such graphs underlie states that might power MBQC. Such graphs might have two or three dimensions. Three-dimensional (3D) graph states offer particular promise. First, they have substantial connectivity, needed for universality [24]. Second, 3D cluster states fuel fault-tolerant universal MBQC. The scheme relies on toplogical quantum error correction [67]. 2. Bob might avoid returning Posners to Alice. The Pos- ners’ triangles (Fig. 1) form the sites in a hexagonal lattice (Fig. 14). Consider coarse-graining two sites into one. Triangle pairs are coarse-grained into Pos- ners. Each Posner forms a site in a square lattice (Fig. 16). 2 2 Imagine Bob measuring the S123 ⊗ S456 of a Posner 1 P . Suppose that the “error” 2 outcome obtains. Bob might discard P . Alternatively, he might measure the zin 0 0 S1...6 of P . The measurement would “terminate the FIG. 17: The AKLT state |AKLThoni as a projected lattice,” forming a boundary. entangled-pair state (PEPS): The two tensors, T + and T −, are repeated to form the PEPS. T + represents the state On average, ≈ 0.64 of the Posners yield the “good” 0 3 27 of one triangle in a Posner (Fig. 1); T represents the other ( 2 ) outcome. The 2D square lattice has a site- triangle’s state. Each tensor has three physical qubits, 28 percolation threshold of p∗ ≈ 0.59. Hence Bob’s + − labeled aj or aj , wherein j = 1, 2, 3. Each tensor has three site-deletion probability lies far above the threshold: + − virtual legs, labeled vj or vj , wherein j = 1, 2, 3. Each of p ≈ 0.64 > 0.59 ≈ p∗. A large, richly connected com- + + − − v1 and v2 , and each of v1 and v2 , has bond dimension two. ponent spans Bob’s graph. Such components underlie + − v3 has bond dimension six, as does v3 . An implicit universality [24]. Kronecker delta δ + − constrains the virtual indices. v3 v3 Bob returns to regarding triangles, rather than Pos- ners, as vertices. The lattice looks hexagonal but con- tains holes. A large connected component spans also VII E. Analysis of the AKLT0 state this graph. Hence the Wei et al. prescription (Sec. VII C 1), or a related prescription, appears likely to turn 0 MBQC motivated the definition of |AKLThoni. Yet the state into a universal cluster state. 0 0 |AKLThoni holds interest in its own right. |AKLThoni To check, one might refer to [68–70]. The authors resembles the AKLT state |AKLThoni on a honeycomb consider faulty lattices: Sites might be deleted or mea- lattice. AKLT states have remarkable properties. We sured. discuss analogous properties, and opportunities to seek | 0 i 3. Universal quantum computation is unnecessary for more analogous properties, of AKLThon . achieving quantum supremacy [71]. Suppose that First, classical resources can represent AKLT states 0 compactly. The 1D AKLT state |AKLT1Di is an |AKLThoni has been converted into a cluster state on MPS [59–61]. |AKLThoni is a PEPS [58, 62, 73]. a 2D square lattice. Consider measuring single qubits 0 nonadaptively. A random distribution P is sampled. |AKLThoni is a PEPS, illustrated in Fig. 17 and detailed in App. J. Classical computers are expected not to be able to 0 sample from P efficiently [72]. Hence |AKLThoni is the ground state of some local, frustration-free Hamiltonian HAKLT0 [26]. The ground Aside from these opportunities, the calculational tech- state is unique [27]. The relationship between HAKLT0 nique in footnote 23 may be rendered more precise. and the Posner Hamiltonian HPos merits study. So does whether HAKLT0 has a constant-size gap [21, 22]. If 0 HAKLT0 has, |AKLThoni can be prepared efficiently via 27 This probability is estimated via the technique in footnote 23. cooling. 28 0 Site percolation is a topic in graph theory and statistical me- Third, |AKLThoni results from deforming |AKLThoni. chanics. Let G denote a graph of N sites. Consider deleting each AKLT states have been deformed via another strat- site v with probability 1 − p. If v is deleted, so are the edges that egy [70, 74–76]: Let H denote the Hamiltonian terminate on v. Let G0 denote the remaining graph. Does a path AKLT of edges traverse G0 from top to bottom? If so, G0 percolates. whose ground state is the AKLT state of interest. HAKLT p∗ denotes the percolation threshold. If p ≥ p∗, in the limit as is transformed with a deformation operator D(a) [70]. 0 0 N → ∞, G percolates. G does not if p < p∗. A phase transition The parameter a is tuned, changing the ground state. occurs at p = p∗. 0 |AKLThoni follows from a different deformation. We 26 start not from a Hamiltonian, but from the Hilbert space (C6)⊗2. Singlets are arranged; then the state is projected onto the τ = 0 eigenspace Hτ=0. Hτ=0 contains the 3 3 2 ⊗ 2 subspace. Projecting onto the latter subspace would yield an AKLT state. Enlarging the projector deforms the state. Wei et al. study an AKLT state’s computational power as a function of a [70]. Our state’s computational power might be studied as a function of the projected-onto space. (a)

VIII. ENTANGLEMENT’S EFFECT ON MOLECULAR-BINDING RATES

Consider two Posners approaching each other head- on, as described below operation 5. The Posners might bind together. They could form subsystems in a many- body entangled system. Entanglement affects the bind- ing probability, Fisher proposes [1]. (b) Fisher supports his proposal with an example [1, p. 5, Fig. 3]. Let a, a0, b, and b0 denote Posners. Let a be FIG. 18: Illustration of entanglement’s effect on entangled with a0, and let b be entangled with b0. Suppose molecular-binding rates: Each maroon circle represents a Posner molecule. Each molecule contains six phosphorus that a has bound to b. Suppose that a0 approaches b0 0 0 nuclear spins, represented by black dots. The dots group head-on. a and b have a higher probability of binding, together into trios (blue triangles) (see Fig. 1). Two dots Fisher argues, than in the absence of entanglement. We connected by a green, wavy line form a singlet |Ψ−i recast this narrative as a quantum circuit in Fig. 13. [Eq. (1)]. Some spins form singlets with spins in external Fisher supports his proposal by analyzing position- molecules. Hence some green, wavy lines extend outside this and-spin states. We use, instead, the Posner-binding Posner pair. Figure 18a shows a Posner pair that contains 43 PVM (31). Checking Fisher’s example lies beyond our no singlets. These Posners have a probability 128 = 0.336 of (classical) computational power. But we check the prin- binding together. Figure 18b shows a Posner pair that shares ciple behind his example quantitatively: Entanglement, one singlet. Each Posner contains one singlet. These Posners have a probability 73 ≈ 0.338 of binding. The entanglement we show, can affect the probability that two Posners bind 216 pattern raises the binding probability by ≈ 0.6%. (Fig. 18 and Sec. VIII A). The effect is small, an ≈ 0.6% increase. Yet this proof of principle paves a path toward large-scale QI-theoretic checking of Fisher’s conjecture. to B. The Posners occupy the state Random tumbling, we find in Sec. VIII B, can eliminate entanglement’s effect on binding rates.  ⊗3  ⊗3 1 ⊗3 1 ρ0 = 2 ⊗ |Ψ−ihΨ−| ⊗ 2 . (66) AB 2 2 VIII A. Illustration: Entanglement’s effect on Suppose that the Posners approach each other head-on. binding rates The Posners have a probability

Let A and B denote two Posners. Suppose that the AB 0 0 73 pAB = Tr (ΠAB ρAB) = ≈ 0.338 (67) system contains no singlets (Fig. 18a). Every phosphorus 216 nuclear spin forms a singlet with an external molecule. of binding. The reduced state of AB equals a product of maximally Entanglement raises the binding probability by a frac- 112 112 mixed states: ρAB = 64 ⊗ 64 . The identity operator tion defined on k is denoted by 1 . C k p0 1168 Suppose that the Posners approach each other head- AB − 1 = − 1 ≈ 0.006 , (68) on. The Posners have a probability pAB 1161 43 or by ≈ 0.6%. This rise is small. But it illustrates quanti- pAB = Tr (ΠAB ρAB) = ≈ 0.336 (65) tatively how, under Fisher’s assumptions, entanglement 128 influences binding rates. of binding, independently of any other event. The technique illustrated here can be scaled up. With Suppose, instead, that AB contains the singlets shown more computational power, one can store larger quantum in Fig. 18b. Each Posner contains one singlet. This sin- states. The conjecture associated with Fig. 13 can be glet links the Posner’s triangles. Another singlet links A checked. 27

