+ All Categories
Home > Documents > arXiv:1711.05899v1 [physics.chem-ph] 16 Nov 2017 we show that the Posner molecule provides an ideal...

arXiv:1711.05899v1 [physics.chem-ph] 16 Nov 2017 we show that the Posner molecule provides an ideal...

Date post: 30-May-2018
Category:
Upload: nguyentram
View: 332 times
Download: 36 times
Share this document with a friend
8
Posner molecules: From atomic structure to nuclear spins Michael W. Swift and Matthew P. A. Fisher Department of Physics, University of California, Santa Barbara, California 93106-9530, USA Chris G. Van de Walle Materials Department, University of California, Santa Barbara, California 93106-5050, USA We investigate “Posner molecules”, calcium phosphate clusters with chemical formula Ca9(PO4)6. Originally identified in hydroxyapatite, Posner molecules have also been observed as free-floating molecules in vitro. The formation and aggregation of Posner molecules have important implications for bone growth, and may also play a role in other biological processes such as the modulation of calcium and phosphate ion concentrations within the mitochondrial matrix. In this work, we use a first-principles computational methodology to study the structure of Posner molecules, their vibrational spectra, their interactions with other cations, and the process of pairwise bonding. Additionally, we show that the Posner molecule provides an ideal environment for the six constituent 31 P nuclear spins to obtain very long spin coherence times. In vitro, the spins could provide a platform for liquid-state nuclear magnetic resonance quantum computation. In vivo, the spins may have medical imaging applications. The spins have also been suggested as “neural qubits” in a proposed mechanism for quantum processing in the brain. I. INTRODUCTION In 1975 Betts and Posner, while examining the x- ray crystal structure of bone mineral—hydroxyapatite, Ca 10 (PO 4 ) 6 (OH) 2 —noticed that within each unit cell there were two calcium-phosphate “structural clusters” with atomic constituents Ca 9 (PO 4 ) 6 . 1 It was subse- quently argued that these so-called “Posner clusters” played an important role in the formation of amorphous calcium phosphate, which can be viewed as a “glass” of Posner clusters. It was over 20 years later that Onuma and Ito, while performing intensity-enhanced dynamic light scattering on simulated body fluids, identified evidence for free- floating calcium phosphate clusters of size roughly 1 nm. 2 They suggested that these clusters, apparently stable in solution for months or longer were, in fact, Posner clus- ters. No spontaneous precipitation was observed even in a supersaturated solution, but it was suggested that these Posner clusters play a role in bone (hydroxyap- atite) nucleation when presented with a seed crystal. 3,4 Additional evidence was provided in 2010 when cryo- genic transmission electron microscopy experiments on bone nucleation in simulated body fluids imaged free- floating nanometer-sized molecules which coalesced near a seed, forming amorphous calcium phosphate before un- dergoing a dramatic transition into the crystalline form of hydroxyapatite. 5 Further evidence for the structural integrity of these clusters in solution was found in AFM images of bone growth on calcite surfaces. 6 Taken to- gether, these remarkable experiments provide evidence that Posner clusters are stable in solution and, as such, should perhaps be called “Posner molecules”—the name we will henceforth adopt. Soon after the Onuma and Ito experiments, sev- eral quantum chemistry calculations examined the pu- tative arrangement of the ions in these Posner molecules, Ca 9 (PO 4 ) 6 , which were indeed found to be stable in vac- uum. 7,8 Their basic form consisted of eight calcium ions on the corners of a cube with the ninth calcium in the center, while the six phosphate ions were situated on the six faces of the cube. Due to the symmetry incompati- bility of the tetrahedral phosphate ions (PO 3- 4 ) with the cube faces, the cubic symmetry was broken in the low- energy molecular configurations. One of the most stable configurations was found to have S 6 symmetry, with a threefold rotational symmetry axis that is aligned along one of the cube diagonals, as depicted in Figure 1. A more detailed exploration of Posner molecules is worthwhile for multiple reasons: 1. Role in biological processes. Since experiments have provided strong evidence that Posner molecules are stable in simulated body fluids, it is likely that they are present in vivo, particularly in the extracellu- lar fluid where the free calcium concentration is appre- ciable. These extracelluar Posner molecules could act as a reservoir for the previously suggested Posner-molecule- mediated mechanism for bone growth. 3,4 The molecules may also be present within cells. Though cytoplasm is unlikely to contain Posner molecules due to its low cal- cium concentration, the mitochondiral matrix is known to contain stable calcium-phosphate complexes with a calcium:phosphate ratio suggestively close to that of Pos- ner molecules (3:2). 9 Posner molecules may form a dy- namic aggregate within the mitochondiral matrix, with formation, aggregation, and dis-aggregation being finely modulated by the changing pH. 10 The Posner molecule is also central to the proposed “quantum brain” concept set forth in Ref. 11. In this pro- posal, clouds of entangled Posner molecules in the brain act as “neural qubits”, serving as a platform for quantum computation in cognitive processes. Many of the proper- ties we discuss here are key to the ongoing exploration of this concept. 2. Long-lived phosphorus nuclear spin states. arXiv:1711.05899v1 [physics.chem-ph] 16 Nov 2017
Transcript