VIII B. Random tumbling can eliminate in Sec. V B 3. The assumption may appear question- entanglement’s effect on molecular-binding rates. able here: Practicalities partially concerned DiVincenzo. Attempting to realize arbitrary qubit rotations through Suppose that six phosphorus nuclear spins occupy the biomolecular collisions appears impractical. Yet the fine- 0 control assumption facilitates a first-step analysis of what joint state ρAB [Eq. (66)]. Suppose that each nucleus ro- tates randomly. Each qubit a evolves under some unitary the model can achieve in principle. Incorporating ran- domness forms an opportunity for future research. Unˆa (θa). The rotation axisn ˆa and the rotation angle θa may be distributed uniformly. 1.“A scalable physical system with well- Suppose that six nuclei form Posner A, while the other characterized qubits”: Different physical DOFs en- nuclei form Posner B. The molecules occupy the joint code QI at different stages of a quantum-cognition com- state putation (Sec. III). Initially, phosphorus-31 (31P) atoms’ nuclear spins serve as qubits. These atoms occupy free- " 12 # " 12 # O O floating phosphate ions. ρ00 ({nˆ } , {θ }) = U (θ ) ρ0 U (θ )† . AB a a nˆa a AB nˆa a Six phosphates can join together, forming a Posner a=1 a=1 molecule. Each state of six independent spins trans- (69) forms into one antisymmetrized state (7). Hence anti- Suppose that A and B approach each other head-on. The symmetrized spin-and-position states store QI. Posners have a probability Third, each Posner has an observable GC (Sec. IV A). The eigenvalue τ assumes one of three possible values: 00 00 τ = 0, ±1. The eigenspaces Hτ have 24-, 20-, and 20-fold pAB ({nˆa} , {θa}) = Tr(ΠAB ρAB ({nˆa} , {θa})) (70) degeneracies. A state |τ=ji can be chosen from each of binding. Hτ=j eigenspace. span {|τ=0i, |τ=1i, |τ=2i} forms an On average over tumbles, the Posners have a probabil- effective qutrit space (Sections IV D 1 and VI). But Pos- ity ners are not associated uniquely with effective qutrits, to our knowledge (footnote 9). Z   QI-storing Posner systems can be scaled up spatially. p00 = Tr Π ρ00 ({nˆ } , {θ }) (71) AB AB AB a a 31P nuclear spins can form singlets distributed across {nˆa},{θa} Posners. Entangled Posners can form lattices that can " 1 ⊗3 Z 2 power universal MBQC (Sec. VII). = Tr ΠAB ⊗ (72) 2 10 10 {nˆa}a=3,{θa}a=3 Just as Posner clusters can be scaled up across space, " 10 # " 10 # 1 ⊗3 #! Posner computations can be scaled up across time. The O − − O † 2 × Unˆ (θa) |Ψ ihΨ | U (θa) ⊗ ability to perform arbitrarily long computations hinges a nˆa 2 a=3 a=3 on the ability to form, dissolve, and reform Posners  ⊗12! (Sec. V B 2). Only when qubits are in Posners can Posner 12 = Tr ΠAB (73) binding create entanglement. Only outside of Posners 2 can qubits rotate independently. Alternating entangle- ment generation with one-qubit rotations is necessary for = pAB (74) performing arbitrary computations. Hence qubits must of binding. Uniformly random rotations effectively de- form Posners, then separate, repeatedly. cohere the internal singlets, on average. The binding 2. “The ability to initialize the state of the probability reduces to its non-singlet-enhanced value. qubits to a simple fiducial state, such as |000 ...i”: Phosphorus nuclear spins can be prepared in singlets |Ψ−i (operation 1). A singlet forms as the enzyme py- IX. MEASURING QUANTUM COGNITION rophosphatase cleaves a diphosphate ion. The resultant AGAINST DIVINCENZO’S CRITERIA FOR two phosphates are projected onto |Ψ−i. QUANTUM COMPUTATION AND 3. “Long relevant decoherence times, much COMMUNICATION longer than the gate-operation time”: Fisher has identified three sources of decoherence and has estimated Consider attempting to realize universal quantum com- coherence times [1, 6, 77]. Magnetic fields B threaten putation and quantum communication with any physical 31P spins most: Nearby spins and electrons can generate platform. Which requirements must the physical com- B’s. The fields can couple to the spins’ magnetic dipole ponents and processes satisfy? DiVincenzo catalogued moments, µ: Hmag = −µ · B. these requirements [28]. Five of diVincenzo’s criteria un- Water molecules threaten 31P the most. But phos- derpin quantum computation. Quantum communication phates and Posners tumble in solution. Tumbling requires another two. Quantum cognition, we find, sat- changes the B experienced by a 31P. On average over isfies DiVincenzo’s criteria, except perhaps universality. tumbles, B is expected to vanish. We continue assuming that Posner operations can be 31P nuclear spins in free phosphates, Fisher estimates, performed with fine control. This assumption is discussed remain coherent for about a second. 31P nuclear spins in 28

Posners remain coherent for ∼ 105 − 106 s. 5. “A qubit-specific measurement capability”: Fisher writes also that τ labels a pseudospin “very iso- Posner binding (operation 5) measures whether τA+τB = lated from the environment, with potentially extremely 0. Whether two molecules have bound together is a clas- long (days, weeks, months, . . . ) decoherence times” [77]. sical property. Each Posner’s center of mass is a classical As explained in Sec. IV A, to our knowledge, τ does not DOF: Water and other molecules bounce off the Posner label any unique quantum three-level pseudospin. We frequently. The bounces measure the Posner’s position, translate Fisher’s statement as follows: Consider prepar- precluding coherence. Hence two Posners’ closeness is ing a GC eigenstate associated with the eigenvalue τ = j. a classical DOF. Closeness serves as a proxy for bind- Consider waiting, then measuring GC . How would long ing. Hence whether two Posners have bound is a classical would you have to wait to have a high probability of ob- DOF. This classical memory records whether τA+τB = 0. taining an outcome τ = k =6 j? At least days. Fisher proposed that the readout might be amplified These coherence times, we expect, exceed entangling- further [1]. Posner binding could impact later Posner gate times. Posner binding (operation 5) consists of elec- binding, then the hydrolyzation of Posners, then neu- tronic dynamics. Electrons have much shorter time scales rons’ Ca2+ concentrations, and then neuron firing. This than nuclei, in general. Posner binding releases about 1 process is overviewed in this paper’s introduction. eV of energy to the environment [1, 9]. A rough esti- 6. “The ability to interconvert stationary and ~ −15 mate for the binding’s time scale is tbind ∼ 1 eV ∼ 10 flying qubits”: Computing is often easiest with unmov- s  105 s. As many as 1020 entangling gates might be ing hardware. Stationary qubits remain approximately performed before the spins decohere. fixed. They undergo computation. Then, their state can Could many single-qubit rotations be performed? A be transferred to flying qubits. Flying qubits move easily. qubit rotates as a phosphate tumbles in an aqueous fluid. They can bring states together for joint processing. Let us estimate a rotation’s time scale. We classically Consider, for example, a that con- approximate tains subroutines. Different labs’ quantum dots could im- plement different subroutines. The quantum dots’ states thermal energy ∼ rotational energy . (75) could be converted into photonic states. The photons −23 could travel down optical fibers to a central lab. There, The thermal energy ∼ kBT , wherein kB ∼ 10 J/K denotes Boltzmann’s constant. The body has a temper- the algorithm’s final steps could be implemented. ature T ∼ 102 K. The rotational energy ∼ Iω2. The Phosphorus nuclear spins could serve as stationary classical moment of inertia is denoted by I. The angular qubits and as flying qubits. Phosphorus atoms occupy frequency is ω ∼ 2π ∼ 10 , wherein t denotes the ro- lone phosphates and Posners. In each setting, the nuclear trot trot rot tational time scale. We approximate a Posner molecule spins undergo computations (Sec. V A). But Posners pro- as a sphere and drop constants: I ∼ mr2, wherein m de- tect the spins from decoherence better than lone phos- notes the molecule’s mass and r denotes the radius. The phates do [1]. Hence Posners form better flying qubits. −24 − Posner has a mass m ∼ 10 kg [78, Supp. Mat., Table The projector Πno-coll. [Eq. (8)] transforms phosphate S1] and a radius r ∼ 10 A˚ = 10−9 m [4]. Substituting states into a Posner state. The projection forms a one- 2 2 2 102 to-one map. Hence “stationary” phosphates’ states are into (75) yields kBT ∼ mr ω ∼ mr t2 . Solving for the rot converted into a “flying” Posner’s state faithfully. rotational time scale yields “The ability faithfully to transmit flying r m −10 qubits between specified locations”: Posners diffuse trot ∼ 10r ∼ 10 s (76) kBT through intracellular and extracellular fluid. A protein could transport Posners into neurons [1, 6]. This pro-  1 s . (77) tein has been called the vesicular glutamate transporter Hence the single-qubit-rotation time scale is much less (VGLUT) [79–82] and the brain-specific (B) sodium- + than the out-of-Posner decoherence time. dependent (Na ) inorganic-phosphate (Pi) cotransporter 4. “A ‘universal’ set of quantum gates”: The bio- (BNPI) [83]. VGLUT sits in cell membranes, through logical qubits can undergo Posner operations, introduced which the protein could ferry Posners. Posners protect in Sec. V A. The operations include arbitrary single-qubit in-transit 31P nuclear spins for ∼ 105 − 106 s [1, 5], as unitaries and entangling operations. Analyses appear in discussed above. Sections V B and VI. For how long do Posners diffuse between neurons? We These qubit operations induce gates on the effective estimate by dimensional analysis. The diffusion constant qutrits. The induced gates depend on how the qutrits D has dimensions of distance2/time: are defined. Whether the gates form a universal set—in transform- `2 D ∼ . (78) ing the qubits or the qutrits—remains an open ques- tdiff tion. Posner operations can efficiently prepare a state 0 |AKLThoni that can fuel universal MBQC (Sec. VII). The time scale over which a Posner diffuses between neu- Whether Posner operations can implement MBQC re- rons is denoted by tdiff . A typical synapse has an area of mains unknown. `2 ∼ 10−2 µm2 [84, Fig. 2]. 29

We estimate D via the Einstein-Stokes relation, might decompose into Posner operations. (ii) Posner op- erations could inspire hitherto-unknown quantum algo- k T D = B . (79) rithms. 6πηr Reverse-engineering: QI processing could guide Equation (79) describes a radius-r sphere in a viscosity-η conjectures about quantum chemistry. Fisher reverse- fluid. Water has a viscosity η ∼ 10−3 N · s/m2. A Posner engineered physical mechanisms by which entanglement molecule has a radius r ∼ 10 A˚ [4]. could impact cognition [1]. Similarly, one might reverse- −23 engineer physical mechanisms by which Posners could We substitute these numbers, with kB ∼ 10 J/K, T ∼ 102 K, and 6π ≈ 10, into Eq. (79): D ∼ 10−10 m2/s. process QI. This paper motivates reverse-engineering op- portunities: We substitute into Eq. (78), upon solving for tdiff : `2 10−2(10−6 m)2 1. Section VII B details how Posner operations can effi- tdiff ∼ ∼ (80) ciently prepare states that can fuel universal MBQC. D 10−10 m2/s To use the states, one performs the operations in 5 = 0.1 ms  10 s . (81) Sec. VII C. Example operations include measurements of the POVM {F ,F ,F , } [Eq. (64)]. Then, one mea- Hence Posners are expected to be able to traverse a x y z sures single qubits adaptively. Could biological sys- synapse before their phosphorus nuclear spins decohere. tems implement these operations?