Posner molecules: From atomic structure to nuclear spins

Michael W. Swift and Matthew P. A. FisherDepartment of Physics, University of California, Santa Barbara, California 93106-9530, USA

Chris G. Van de WalleMaterials Department, University of California, Santa Barbara, California 93106-5050, USA

We investigate “Posner molecules”, calcium phosphate clusters with chemical formula Ca9(PO4)6.Originally identified in hydroxyapatite, Posner molecules have also been observed as free-floatingmolecules in vitro. The formation and aggregation of Posner molecules have important implicationsfor bone growth, and may also play a role in other biological processes such as the modulationof calcium and phosphate ion concentrations within the mitochondrial matrix. In this work, weuse a first-principles computational methodology to study the structure of Posner molecules, theirvibrational spectra, their interactions with other cations, and the process of pairwise bonding.Additionally, we show that the Posner molecule provides an ideal environment for the six constituent31P nuclear spins to obtain very long spin coherence times. In vitro, the spins could provide aplatform for liquid-state nuclear magnetic resonance quantum computation. In vivo, the spins mayhave medical imaging applications. The spins have also been suggested as “neural qubits” in aproposed mechanism for quantum processing in the brain.

I. INTRODUCTION

In 1975 Betts and Posner, while examining the x-ray crystal structure of bone mineral—hydroxyapatite,Ca10(PO4)6(OH)2—noticed that within each unit cellthere were two calcium-phosphate “structural clusters”with atomic constituents Ca9(PO4)6.1 It was subse-quently argued that these so-called “Posner clusters”played an important role in the formation of amorphouscalcium phosphate, which can be viewed as a “glass” ofPosner clusters.

It was over 20 years later that Onuma and Ito, whileperforming intensity-enhanced dynamic light scatteringon simulated body fluids, identified evidence for free-floating calcium phosphate clusters of size roughly 1 nm.2

They suggested that these clusters, apparently stable insolution for months or longer were, in fact, Posner clus-ters. No spontaneous precipitation was observed evenin a supersaturated solution, but it was suggested thatthese Posner clusters play a role in bone (hydroxyap-atite) nucleation when presented with a seed crystal.3,4

Additional evidence was provided in 2010 when cryo-genic transmission electron microscopy experiments onbone nucleation in simulated body fluids imaged free-floating nanometer-sized molecules which coalesced neara seed, forming amorphous calcium phosphate before un-dergoing a dramatic transition into the crystalline formof hydroxyapatite.5 Further evidence for the structuralintegrity of these clusters in solution was found in AFMimages of bone growth on calcite surfaces.6 Taken to-gether, these remarkable experiments provide evidencethat Posner clusters are stable in solution and, as such,should perhaps be called “Posner molecules”—the namewe will henceforth adopt.

Soon after the Onuma and Ito experiments, sev-eral quantum chemistry calculations examined the pu-tative arrangement of the ions in these Posner molecules,

Ca9(PO4)6, which were indeed found to be stable in vac-uum.7,8 Their basic form consisted of eight calcium ionson the corners of a cube with the ninth calcium in thecenter, while the six phosphate ions were situated on thesix faces of the cube. Due to the symmetry incompati-bility of the tetrahedral phosphate ions (PO3−

4 ) with thecube faces, the cubic symmetry was broken in the low-energy molecular configurations. One of the most stableconfigurations was found to have S6 symmetry, with athreefold rotational symmetry axis that is aligned alongone of the cube diagonals, as depicted in Figure 1.

A more detailed exploration of Posner molecules isworthwhile for multiple reasons:

1. Role in biological processes. Since experimentshave provided strong evidence that Posner moleculesare stable in simulated body fluids, it is likely thatthey are present in vivo, particularly in the extracellu-lar fluid where the free calcium concentration is appre-ciable. These extracelluar Posner molecules could act asa reservoir for the previously suggested Posner-molecule-mediated mechanism for bone growth.3,4 The moleculesmay also be present within cells. Though cytoplasm isunlikely to contain Posner molecules due to its low cal-cium concentration, the mitochondiral matrix is knownto contain stable calcium-phosphate complexes with acalcium:phosphate ratio suggestively close to that of Pos-ner molecules (3:2).9 Posner molecules may form a dy-namic aggregate within the mitochondiral matrix, withformation, aggregation, and dis-aggregation being finelymodulated by the changing pH.10

The Posner molecule is also central to the proposed“quantum brain” concept set forth in Ref. 11. In this pro-posal, clouds of entangled Posner molecules in the brainact as “neural qubits”, serving as a platform for quantumcomputation in cognitive processes. Many of the proper-ties we discuss here are key to the ongoing exploration ofthis concept.

2. Long-lived phosphorus nuclear spin states.

arX

iv:1

711.

0589

9v1

[ph

ysic

s.ch

em-p

h] 1

6 N

ov 2

017

2

FIG. 1. Posner molecule, Ca9(PO4)6, with S6 symmetry. Thesymmetry axis is coming out of the page, tilted at a slightangle to show underlying calcium ions. The symmetry axismay be thought of as the (111) axis of a cube. A phosphateion tetrahedrally coordinated by oxygens is at each face ofthe cube, a calcium ion is at each vertex, and a final calciumion is in the center. The cube is flattened by about 3% alongthe symmetry axis. Calcium ions are shown in blue, andphosphate ions are represented as purple tetrahedra with redoxygen ions at the vertices.

Molecules or ions with isolated nuclear spins that ex-hibit macroscopic coherence times are of great physicalinterest, both theoretically and practically. Nuclear spinsin liquid-state nuclear magnetic resonance (NMR) havesome of the longest known coherence times of any sys-tem in physics, especially nuclei with spin 1/2 that donot couple to electric fields and interact with the envi-ronment only through magnetic fields (e.g., dipole fieldsfrom other nuclear spins) and intra-molecular electron-mediated exchange interactions.12 Motional narrowingdue to the rapid rotation of the molecules in solution av-erages out dipole-dipole magnetic field interactions, giv-ing spin-1/2 nuclei long coherence times.13

Since the most abundant isotopes of both calcium andoxygen have zero nuclear spin, we predict that the six31P nuclear spins with S=1/2 in a Posner molecule areespecially long lived, making them an intriguing platformfor liquid-state NMR quantum computation.14 If the nu-clear spins can be hyper-polarized, the spins could havemedical imaging applications, since the 31P NMR sig-nal is quite robust. The Posner molecule’s nuclear spinsare also central to the quantum brain concept,11 in whichlong coherence times are key to their role as neural qubits.