2. Reverse-engineer a measurement of the generator G X. OUTLOOK C of the permutation operator C. If GC can be measured, incoherently teleported random variables can be used This paper establishes a framework for the QI-theoretic easily (Sec. VI B 2). analysis of Posner chemistry. The paper also presents ap- plications of Posners to QI processing: to QI storage and Quantum computational complexity and uni- protection, to quantum communication, and to quantum versality: Posner operations (Sec. V) constitute a model computation. Many QI applications of Posners await of quantum computation. Which set of problems can this discovery, we expect. In turn, QI motivates quantum- model solve efficiently? Let PosQP denote the class of chemistry questions. Opportunities are discussed below. computational problems solvable efficiently with Posner Quantum error-correcting and -detecting quantum computation. codes: We presented one quantum error-detecting code Whether Posner quantum computation is universal re- and one error-correcting code accessible to Posners. mains an open question. (See Sec. V B 1 for an elabo- Other accessible codes might protect more information ration.) Suppose that the model were universal. PosQP against more errors. would equal BQP (the class of problems that a quantum Furthermore, one conserved charge “protects” each of computer can solve in polynomial time [7]). But perhaps our codes. In the error-detecting code, for example, the PosQP ⊂ BQP. PosQP merits characterization. 0 codewords |jτ i correspond to distinct eigenvalues of GC . AKLT state and MBQC protocol: Posner opera- 0 The natural dynamics protect GC . Hence the dynamics tions can efficiently prepare a state |AKLThoni that fuels should not map any codeword |jτ i into any other |kLi. universal MBQC (Sec. VII). The state preparation may 0 But the dynamics could map |jτ i to another state |jτ i in be simplified. Opportunities are detailed in Sec. VII D. 0 the τ = j eigenspace. Also, |AKLThoni holds interest outside of MBQC. Prop- Imagine a more robust code: A complete set of quan- erties to explore are discussed in Sec. VII E. tum numbers (e.g., {τ, m1...6,...}) would label each code- Entanglement’s effect on binding rates and bi- word. The dynamics could not map any codeword ological Bell tests: Entanglement between Posners af- 0 0 |τ, m1...6,...i into any other codeword |τ , m1...6,...i. fects binding rates. So Fisher conjectured in [1]. The Such a code would enjoy considerable protection by conjecture grew from analyses of spin-and-orbital states. charge preservation. We supported the conjecture with a small example, using Relatedly, quantum codes have been cast as the ground a PVM (Sec. VIII). The example illustrates how to check spaces of Hamiltonians. Every code’s states, |ψ¯i, occupy Fisher’s conjecture with the formalism of QI. Larger-scale a Hilbert space H¯. Suppose that H¯ is the ground space of calculations could test Fisher’s conjecture more directly. a Hamiltonian H. Suppose that the system is in thermal Moreover, the QI formalism could lead to a frame- equilibrium at a low temperature T = 1 . The sys- kBT work for biological Bell tests. Such tests might be cast as tem has a high probability of remaining in H¯. Entropy nonlocal games [86]. The Clauser-Holt-Shimony-Hauser suppresses errors. Equivalently, the code detects errors. (CHSH) game, which illustrates Bell’s theorem [87, 88], The Posner Hamiltonian HPos has yet to be character- can serve as a model. ized. The ground space might point to an entropically Quantum chemistry: Physical conjectures populate preserved a code. Sections II C, II D, and III C. These conjectures merit Quantum algorithms: Posners might perform quan- testing and refinement. First, Posner creation was mod- tum algorithms of two types: (i) Known algorithms [85] eled with a Lennard-Jones potential. Second, pre-Posner 30

spin states were assumed to transform deterministically Wei for discussions. We thank Fernando Pastawski for into antisymmetric Posner states. The pre-Posner orbital help with constructing the quantum error-detecting code. state was assumed to determine the map. Third, Posner NYH thanks John Preskill for nudges toward this paper’s creation was assumed to preserve each spin’s Szlab essen- topic and for feedback about drafts. We are grateful for tially. Fourth, Posner dynamics were assumed to preserve funding from the Institute for Quantum Information and zin C and S1...6. Matter, an NSF Physics Frontiers Center (NSF Grant Randomness: Our QI-processing protocols involve PHY-1125565) with support from the Gordon and Betty perfect executions of Posner operations. But Posners Moore Foundation (GBMF-2644). This research was par- randomly collide with other molecules, tumble, and ex- tially supported by the NSF also under Grant No. NSF perience electric and magnetic fields. Randomness could PHY-1125915. NYH is grateful for partial support from hinder some, and improve some, QI processing. the Walter Burke Institute for Theoretical Physics at For example, Sec. VII features a honeycomb lattice. Caltech, for a Graduate Fellowship from the Kavli In- Singlets would likely not form a honeycomb in solution. stitute for Theoretical Physics, and for a Barbara Groce They would form a random graph. Randomness could Women’s Graduate Fellowship. improve the state’s connectivity. Improved connectivity might lower the bar for fueling universal MBQC (Sec. VII D). What randomness helps, and what randomness hin- ders, merits investigation.

ACKNOWLEDGMENTS

The authors thank Ning Bao, Philippe Faist, Matthew Fisher, Steve Flammia, Leo Radzihovsky, and Tzu-Chieh

Appendix A BACKGROUND: QUANTUM INFORMATION THEORY

Quantum systems can process information more efficiently, transmit information more compactly, and secure infor- mation more reliably than classical systems can. Consider a system of N qubits, e.g., N phosphorus nuclear spins. N The system corresponds to a Hilbert space H of dimensionality 2 . Let {|φji} denote an orthonormal basis for H. P N The system can occupy a quantum state |ψi = j cj|φji ∈ H. The 2 coefficients cj ∈ C satisfy the normalization P 2 N N condition j |cj| = 1. Consider specifying one of the 2 basis elements |φji. One must use 2 bits (two-level units of classical information). The specification requires only N qubits. One can leverage this discrepancy to process information quickly, using quantum systems. The state |ψi constitutes QI. QI can be processed with help from entanglement [7, 8]. Entanglement manifests in correlations stronger than any shareable by classical systems. Entanglement facilitates quantum computation, communication, and cryptography. We briefly review efficiency, quantum computational models and universality, and quantum error correction. Readers seeking more background are referred to [7, 8].

A 1 Efficiency

Quantum computers can efficiently solve certain problems that, according to widespread belief, classical computers cannot. Efficiently loosely means the following. Consider a family F of computational problems. For example, consider receiving a number N whose prime factors you must identify. An instance of F consists of, e.g., the number N to be factored. Let n quantify the resources required to specify an instance of F . For example, n might equal the number of bits needed to represent N . Let t denote the time required to solve the instance. Suppose that the time grows, at most, polynomially in the amount of resources: t ∼ (const.)nk, for some k ≥ 0. The problems in F can be solved efficiently. Quantum computers can factor arbitrary numbers more quickly than classical computers can [89]. Imagine using a quantum computer to solve a problem more quickly than any classical computer. One would achieve quantum speedup, or quantum supremacy [71].29

29 Preskill coined the term “quantum supremacy” in [71]. The pa- quantum computers change the world? Predictions are never per concludes with ’s potential: “How might 31

A 2 Quantum-computation models and universality

A general quantum process consists of state preparations, evolutions, and measurements. Which operations can be implemented easily (which states |ψi can be prepared easily, etc.) varies from platform to platform. Consider, for example, a nuclear-magnetic resonance (NMR) experiment. Let N denote the number of nuclear spins. Preparing the pure state |0i⊗N is difficult. Preparing a maximally mixed state 1/2N−1 of N − 1 spins, tensored with one pure |0i, is easier [90]. A set of quantum resources—of performable quantum operations—forms a model for quantum computation. DiVincenzo catalogued the ingredients needed to realize a quantum-computation model physically [28]. Certain computational models are universal [91]. A universal quantum computer can perform every conceivable quantum computation. Every universal model can simulate every other universal model efficiently. Many quantum-computation models exist. Two prove most pertinent to this paper: the circuit model [15] and measurement-based quantum computation (MBQC) [12–14]. Other models include the quantum Turing ma- chine [91], the one-clean-qubit model [90], adiabatic quantum computation [92], anyonic quantum computation [93], teleportation-based quantum computation [52–55], quantum walks on graphs [94], and permutational quantum com- putation [95]. The circuit model is used most widely [15]. One solves a problem by running a quantum circuit, illustrated by a circuit diagram (e.g., Fig. 13). Wires represents the qubits, which are often prepared in pure states |0i. Rectangles represent unitary operations U. The U’s evolve the qubits, implementing gates. A rectangle inscribed with a dial represents a measurement. Single qubits can be measured with respect to some orthonormal basis, e.g., {|0i, |1i}, wherein h0|1i = 0. Depth quantifies a circuit’s length, or complexity. Consider grouping together the operations that can be performed simultaneously. For example, qubit 1 can interact with qubit 2 while qubit 3 interacts with qubit 4. Each group of gates occurs during one time slice. The number of time slices in a circuit equals the circuit’s depth. Suppose that the depth does not depend on the number of qubits. Such a circuit has constant depth. Primitive unitaries can be implemented directly. Composing primitives simulates more-complicated operations. One universal primitive set [7, 96] is natural to compare with 31P dynamics: (i) Each qubit’s state can rotate through a fixed angle θ about a fixed axisn ˆ of the Bloch sphere.30 θ must be an irrational multiple of 2π. (ii) Each qubit can rotate through a fixed angle θ0 about a fixed axisn ˆ0 =6 n ˆ. (iii) Any two qubits can be entangled via some fixed unitary. No unitary is known to entangle Posners’ phosphorus nuclear spins. Hence we turn from the circuit model to MBQC [12–14]. To implement MBQC, one prepares a many-qubit entangled state |ψi. One measures single qubits adaptively. Measurements are adaptive if earlier measurements’ outcomes dictate later measurements’ forms. Certain states |ψi enable one to simulate efficiently, via MBQC, a universal quantum computer. Example states include the Affleck-Kennedy-Lieb-Tasaki (AKLT) state on a honeycomb lattice, |AKLThoni [12, 13, 19, 20]. Posners 0 0 can occupy a similar state, |AKLThoni. |AKLThoni can fuel universal MBQC (Sec. VII).

A 3 Quantum error correction

Two sources of error threaten quantum computers. First, the operations performed might differ from the target π operations. Consider, for example, trying to rotate a qubit through an angle 2 about the z-axis. One might overshoot π or undershoot. The qubit would rotate through an angle 2 + , for some  =6 0. Second, a quantum computer might entangle with its environment. The environment decoheres the computer’s state. QI leaks from the computer into the environment. Quantum error correction preserves QI. Imagine wishing to process a state |ψi of k qubits. One chooses an error- correcting code. The code maps |ψi to a state |ψ¯i of n > k qubits. |ψ¯i undergoes physical processes that effect logical operations on the encoded state. The logical operations constitute a computation. Throughout the computation, certain observables O are measured. Which O’s depends on the code. The mea- surements’ outcomes imply whether an error has occurred and, if so, which sort of error. The code dictates how to counteract the error. The state is typically corrected with some unitary U. After the computation and correction terminate, the state is decoded. The computational problem’s answer is read out.

easy, but it would be especially presumptuous to believe that eiϕ sin θ |1i, wherein θ, ϕ ∈ [0, 2π). The state is equivalent to our limited classical minds can divine the future course of quan- 2 the Bloch vector (sin θ cos ϕ, sin θ sin ϕ, cos θ). The Bloch vec- tum information science.” Posners suggest that we have better tor lies on the unit sphere, or Bloch sphere. Points inside the chances than Preskill expected. sphere represent mixed states ρ 6= |ψihψ|. 30 The Bloch sphere represents pure qubit states geometrically [7]. θ A general pure qubit state has the form |ψi = cos 2 |0i + 32

A code can detect more errors than it can correct. Suppose that, according to the O measurements, many errors have corrupted |ψi. Suppose that the code cannot correct all those errors. The state must be scrapped; and the computation, reinitiated. We present a quantum error-detecting code and an error-correcting code formed from states accessible to Posners (Sec. IV D). Let us review the mathematics of quantum error correction and detection (QECD). Consider encoding k < n logical qubits in n physical qubits. The physical Hilbert space C2n has dimensionality 2n. A QECD code is a 2n k n comp subspace HL ⊂ C of dimensionality 2 < 2 . Let BL = {|jLi} denote the code’s computational basis. (See Sec. III B for an introduction to computational bases.) Each quantum error-correcting/-detecting code corresponds to a set {Eα} of correctable/detectable errors. For example, a code of n = 9 physical qubits has been constructed [97]. This code corrects the set of single-qubit Pauli x x x z z α 1⊗(j−1) α 1⊗(n−j) y errors, {σ1 , σ2 , . . . , σ9 , σ1 , . . . , σ9 } . The shorthand σj ≡ ⊗ σj ⊗ . The ability to correct σ errors follows from the ability to correct σx and σz. Under what conditions can a code HL detect a set {Eα} of errors? The code and set must satisfy the quantum error-detection criteria,

hjL|Eα|kLi = Cα δjk ∀j, k, α . (A1)

The Kronecker delta is denoted by δjk. Cα denotes a constant dependent only on the error Eα, not on the codeword labels j and k. Equation (A1) decomposes into two subcriteria: the off-diagonal criterion, in which j =6 k, and the diagonal criterion, in which j = k. The off-diagonal error-detecting criterion has the form

hjL|Eα|kLi = 0 ∀j =6 k . (A2)

No Eα maps any codeword |kLi into any other codeword |jLi. The logical states retain their integrity under detectable errors. The diagonal criterion has the form

hjL|Eα|jLi = Cα ∀j, α . (A3)

Suppose that |jLi is prepared. The environment might effectively measure hEαi. The environment gains no infor- mation about the state, according to Eq. (A3): Every codeword’s expectation value equals every other codeword’s. Typical detectable errors Eα operate nontrivially on just a few close-together qubits. The codewords are locally indistinguishable with respect to {Eα}. Local indistinguishability protects QI: Suppose that the environment had “learned” about |jLi. Information would have leaked out of the system. Highly entangled states are locally indistinguishable: Entanglement distributes infor- mation throughout the system. Local operations cannot extract the distributed information. We have reviewed the error-detection criteria. Under what conditions can a code HL correct {Eα}? The code must satisfy the quantum error-correction criteria [8, 36–40],

† hjL|EβEα|kLi = Cαβ δjk ∀j, k, α, β . (A4) Equation (A4) is interpreted similarly to Eq. (A1). A code that corrects (d − 1)/2 errors detects (d − 1) errors. We refer readers to [8] for more background.

Appendix B MULTIPLICITY OF (NO-COLLIDING-NUCLEI ANTISYMMETRIC) SUBSPACES ACCESSIBLE TO A POSNER MOLECULE

− A subtlety about Hno-coll. was glossed over in Sec. III C. Consider Eq. (7). In every term, the spin quantum number mπα(j) appears alongside the position rπα(j). The tuple (mπα(j), rπα(j)) occupies different kets in different terms. But mπα(j) remains hitched to the same position rπα(j) throughout the terms. How are the mπα(j)’s assigned to positions? This question has a two-part answer. The choice of coordinate system partially determines the assignments. So do initial conditions, the pre-Posner phosphates’ positions and momenta. The choice of coordinate system determines the ϕ-value associated with a given m-value. For example, suppose that m1 = 0. Should this spin variable be assigned to r1 = (φ, h), to r1 = (φ+2π/3, h), or to r1 = (φ+4π/3, h)? (Whether h = h+ or h = h− is irrelevant.) This assignment is a convention, because the orientation ofx ˆin is a convention. We illustrate the answer’s second part with an example. Suppose that three singlets, 1 1 1 |Ψ−i⊗3 = √ (|↑i|↓i − |↓i|↑i) ⊗ √ (|↑i|↓i − |↓i|↑i) ⊗ √ (|↑i|↓i − |↓i|↑i) , (B1) 2 2 2 33 join together to form a Posner. Molecule creation is assumed to preserve the entanglement within each pair of spins (Sec. II D). Posner creation maps each six-spin term in (B1) to a sum (7). Suppose we choose an intra-Posner coordinate system such that r1 = (0, h), for h = h+ or h = h−. Given that coordinate system, which value should r2 assume? Should the spin at (0, h±) form a singlet with the spin at (0, h∓), with the spin at (2π/3, h±), etc.? Different zin answers generate qualitatively different Posner states: The states transform differently under S1...6. Direct calculation supports this claim. The correct answer, we posit, is determined by the positions and momenta that the phosphates had at the lip of the Lennard-Jones potential (Sec. II C). Different projections of the same initial state, we posit, would release different amounts of heat to the environment. The PVM model in Sec. II D can now be refined: Posner creation projects the phosphates’ state with the projector − − Πno-coll. onto some no-colliding-nuclei subspace Hno-coll. of the antisymmetric subspace. 6! such subspaces exist; 6! − possible forms are available to Πno-coll.. One subspace and projector correspond to entanglement between the (0, h±) spin and the (0, h∓) spin; one subspace and projector correspond to entanglement between the (0, h±) spin and (2π/3, h±) spin; etc. Hence pre-Posner positions and momenta, with a choice of coordinate system, determine to which position each spin variable (e.g., m1) is assigned during Posner creation.

− Appendix C THE POSNER-MOLECULE HILBERT SPACE Hno-coll. HAS DIMENSIONALITY 64.

− − When a Posner forms, we posit, Πno-coll. projects the phosphorus nuclei’s joint state [Eq. (8)]. Πno-coll. defines a map that preserves the dimensionality of the space available for storing QI, 64. A counting argument shows why. th Imagine that the phosphorus nuclei were classical and distinguishable. A tuple (mj, rj) would label the j nucleus’s state. The spin variable mj could assume one of two possible values. The position rj could assume one of six possible values, ↑ or ↓. The tuple could therefore assume one of twelve possible values:

(↑, (φ, h+)), (↑, (φ, h−)), (↑, (φ + 2π/3, h+)), (↑, (φ + 2π/3, h−)), (↑, (φ + 4π/3, h+)), (↑, (φ + 4π/3, h−)), (C1)

(↓, (φ, h+)), (↓, (φ, h−)), (↓, (φ + 2π/3, h+)), (↓, (φ + 2π/3, h−)), (↓, (φ + 4π/3, h+)), or (↓, (φ + 4π/3, h−)) .

The nuclei would be “dodequits”: dim(Hnuc) would equal 2 × 6 = 12. Let us return to reality: The phosphorus nuclei are indistinguishable fermions. A hextuple of nuclei can occupy the antisymmetric basis state (7). This state is labeled by a set of six tuples. Each tuple must differ from each other tuple, for the state to be antisymmetric. To label a joint state, we choose six of the twelve possible tuples. But we cannot choose six arbitrary tuples. No two tuples can contain the same position: No two nuclei can coincide. Hence we pair up the twelve possible tuples. Each pair’s constituent tuples have the same positions and different spin states:

1. (↑, (φ, h+)), (↓, (φ, h+))

2. (↑, (φ, h−)), (↓, (φ, h−))

3. (↑, (φ + 2π/3, h+)), (↓, (φ + 2π/3, h+))

4. (↑, (φ + 2π/3, h−)), (↓, (φ + 2π/3, h−))

5. (↑, (φ + 4π/3, h+)), (↓, (φ + 4π/3, h+))

6. (↑, (φ + 4π/3, h−)), (↓, (φ + 4π/3, h−)) We have formed six pairs of tuples. We choose one tuple from each pair, to label an antisymmetric joint basis state. Let us count the ways in which we can choose the six tuples. We can choose one tuple from each pair in two ways. We choose from each of six pairs. Hence we have 26 = 64 choices of labels for an antisymmetric joint state.

zin Appendix D WHY A POSNER MOLECULE’S S1...6 IS EXPECTED TO BE CONSERVED

A Posner’s phosphorus nuclear spins resist decoherence for long times, according to Fisher [1]. We interpret this zin claim as meaning that the Posner’s dynamics preserve S1...6 (the z-component, relative to the Posner’s internal frame, of the six phosphorus nuclei’s total spin) for long times. Fisher supports his claim by arguing that the spins (i) fail to couple to electric fields, (ii) couple to magnetic fields weakly, and (iii) cannot couple to the Posner’s calcium and oxygen nuclear spins (as those atoms have spin quantum numbers s = 0) [1, 6]. 34

We supplement Fisher’s analysis to support our interpretation. Appendix D 1 concerns how phosphorus nuclear spins might interact. We identify candidate interactions that preserve the Posner’s C3 symmetry. These interactions zin preserve S1...6. So might collisions with other molecules (Sec. D 2). Collisions are expected to decohere just irrelevant orbital DOFs. Throughout the rest of this section, components are implicitly defined with respect to the internal coordinate system.

zin D 1 Why interactions amongst phosphorus nuclear spins are expected to conserve S1...6