3. Doping with other cations. We shall present evi-dence that replacing the calcium ion at the center of aPosner molecule with either another divalent cation or apair of monovalent cations can further stabilize the Pos-ner molecule. This provides a possible mechanism for theknown impact of elements such as lithium, magnesium,

and iron on bone health.15–17 Impurities are also relevantto the quantum brain hypothesis,11 which suggests thatthe cognitive effects of lithium (and the isotope depen-dence of these effects) may arise from the interaction ofthe lithium nuclear spin with the neural qubits providedby the 31P nuclear spins.

4. Aggregation of Posner molecules. The presence oftwo Posner clusters within the unit cell of hydroxyap-atite suggests the possible importance of aggregation ofPosner molecules in bone growth. Our quantum chem-istry calculations provide evidence that a pair of Posnermolecules can chemically bind together (in vacuum) witha substantial binding energy of roughly 5 eV. This pair-wise bonding is the first step towards larger-scale aggre-gation, which may lead to the formation of amorphouscalcium phosphate and eventually hydroxyapatite.3–6

5. Quantum dynamics of the six phosphorus nu-clear spins in a Posner molecule. Provided the Posnermolecules do have an S6 symmetry, the 26 = 64 nuclearspin eigenstates in each molecule can be chosen also aseigenstates under a 3-fold rotation about the symmetryaxis, with eigenvalues of the form ei2πτ/3 with the “pseu-dospin” τ taking one of 3 allowed values, τ = 0,±1. Un-derstanding the dynamics of these spins is important forany application of Posner molecules to medical imagingor quantum computation as discussed above. The spindynamics are also key to the quantum brain concept.11

Our paper is organized as follows. After a brief descrip-tion of the computational methods (Sec. II), we treatthe structural properties of the Posner molecule: sym-metry (Sec. III A), vibrational spectra (Sec. III B), im-purity incorporation (Sec. III C), and pairwise bonding(Sec. III D), discussing the results and implications ofeach in turn. We then move on to the spin properties,using first-principles methods to calculate the indirectspin-spin coupling between 31P nuclear spins in a Pos-ner molecule. These values become input to an effectiveHamiltonian for the spins. We study this effective model,paying special attention to the implications for coherenceof encoded quantum information (Sec. III E).

II. COMPUTATIONAL METHODS

Our first-principles calculations are based on densityfunctional theory (DFT).

Molecular symmetry, impurity incorporation, andpairwise bonding were studied with the Projector-Augmented Wave (PAW) method18 as implemented inthe Vienna Ab initio Simulation Package (VASP).19 Thehybrid exchange-correlation functional of Heyd, Scuseria,and Ernzerhof20 was employed with the standard 25%mixing and screening parameter ω = 0.20 A−1 (a combi-nation of parameters commonly referred to as HSE06).21

These calculations used a plane-wave energy cutoff of 400eV. Single Posner molecule calculations were performedin a vacuum supercell 16 A on a side. Calculations explor-ing the chemical bonding between two Posner molecules

3

used a 16 A×16 A×32 A cell.The vibrational spectrum of Ca9(PO4)6 was calcu-

lated via density functional perturbation theory (DFPT)with the Quantum Espresso package22 using ultra-soft pseudopotentials23 and the Perdew-Burke-Ernzerhofexchange-correlation functional.24 The acoustic sum rulewas applied for the diagonalization of the dynamical ma-trix in order to account for the molecule’s translationaland rotational degrees of freedom.

Calculations of the pairwise nuclear spin-spin interac-tions between 31P nuclear spins inside a Posner moleculewere performed in Dalton201525,26 with the B3LYP hy-brid functional27 using the method of Refs. 28 and 29.The electronic states around oxygen and calcium wereexpanded in the popular 6-31G** basis,30,31 while thosearound phosphorus were expanded in the pcJ-4 basis,which is optimized for the calculation of indirect spin-spin couplings.32

Molecular visualizations were produced usingVESTA.33

III. RESULTS AND DISCUSSION

A. Symmetry

Previous computational work7,8 used DFT to study thepossible symmetries of a Posner molecule. The authorsfound various candidates for the lowest-energy molecularstructure, all within the numerical accuracy of the tech-nique. They concluded that the S6 structure, the can-didate ground state with the highest symmetry, is theprototypical Posner molecule.