The nuclei within a molecule can interact, in general. Intramolecule interactions include the Coulomb exchange, kinetic exchange, and superexchange [34].31 These interactions have the Heisenberg form

z z + − − + Sj · Sk = Sj Sk + Sj Sk + Sj Sk . (D1)

th ± 1 x The j single-nucleus spin operator is denoted by Sj. Raising and lowering operators are denoted by Sj := 2 (Sj ± y iSj ). Suppose that arbitrary phosphorus nuclear spins in a Posner interact via Eq. (D1):

6 X X Hint = Jjk Sj · Sk . (D2) j=1 k

The pair-dependent interaction strength is denoted by Jjk. This Hint remains invariant under permutations of the spins via C. C represents the rotation that preserves the Posner’s geometry. Hence the Posner’s intrinsic Hamiltonian might contain Hint. z The first term in Eq. (D1) conserves each spin’s Sj . The second term does not. But suppose that any spin flips + − upward via Sj . Another spin flips downward via Sk . The compensation preserves the total spin’s zin-component.

zin D 2 Why collisions with other molecules are expected to conserve a Posner molecule’s S1...6

− The Posner Hilbert space Hno-coll. has the computational basis specified by Eq. (10). Each basis element depends on a collective coordinate φ. If one spin makes an angle φ with the xin-axis (Fig. 1b), the other spins make angles φ + 2π/3 and φ + 4π/3. But spins are quantum objects. The first spin need not localize at any particular φ-value. Rather, the spin can delocalize across multiple φ-values, via a superposition. Delocalizing the state (10) yields Z dφ Ψ(φ) |(m1, r1)(m2, r2)(m3, r3); (m4, r4)(m5, r5)(m6, r6)i . (D3)

The kets depend on φ only through the rj’s. The coefficients Ψ(φ) ∈ C satisfy the normalization condition R dφ |Ψ(φ)|2 = 1.32 The state (D3) reduces to the basis element (10) when Ψ(φ) equals a Dirac delta function. A more general state of the Posner’s phosphorus nuclei33 has the form

Z X |ΨPosi = dφ Ψ(φ) |(m1, r1)(m2, r2)(m3, r3); (m4, r4)(m5, r5)(m6, r6)i . (D4)

((m1,r1)...(m6,r6))

The coefficients Ψ(φ) are wave functions, relative to the position basis, that represent an orbital DOF’s quantum state. Molecular collisions are expected to alter the Ψ(φ)’s: Jostling may change how tightly a spin is localized. Jostling is not expected to change the directions in which the spins point (since the spins are expected to have long zin coherence times). Hence molecular collisions are expected to preserve S1...6.

31 Superexchange within Posners is expected to be detailed in [9], We cleave to the kets prevalent in QI theory. according to [6]. 33 32 We continue to suppress the molecule’s coordinates, and most F&R introduce states of the form (D3) [1, 5]. They use second components of the molecule’s orientation, relative to the lab quantization and discuss transformation properties of the Ψ(φ)’s. frame. 35

− Appendix E DECOMPOSITION OF THE POSNER-MOLECULE HILBERT SPACE Hno-coll. IN TERMS OF COMPOSITE SPIN OPERATORS

34 A Posner encodes logical qubits (Sec. III). Three qubits correspond to the h+ triangle in Fig. 1a, via Eq. (10). We label these qubits 1, 2, and 3. The h− triangle corresponds to logical qubits 4, 5, and 6. The 64-dimensional logical space decomposes into a direct sum of subspaces. Different subspaces transform in different ways under 2 2 S123 + S456, the composite spin-squared operator (13). Let us derive the decomposition. We refer readers to standard quantum-mechanics textbooks, such as [66], for background. Let us focus on one triangle (one trio of qubits) first. Each trio corresponds to a Hilbert space C2 ⊗ C2 ⊗ C2. Each factor is replaced with the corresponding subsystem’s spin quantum number, in useful conventional notation: 1 1 1 s1 ⊗ s2 ⊗ s3 = 2 ⊗ 2 ⊗ 2 . This tensor product can be rewritten as a direct sum. To derive the direct sum, we follow rules for adding angular-momentum quantum numbers. Two spin quantum numbers, s1 and s2, sum as

stot = |s1 − s2|, |s1 − s2| + 1, . . . s1 + s2 − 1, s1 + s2 . (E1)

Two magnetic spin quantum numbers, m1 and m2, sum as

mtot = m1 + m2 . (E2)

We need not use Eq. (E2) here, however. Since tensor products distribute across direct sums, 1 1 1 1 ⊗ ⊗ = (0 ⊕ 1) ⊗ (E3) 2 2 2 2 1 1 3 = ⊕ ⊕ . (E4) 2 2 2

We can check Eq. (E4): A space that transforms with spin quantum number s has dimensionality 2s + 1. That is, s corresponds to 2s + 1 possible magnetic spin quantum numbers m. According to the LHS of Eq. (E3), therefore, 1 3 3 each triangle corresponds to a space of dimensionality 2 × 2 + 1 = 2 = 8. Equation (E4) implies the same dimensionality: 2 + 2 + 4 = 8. 1 1 3 ⊗2 Each Posner consists of two triangles. A triangle pair corresponds to the Hilbert space 2 ⊕ 2 ⊕ 2 . Distributing the tensor product across the direct sums yields

1 1 3⊗2 1 1⊕4 1 3⊕4 3 3 ⊕ ⊕ = ⊗ ⊕ ⊗ ⊕ ⊗ (E5) 2 2 2 2 2 2 2 2 2 = (0 ⊕ 1)⊕4 ⊕ (1 ⊕ 2)⊕4 ⊕ (0 ⊕ 1 ⊕ 2 ⊕ 3) (E6) = 0⊕5 ⊕ 1⊕9 ⊕ 2⊕5 ⊕ 3 . (E7)

Let us check Eq. (E7). According to the LHS of Eq. (E5), a Posner corresponds to a space of dimensionality (2 + 2 + 4)2 = 82 = 64. Equation (E7) implies the same dimensionality: 5 + (3 × 9) + (5 × 5) + 7 = 64.

Appendix F PREFERRED EIGENBASIS OF THE PERMUTATION OPERATOR C

The permutation operator C was introduced in Sec. IV A. The Posner dynamics is assumed to conserve C, as well zin zin as S1...6. An eigenbasis shared by C and S1...6 can facilitate the construction of natural quantum error-correcting codes (Sec. IV D). zin 2 2 Several eigenbases of C are eigenbases of S1...6. The operator S123 ⊗ S456 breaks the degeneracy satisfactorily, as zin 2 2 discussed in Sections IV C and VII. C, S1...6, and S123 ⊗ S456 share the eigenbasis in Tables II, III, and IV. Each table corresponds to one value of τ = 0, ±1 (equivalently, τ = 0, 1, 2).

34 No particular nucleus can be associated with any particular pure occupies the spin state |mj i if and only if A occupies the position spin-and-position state, by Pauli’s principle. But a spin can be state |rj i. associated with a position. Loosely speaking, some nucleus A 36

State s123 ⊗ s456 m123 m456 m1...6 τ123 ⊗ τ456 Decomposition 1 3 3 3 3 |cτ=0i 2 ⊗ 2 2 2 3 1 ⊗ 1 |000i|000i 2 3 3 3 1 |cτ=0i 2 ⊗ 2 2 2 2 1 ⊗ 1 |000i|W i 3 3 3 3 1 ¯ |cτ=0i 2 ⊗ 2 2 − 2 1 1 ⊗ 1 |000i|W i 4 3 3 3 3 |cτ=0i 2 ⊗ 2 2 − 2 0 1 ⊗ 1 |000i|111i 5 3 3 1 3 |cτ=0i 2 ⊗ 2 2 2 2 1 ⊗ 1 |W i|000i 6 3 3 1 1 |cτ=0i 2 ⊗ 2 2 2 1 1 ⊗ 1 |W i|W i 7 3 3 1 1 ¯ |cτ=0i 2 ⊗ 2 2 − 2 0 1 ⊗ 1 |W i|W i 8 3 3 1 3 |cτ=0i 2 ⊗ 2 2 − 2 -1 1 ⊗ 1 |W i|111i 9 3 3 1 3 ¯ |cτ=0i 2 ⊗ 2 − 2 2 1 1 ⊗ 1 |W i|000i 10 3 3 1 1 ¯ |cτ=0i 2 ⊗ 2 − 2 2 0 1 ⊗ 1 |W i|W i 11 3 3 1 1 ¯ ¯ |cτ=0i 2 ⊗ 2 − 2 − 2 -1 1 ⊗ 1 |W i|W i 12 3 3 1 3 ¯ |cτ=0i 2 ⊗ 2 − 2 − 2 -2 1 ⊗ 1 |W i|111i 13 3 3 3 3 |cτ=0i 2 ⊗ 2 − 2 2 0 1 ⊗ 1 |111i|000i 14 3 3 3 1 |cτ=0i 2 ⊗ 2 − 2 2 -1 1 ⊗ 1 |111i|W i 15 3 3 3 1 ¯ |cτ=0i 2 ⊗ 2 − 2 − 2 -2 1 ⊗ 1 |111i|W i 16 3 3 3 3 |cτ=0i 2 ⊗ 2 − 2 − 2 -3 1 ⊗ 1 |111i|111i 17 1 1 1 1 2 2 |cτ=0i 2 ⊗ 2 2 2 1 ω ⊗ ω |ωi|ω i 18 1 1 1 1 2 2 |cτ=0i 2 ⊗ 2 2 − 2 0 ω ⊗ ω |ωi|ω i 19 1 1 1 1 2 2 |cτ=0i 2 ⊗ 2 2 2 1 ω ⊗ ω |ω i|ωi 20 1 1 1 1 2 2 |cτ=0i 2 ⊗ 2 2 − 2 0 ω ⊗ ω |ω i|ω¯i 21 1 1 1 1 2 2 |cτ=0i 2 ⊗ 2 − 2 2 0 ω ⊗ ω |ω¯i|ω i 22 1 1 1 1 2 2 |cτ=0i 2 ⊗ 2 − 2 − 2 -1 ω ⊗ ω |ω¯i|ω i 23 1 1 1 1 2 2 |cτ=0i 2 ⊗ 2 − 2 2 0 ω ⊗ ω |ω i|ωi 24 1 1 1 1 2 2 |cτ=0i 2 ⊗ 2 − 2 − 2 -1 ω ⊗ ω |ω i|ω¯i

TABLE II: Preferred eigenbasis for the τ = 0 eigenspace of the permutation operator C: Twenty-four states span the eigenspace. Each basis element equals a product of two three-qubit states. The final column displays the product, explained in Sec. IV C. The state’s first factor represents a state of the qubits (labeled j = 1, 2, 3) in the top triangle in Fig. 1a. The second factor represents a state of the qubits (labeled j = 4, 5, 6) in the bottom triangle. Each factor is an 2 zin 2 zin eigenstate shared by the total-spin operators S123 and S123 or by S456 and S456. The operators are defined in Sec. IV C. 2 Table I displays the three-qubit eigenstates. The spin quantum number s123 denotes the eigenvalue of S123. The magnetic zin spin quantum number m123 denotes the eigenvalue of S123. s456 and m456 are defined analogously. The total magnetic spin 2 2 quantum number m1...6 = m123 + m456. The notation in column two follows from [66]: Eigenspaces of S123 ⊗ S456 bear the label s123 ⊗ s456. Column six is notated similarly. τ123 denotes the eigenvalue of the permutation operator that cyclically permutes qubits 1, 2, and 3. τ456 is defined analogously. The permutation eigenvalues multiply to τ123 × τ456 = τ.

Appendix G QUANTIFICATION OF THE INFORMATION ENCODED IN THE OUTCOME OF A POSNER-BINDING MEASUREMENT: ANALYSIS 2

The Posner-binding measurement is analyzed in Sec. VI. The measurement yields an outcome that encodes classical information. This information is quantified in Sec. VI A. The quantification is explained alternatively here. Imagine wishing to measure the τA and τB of Posners A and B. Each measurement would yield one of three possible outcomes (0, 1, or 2). The pair of measurements would yield one of nine possible outcomes. The pair of outcomes could be recorded in dlog2(9)e = 4 bits. Whether two Posners bind implements a measurement of whether τA + τB = 0. The yes-or-no answer constitutes one bit. You forfeit three of the bits you wanted, measuring just whether the Posners bind. Three is the number of bits you would need to specify the value of (τA, τB), given that τA + τB =6 0. Why? Suppose that τA + τB =6 0. 37

State s123 ⊗ s456 m123 m456 m1...6 τ123 ⊗ τ456 Decomposition 1 3 1 3 1 |cτ=1i 2 ⊗ 2 2 2 2 1 ⊗ ω |000i|ωi 2 3 1 3 1 |cτ=1i 2 ⊗ 2 2 − 2 1 1 ⊗ ω |000i|ω¯i 3 3 1 1 1 |cτ=1i 2 ⊗ 2 2 2 1 1 ⊗ ω |W i|ωi 4 3 1 1 1 |cτ=1i 2 ⊗ 2 2 − 2 0 1 ⊗ ω |W i|ω¯i 5 3 1 1 1 ¯ |cτ=1i 2 ⊗ 2 − 2 2 0 1 ⊗ ω |W i|ωi 6 3 1 1 1 ¯ |cτ=1i 2 ⊗ 2 − 2 − 2 -1 1 ⊗ ω |W i|ω¯i 7 3 1 3 1 |cτ=1i 2 ⊗ 2 − 2 2 -1 1 ⊗ ω |111i|ωi 8 3 1 3 1 |cτ=1i 2 ⊗ 2 − 2 − 2 -2 1 ⊗ ω |111i|ω¯i 9 1 3 1 3 |cτ=1i 2 ⊗ 2 2 2 2 ω ⊗ 1 |ωi|000i 10 1 3 1 1 |cτ=1i 2 ⊗ 2 2 2 1 ω ⊗ 1 |ωi|W i 11 1 3 1 1 ¯ |cτ=1i 2 ⊗ 2 2 − 2 0 ω ⊗ 1 |ωi|W i 12 1 3 1 3 |cτ=1i 2 ⊗ 2 2 − 2 -1 ω ⊗ 1 |ωi|111i 13 1 1 1 1 2 2 2 2 |cτ=1i 2 ⊗ 2 2 2 1 ω ⊗ ω |ω i|ω i 14 1 1 1 1 2 2 2 2 |cτ=1i 2 ⊗ 2 2 − 2 0 ω ⊗ ω |ω i|ω i 15 1 3 1 3 |cτ=1i 2 ⊗ 2 − 2 2 1 ω ⊗ 1 |ω¯i|000i 16 1 3 1 1 |cτ=1i 2 ⊗ 2 − 2 2 0 ω ⊗ 1 |ω¯i|W i 17 1 3 1 1 ¯ |cτ=1i 2 ⊗ 2 − 2 − 2 -1 ω ⊗ 1 |ω¯i|W i 18 1 3 1 3 |cτ=1i 2 ⊗ 2 − 2 − 2 -2 ω ⊗ 1 |ω¯i|111i 19 1 1 1 1 2 2 2 2 |cτ=1i 2 ⊗ 2 − 2 2 0 ω ⊗ ω |ω i|ω i 20 1 1 1 1 2 2 2 2 |cτ=1i 2 ⊗ 2 − 2 − 2 -1 ω ⊗ ω |ω i|ω i

TABLE III: Preferred eigenbasis for the τ = 1 eigenspace of the rotation symmetry operator C: The notation is defined below Table II.

(τA, τB) can equal one of six possible values, (0, 1), (0, 2), (1, 0), (1, 1), (2, 0), or (2, 2). Specifying one of six possible 35 values requires dlog2(6)e = 3 bits. Hence measuring Posner binding is equivalent to each of two QI processes:

1. Measuring (τA, τB) and coarse-graining away three bits (all information except whether τA + τB = 0).

2. Measuring the Bell basis and coarse-graining away one bit (whether a + outcome or a − outcome occurred).

Appendix H HOW TO PREPARE, WITH POSNER OPERATIONS, STATES USED IN INCOHERENT TELEPORTATION

Section VI B details how Posners can teleport QI incoherently. The protocol involves states |+ i = √1 (|0 i + τ 3 τ |1τ i + |2τ i) and |ψi = c0|0τ i + c1|1τ i + c2|2τ i. Each |jτ i denotes an arbitrary state in the τ = j subspace. How can Posner operations (Sec. V A) prepare a |+τ i and a |ψi? One protocol is described below. Other protocols may await discovery. Each state is of one Posner and is pure. Hence the Posner contains three singlets. Consider preparing three singlets via operation 1. Consider rotating one spin via operation 2. Let the rotation be about the ylab-axis, through an angle θ.

35 Imagine learning, instead, that τA +τB = 0. Given this informa- would need dlog2(3)e = 2 bits. But you could encode the tuple’s tion, would you need three bits to specify the value of (τA, τB )? value in three bits. No: (τA, τB ) can assume one of three possible values. Hence you 38

State s123 ⊗ s456 m123 m456 m1...6 τ123 ⊗ τ456 Decomposition 1 3 1 3 1 2 2 |cτ=−1i 2 ⊗ 2 2 2 2 1 ⊗ ω |000i|ω i 2 3 1 3 1 2 2 |cτ=−1i 2 ⊗ 2 2 − 2 1 1 ⊗ ω |000i|ω i 3 3 1 1 1 2 2 |cτ=−1i 2 ⊗ 2 2 2 1 1 ⊗ ω |W i|ω i 4 3 1 1 1 2 2 |cτ=−1i 2 ⊗ 2 2 − 2 0 1 ⊗ ω |W i|ω i 5 3 1 1 1 2 ¯ 2 |cτ=−1i 2 ⊗ 2 − 2 2 0 1 ⊗ ω |W i|ω i 6 3 1 1 1 2 ¯ 2 |cτ=−1i 2 ⊗ 2 − 2 − 2 −1 1 ⊗ ω |W i|ω i 7 3 1 3 1 2 2 |cτ=−1i 2 ⊗ 2 − 2 2 −1 1 ⊗ ω |111i|ω i 8 3 1 3 1 2 2 |cτ=−1i 2 ⊗ 2 − 2 − 2 −2 1 ⊗ ω |111i|ω i 9 1 1 1 1 |cτ=−1i 2 ⊗ 2 2 2 1 ω ⊗ ω |ωi|ωi 10 1 1 1 1 |cτ=−1i 2 ⊗ 2 2 − 2 0 ω ⊗ ω |ωi|ω¯i 11 1 3 1 3 2 2 |cτ=−1i 2 ⊗ 2 2 2 2 ω ⊗ 1 |ω i|000i 12 1 3 1 1 2 2 |cτ=−1i 2 ⊗ 2 2 2 1 ω ⊗ 1 |ω i|W i 13 1 3 1 1 2 2 ¯ |cτ=−1i 2 ⊗ 2 2 − 2 0 ω ⊗ 1 |ω i|W i 14 1 3 1 3 2 2 |cτ=−1i 2 ⊗ 2 2 − 2 −1 ω ⊗ 1 |ω i|111i 15 1 1 1 1 |cτ=−1i 2 ⊗ 2 − 2 2 0 ω ⊗ ω |ω¯i|ωi 16 1 1 1 1 |cτ=−1i 2 ⊗ 2 − 2 − 2 −1 ω ⊗ ω |ω¯i|ω¯i 17 1 3 1 3 2 2 |cτ=−1i 2 ⊗ 2 − 2 2 1 ω ⊗ 1 |ω i|000i 18 1 3 1 1 2 2 |cτ=−1i 2 ⊗ 2 − 2 2 0 ω ⊗ 1 |ω i|W i 19 1 3 1 1 2 2 ¯ |cτ=−1i 2 ⊗ 2 − 2 − 2 −1 ω ⊗ 1 |ω i|W i 20 1 3 1 3 2 2 |cτ=−1i 2 ⊗ 2 − 2 − 2 −2 ω ⊗ 1 |ω i|111i

TABLE IV: Preferred eigenbasis for the τ = −1 eigenspace (equivalently, the τ = 2 eigenspace) of the rotation symmetry operator C: The notation is defined below Table II.

Consider forming a Posner from the spins, via operation 3. Let the singlets be arranged as in Fig. 19. Recall that a Posner contains two triangles of phosphorus nuclear spins (Sec. II B). One triangle sits at zin = h+; and the other triangle, at zin = h−. Each triangle contains one singlet (illustrated with a green, wavy line). One singlet extends from the h+ triangle to the h− triangle. (How a singlet corresponds to positions in a Posner is discussed in Sec. III C and App. B.) The red hoop encircles the rotated spin. The rotated spin is entangled with a spin in the same triangle. Let |φ(θ)i denote the Posner’s state. |φ(θ)i can have weight on each τ = j eigenspace:

2 dj X X λj |φ(θ)i = Cj,λj (θ)|cτ=ji . (H1) j=0 λj =1

The τ = j eigenspace has degeneracy dj. The degeneracy parameter is denoted by λj. The coefficients Cj,λj (θ) satisfy P P 2 the normalization condition j λj Cj,λj (θ) = 1. The dependence on θ can be calculated analytically: The state has an amount

24 X 2 2 |C0,λ0 (θ)| = cos θ (H2)

λ0=1 of weight on the τ = 0 eigenspace, an amount

20 X 2 1 2 |C ,λ (θ)| = sin θ (H3) 1 1 2 λ1=1 39

zˆin

h+

h-

FIG. 19: Posner-molecule state usable in incoherent teleportation: Each black dot represents a phosphorus nuclear spin. The internal z-axisz ˆin remains fixed with respect to the atoms’ positions. Three spins sit at zin = h+; and three spins, at zin = h−. The spins occupy a pure state of three singlets. Each green, wavy line represents one singlet. The red hoop encircles a spin that has been rotated through an angle θ. The rotation is about the ylab-axis, which remains fixed relative to the lab that contains the Posner. The angle labels the Posner’s state, |φ(θ)i. Instances of |φ(θ)i can serve as the |+τ i and the |ψi in incoherent teleportation (Sec. VI B 2). on the τ = 1 eigenspace, and an amount 20 X 2 1 2 |C ,λ (θ)| = sin θ (H4) 2 2 2 λ2=1 on the τ = 2 eigenspace. At which θ-value does the weight on each eigenspace√ equal the weight on every other? Let us equate (H2), (H3), and (H4). Solving for the angle yields θ = tan−1( 2). The corresponding state can serve as the equal-weight superposition |+τ i: √ −1 |+τ i = |φ(tan ( 2))i . (H5)

The basis vectors |jτ i inherit the definition dj √ X −1 λj |jτ i = Cj,λj (tan ( 2))|cj i . (H6) λj =1 Now, let θ assume an arbitrary value. Information about |φ(θ)i can be teleported incoherently: |ψi = |φ(θ)i . (H7) P2 Granted, |φ(θ)i might not decompose as j=0 cj|jτ i, in terms of the |jτ i’s defined in Eq. (H6). Yet the incoherent- teleportation protocol continues to work: Equation (H7) defines new basis elements |jτ (θ)i:

dj X λj |jτ (θ)i = Cj,λj (θ)|cj i . (H8) λj =1

States |jτ i of Posner A appear in Eqs. (47) and (50). Each such |jτ i must be replaced with a |jτ (θ)i. The projector ΠAB transforms the |jτ (θ)i’s as it would transform the |jτ i’s.

Appendix I FRUSTRATED-LATTICE INTUITION ABOUT PROJECTING ONTO THE τ = 0 SUBSPACE

We can understand Eq. (61) in terms of a frustrated lattice. Consider a triangular lattice of three sites, A, B, and C. Let a spin-1 DOF occupy each site. The site-K magnetic spin quantum number mK = 0, ±1 stands in place of τK . 40

Let us regress to Eq. (60). We ignore the final m − 3 identity operators in each term. How does Π123 transform the lattice’s state? Consider multiplying out the terms in the RHS. We label as a cross-term each term that contains at least one ΠτK =0 and one ΠτK =±1, for some K = A, B, C. These projectors annihilate each other; the cross-terms vanish. Each surviving term in Π123 contains only τK = 0 projectors or only τK = ±1 projectors. Each τK = ±1 projector represents an antiferromagnetic interaction between two lattice sites. The τK = ±1 projectors form a term that represents a frustrated lattice. No set (τA, τB, τC ) satisfies all the constraints encoded in the frustration term. Hence the lattice must occupy its τA = τB = τC = 0 subspace.

0 Appendix J PEPS REPRESENTATION OF |AKLThoni

The AKLT0 PEPS is a repeating pattern of two tensors, T + and T − (Fig. 17). We will focus primarily on T +. The + + + + + + tensor has six indices. Three (v1 , v2 , and v3 ) are virtual. Three more indices (a1 , a2 , and a3 ) are physical. Each small, black dot represents a virtual spin. Each short leg, extending upward from the plane occupied by the + + + large circle, represents a physical qubit. We denote the physical qubits’ computational-basis states by |a1 a2 a3 i. For + each j = 1, 2, 3, the physical index aj = 0, 1. + + Each long leg, extended across the plane occupied by the large circle, represents a virtual index. The v1 and v2 + lines represents singlets. Consider, as an illustration, the physical qubit associated with a1 . This qubit forms a singlet + + with some physical qubit in another tensor. Suppose that a1 = 0. The T physical qubit points upward. Hence the partner physical qubit must point downward: The partner qubit’s a must equal one. This necessity is conveyed to + + + + the second tensor by the virtual index v1 : If a1 = 0, T + + + + + + =6 0 only if v1 = 0. a1 ,a2 ,a3 ,v1 ,v2 ,v3 + + + + The virtual index v3 differs from v1 and v2 : The tensor lacks isotropy. v3 connects two tensors associated with the same Posner, T + and T −. The two tensors, together, determine which C eigenspace the Posner occupies. Hence + + + v3 carries not only “singlet” information about one physical qubit. v3 conveys also how the T qubit trio transforms − under C3 (the final column in Table I). This C information dictates how the T physical qubits must transform, such that the Posner occupies the τ = 0 eigenspace. + + + + We ascribe to v3 a tuple (˜v3 , τ ). The first entry conveys information about the a3 physical qubit. Only if + + + + v˜3 = a3 can the tensor have a nonzero value. The second entry, τ , equals 0, 1, or 2. Hence v3 assumes one of six possible values:

+ + + v3 = (˜v3 , τ ) = {(0, 0), (0, 1), (0, 2), (1, 0), (1, 1), (1, 2)} (J1) = {0, 1, 2, 3, 4, 5} . (J2)

+ Hence v3 has a bond dimension of six. + Having overviewed the tensor’s six indices, we consider the whole tensor, T + + + + + + . This tensor equals the a1 ,a2 ,a3 ,v1 ,v2 ,v3 + + + + + coefficient that multiplies the computational-basis state |a1 , a2 , a3 i when the virtual indices have the values v1 , v2 , + + and v3 . Suppose, for simplicity, that the T triangle lacked connections to any other triangles. The triangle would occupy the physical state

X + + + + (const.) T + + + + + + |a1 , a2 , a3 i . (J3) a1 ,a2 ,a3 ,v1 ,v2 ,v3 + + + + + + a1 ,a2 ,a3 ,v1 ,v2 ,v3 The v’s do not label the ket, because they are virtual. The tensor can be evaluated, with help from Table I, after a normalization convention is chosen. We illustrate with three examples. + + + + + First, let us evaluate T000000. Since v3 = 0, Eq. (J2) implies that T000000 can =6 0 only if a3 = 0. Indeed, a3 = 0. + + + In fact, every a vanishes. This tensor equals the coefficient of the one-triangle state |a1 a2 a3 i = |000i. This state occupies the τ = 0 eigenspace, according to Table I. We choose the following normalization condition: |000i appears + once, with a unit coefficient, in the table’s second column. Hence we choose for T000000 to equal one. + The second example consists of T + + + , wherein the a’s have arbitrary values. According to the final three a1 a2 a3 001 + + indices [and Eq. (J2)], the tensor can be nonzero only if aj = 0 for all j = 1, 2, 3. That is, T + + + = 0 except, a1 a2 a3 001 + + + perhaps, if the coefficient of |a1 a2 a3 i = |000i. + + The tensor’s final index implies that v3 = 1. Hence, by Eq. (J2), the qubit trio transforms under C with τ = 1. No qubit-trio state (i) transforms with τ + and (ii) equals a linear combination of computational-basis states including + |000i, by Table (I). Hence T000001 = 0. + + + + + The final example consists of T100100. The physical indices “agree with” the virtue indices: a1 = v1 , a2 = v2 , and + + [by Eq. (J2)] a3 =v ˜3 . Hence the tensor does not necessarily vanish. This tensor multiplies the physical one-triangle 41

+ + + ket |a1 a2 a3 i = |100i. This ket appears three times in the second column of Table I. Only one of those appearances + is relevant: Since v3 = 0, Eq. (J2) implies that τ+ = 0. Hence the physical qubit trio occupies the first ket in the |W i decomposition (in the third row of Table I). This ket multiples a √1 in the table’s second column. We might wish to 3 ascribe the value √1 to T + . 3 100100 But the physical qubits’ state is constructed from singlets. Singlets carry minus signs. We must incorporate these a+ minus signs into our convention. We choose for the tensor to carry a factor of (−1) j for each j = 1, 2, 3. Hence T + = − √1 . 100100 3

REFERENCES

[1] M. P. A. Fisher, Annals of Physics 362, 593 (2015). [2] G. Treboux, P. Layrolle, N. Kanzaki, K. Onuma, and A. Ito, The Journal of Physical Chemistry A 104, 5111 (2000), http://dx.doi.org/10.1021/jp994399t. [3] N. Kanzaki, G. Treboux, K. Onuma, S. Tsutsumi, and A. Ito, Biomaterials 22, 2921 (2001). [4] X. Yin and M. Stott, J. Chem. Phys. 118, 3717 (2003). [5] M. P. A. Fisher and L. Radzihovsky, ArXiv e-prints (2017), 1707.05320. [6] M. P. A. Fisher, private communication. [7] M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information (Cambridge University Press, 2010). [8] J. Preskill, Ch. 7: Quantum error correction, in Phys. 219: Quantum Computation and Information, Lecture Notes, 1999. [9] M. Swift, C. Van de Walle, and M. Fisher, (in prep). [10] C. H. Bennett et al., Phys. Rev. Lett. 70, 1895 (1993). [11] C. H. Bennett and S. J. Wiesner, Phys. Rev. Lett. 69, 2881 (1992). [12] H. J. Briegel and R. Raussendorf, Phys. Rev. Lett. 86, 910 (2001). [13] R. Raussendorf, D. E. Browne, and H. J. Briegel, Phys. Rev. A 68, 022312 (2003). [14] H. J. Briegel, D. E. Browne, W. Dur, R. Raussendorf, and M. Van den Nest, Nat Phys , 19 (2009). [15] D. Deutsch, Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 425, 73 (1989), http://rspa.royalsocietypublishing.org/content/425/1868/73.full.pdf. [16] D. Gross, S. T. Flammia, and J. Eisert, Phys. Rev. Lett. 102, 190501 (2009). [17] R. Raussendorf and H. J. Briegel, Phys. Rev. Lett. 86, 5188 (2001). [18] M. Hein et al., Entanglement in Graph States and its Applications, in Proceedings of the International School of Physics ”Enrico Fermi” on ”Quantum Computers, Algorithms and Chaos”, 2006. [19] M. Van den Nest, A. Miyake, W. D¨ur,and H. J. Briegel, Phys. Rev. Lett. 97, 150504 (2006). [20] A. Miyake, Annals of Physics 326, 1656 (2011), July 2011 Special Issue. [21] I. Affleck, T. Kennedy, E. H. Lieb, and H. Tasaki, Phys. Rev. Lett. 59, 799 (1987). [22] I. Affleck, T. Kennedy, E. H. Lieb, and H. Tasaki, Communications in Mathematical Physics 115, 477 (1988). [23] T. Kennedy, E. H. Lieb, and B. S. Shastry, Journal of Statistical Physics 53, 1019 (1988). [24] T.-C. Wei, I. Affleck, and R. Raussendorf, Phys. Rev. A 86, 032328 (2012). [25] F. Verstraete, V. Murg, and J. Cirac, Advances in Physics 57, 143 (2008), http://dx.doi.org/10.1080/14789940801912366. [26] D. Perez-Garcia, F. Verstraete, J. I. Cirac, and M. M. Wolf, ArXiv e-prints (2007), 0707.2260. [27] A. Molnar, Y. Ge, N. Schuch, and J. I. Cirac, ArXiv e-prints (2017), 1706.07329. [28] D. P. Divincenzo, Fortschritte der Physik 48, 771 (2000), quant-ph/0002077. [29] K. Onuma and A. Ito, Chemistry of Materials 10, 3346 (1998), http://dx.doi.org/10.1021/cm980062c. [30] A. Oyane, K. Onuma, T. Kokubo, and I. Atsuo, The Journal of Physical Chemistry B 103, 8230 (1999), http://dx.doi.org/10.1021/jp9910340. [31] A. Dey et al., Nat. Mat. 9, 1010 (2010). [32] D. C. Rapaport, The Art of Molecular Dynamics Simulation, 2 ed. (Cambridge University Press, 2004). [33] B. Misra and E. C. G. Sudarshan, Journal of Mathematical Physics 18, 756 (1977), http://dx.doi.org/10.1063/1.523304. [34] N. W. Ashcroft and N. D. Mermin, Solid State Physics (Brooks/Cole, Belmont, CA, 1976). [35] C. Cohen-Tannoudji, B. Diu, and F. Lalo¨e, Quantum Mechanics (Wiley, 1991). [36] E. Knill and R. Laflamme, Phys. Rev. A 55, 900 (1997). [37] C. H. Bennett, D. P. DiVincenzo, J. A. Smolin, and W. K. Wootters, Phys. Rev. A 54, 3824 (1996). [38] D. Kribs, R. Laflamme, and D. Poulin, Phys. Rev. Lett. 94, 180501 (2005). [39] D. W. Kribs, R. Laflamme, D. Poulin, and M. Lesosky, eprint arXiv:quant-ph/0504189 (2005), quant-ph/0504189. [40] M. A. Nielsen and D. Poulin, Phys. Rev. A 75, 064304 (2007). [41] J. Watrous, Ch. 16: Quantum error correction, in CPSC 519/619: Quantum Computation, Lecture Notes, 2006. [42] R. W. Spekkens, Phys. Rev. Lett. 101, 020401 (2008). [43] M. Howard, J. Wallman, V. Veitch, and J. Emerson, Nature 510, 351 (2014). [44] and and, Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 449, 669 (1995), http://rspa.royalsocietypublishing.org/content/449/1937/669.full.pdf. [45] S. Lloyd, Phys. Rev. Lett. 75, 346 (1995). [46] J.-L. Brylinski and R. Brylinski, eprint arXiv:quant-ph/0108062 (2001), quant-ph/0108062. 42

[47] M. Freedman, A. Kitaev, and J. Lurie, eprint arXiv:quant-ph/0209113 (2002), quant-ph/0209113. [48] A. W. Harrow, ArXiv e-prints (2008), 0806.0631. [49] L. Eldar and A. W. Harrow, ArXiv e-prints (2015), 1510.02082. [50] S. Aaronson, Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 461, 3473 (2005), http://rspa.royalsocietypublishing.org/content/461/2063/3473.full.pdf. [51] R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki, Rev. Mod. Phys. 81, 865 (2009). [52] D. Gottesman and I. L. Chuang, Nature 402, 390 (1999). [53] E. Knill, R. Laflamme, and G. J. Milburn, Nature 409, 46 (2001). [54] M. A. Nielsen, Physics Letters A 308, 96 (2003). [55] X. Zhou, D. W. Leung, and I. L. Chuang, Phys. Rev. A 62, 052316 (2000). [56] D. Gross and J. Eisert, Phys. Rev. Lett. 98, 220503 (2007). [57] D. Gross, J. Eisert, N. Schuch, and D. Perez-Garcia, Phys. Rev. A 76, 052315 (2007). [58] F. Verstraete and J. I. Cirac, Phys. Rev. A 70, 060302 (2004). [59]S. Ostlund¨ and S. Rommer, Phys. Rev. Lett. 75, 3537 (1995). [60] M. Fannes, B. Nachtergaele, and R. F. Werner, Comm. Math. Phys. 144, 443 (1992). [61] A. Kl¨umper, A. Schadschneider, and J. Zittartz, Journal of Physics A: Mathematical and General 24, L955 (1991). [62] F. Verstraete and J. I. Cirac, eprint arXiv:cond-mat/0407066 (2004), cond-mat/0407066. [63] G. K. Brennen and A. Miyake, Phys. Rev. Lett. 101, 010502 (2008). [64] A. Miyake, Phys. Rev. Lett. 105, 040501 (2010). [65] S. D. Bartlett, G. K. Brennen, A. Miyake, and J. M. Renes, Phys. Rev. Lett. 105, 110502 (2010). [66] R. Shankar, Principles of Quantum Mechanics (Plenum Press, New York, 1994). [67] R. Raussendorf, J. Harrington, and K. Goyal, Annals of Physics 321, 2242 (2006). [68] D. E. Browne et al., New Journal of Physics 10, 023010 (2008). [69] T.-C. Wei, P. Haghnegahdar, and R. Raussendorf, Phys. Rev. A 90, 042333 (2014). [70] T.-C. Wei and R. Raussendorf, Phys. Rev. A 92, 012310 (2015). [71] J. Preskill, Quantum computing and the entanglement frontier, in The Theory of the Quantum World: Proceedings of the 25th Solvay Conference on Physics, edited by D. Gross, M. Henneaux, and A. Sevrin, World Scientific Publishing, 2013. [72] E. Farhi and A. W. Harrow, ArXiv e-prints (2016), 1602.07674. [73] T. Nishino et al., Progress of Theoretical Physics 105, 409 (2001). [74] H. Niggemann, A. Kl¨umper, and J. Zittartz, Zeitschrift f¨urPhysik B Condensed Matter 104, 103 (1997). [75] F. Verstraete, M. A. Mart´ın-Delgado,and J. I. Cirac, Phys. Rev. Lett. 92, 087201 (2004). [76] A. S. Darmawan, G. K. Brennen, and S. D. Bartlett, New Journal of Physics 14, 013023 (2012). [77] M. P. A. Fisher, International Journal of Modern Physics B 31, 1743001 (2017), http://www.worldscientific.com/doi/pdf/10.1142/S0217979217430019. [78] A. Heiss, V. Pipich, W. Jahnen-Dechent, and D. Schwahn, Biophysical Journal 99, 3986 (2010). [79] E. Bellocchio, R. J. Reimer, R. T. Fremeau, and R. H. Edwards, Science 289, 957 (2000). [80] M. Liguz-Lecznar and J. Skangiel-Kramska, Acta Neurobiol Exp 67, 207 (2007). [81] S. Sreedharan et al., BMC Genomics 11, 17 (2010). [82] R. Reimer, Molecular Aspects of Medecine 34, 350 (2013). [83] B. Ni, P. R. Rosteck, N. S. Nadi, and S. M. Paul, Proc. Nod. Acad. Sci. 91, 5607 (1994). [84] R. Milo and R. Phillips, How big is a synapse?, in Cell Biology by the Numbers, Garland Science, 2015. [85] S. Jordan, Quantum algorithm zoo, Online resource. [86] C. Palazuelos and T. Vidick, Journal of Mathematical Physics 57, 015220 (2016). [87] J. F. Clauser, M. A. Horne, A. Shimony, and R. A. Holt, Phys. Rev. Lett. 23, 880 (1969). [88] J. Preskill, Ch. 4: , in Phys. 219: Quantum Computation and Information, Lecture Notes, 2001. [89] P. W. Shor, SIAM Journal on Computing 26, 1484 (1997), https://doi.org/10.1137/S0097539795293172. [90] E. Knill and R. Laflamme, Phys. Rev. Lett. 81, 5672 (1998). [91] D. Deutsch, Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 400, 97 (1985), http://rspa.royalsocietypublishing.org/content/400/1818/97.full.pdf. [92] E. Farhi et al., Science 292, 472 (2001), http://science.sciencemag.org/content/292/5516/472.full.pdf. [93] A. Kitaev, Annals of Physics 303, 2 (2003). [94] A. M. Childs, Phys. Rev. Lett. 102, 180501 (2009). [95] S. P. Jordan, Quantum Information and Computation 10, 470 (2010). [96] A. Y. Kitaev, Russian Math. Surveys 52, 1191 (1997). [97] P. W. Shor, Phys. Rev. A 52, R2493 (1995).