Our calculations agree with these results. Severalstructures are illustrated in Fig. 2. The lowest-energystructures have C1 (no) symmetry but are only slightlydistorted from S6 symmetry. The S6 structure is higherin energy by only 0.06 eV, or less than 2 meV per atom,while the Th structure is a full 1.98 eV higher in energythan the S6. This is consistent with earlier work, whichidentified S6 as the prototypical structure.8 To furthertest this identification, five C1 structures were generatedby random 0.1 A perturbations of an S6 structure. Afterrelaxation, the atomic positions of these structures arefound to deviate from S6 symmetry by at most 0.008 A,while the average positions across all five structures devi-ate from S6 symmetry by at most 0.002 A. This demon-strates that the S6 structure is indeed correct on average,and we expect thermal fluctuations will wash out anysymmetry breaking. We therefore assume the S6 sym-metry for the Posner molecule in the remainder of thiswork.

B. Vibrational Spectra

Experiments find evidence for calcium phosphate clus-ters roughly 1 nm in size in simulated body fluid.6,34

Th

S6 C1

FIG. 2. Different symmetries of a Posner molecule,Ca9(PO4)6. All structures were relaxed in vacuum. The Th

structure is unstable to distortion (compression along a diag-onal of the cube and rotation of the phosphate tetrahedra),lowering the symmetry to S6 about the compressed diagonal.The reduction of symmetry from Th to S6 lowers the energysubstantially, by 1.98 eV. While the S6 structure is locallystable, a C1 structure corresponding to an off-centering of thecalcium atoms lowers the energy by only 0.06 eV. As discussedin the text, we identify the S6 structure as the prototypicalPosner molecule.

While it is suspected that these clusters are Pos-ner molecules,3 this has not been shown conclusively.Spectroscopic probes such as infrared absorption spec-troscopy, Raman scattering, or wide-angle X-ray scatter-ing could identify the molecules definitively. IR spec-troscopy is based on the absorption of incoming photonsresonant with a dipole-inducing vibrational mode of amolecule.35 A calculation of vibrational modes and theirassociated dipole moments indicates the wavelengths oflight that could be absorbed by the molecule. The IRactivity of the associated vibrational mode, which is pro-portional to the square of the induced dipole moment,corresponds to the strength of the peak in an absorp-tion experiment. Practically speaking, IR spectroscopyin solution is challenging, since the H2O peak at 1575cm−1 with an IR activity of 1.659 (D/A)2/amu 36 tendsto wash out the IR absorption signal of any species inaqueous solution. Nevertheless, spectroscopic methodsremain one of the best ways to conclusively identify Pos-ner molecules, and calculations of vibrational spectra arean essential step in this process.

In addition to their spectroscopic relevance, the vibra-tional modes of the molecule also couple to the interac-tions between the six phosphorus nuclear spins, whichwill be discussed in section III E.

The vibrational spectra of both the S6 structure and asymmetry-broken C1 structure are shown in Fig. 3. Note

4

a)

b)

025050075010000

10

20

30

40

ν (cm-1 )

IRActivity

025050075010000

10

20

30

40

ν (cm-1 )

IRActivity

FIG. 3. Vibrational spectrum of a Posner molecule with a)S6 and b) C1 symmetry. The IR Activity [in (D/A)2/amu] ofeach mode is plotted versus frequency.

that the S6 symmetry guarantees that a number of themodes will induce no dipole moment, and thus have zeroIR activity. The corresponding vibrational modes of theC1 structure are shifted slightly and have a small IR ac-tivity, but the overall shape of the spectrum remains thesame. In addition to the plotted modes, we also foundmodes with small imaginary frequencies. The S6 struc-ture has two such modes, and the C1 has one. Theseimaginary frequencies represent soft modes which movebetween different low-energy states. Their existence isunsurprising given the presence of many nearly degener-ate C1 structures similar to the S6 structure.

C. Impurity substitution

During the formation of Posner molecules, ions otherthan Ca2+ and PO3−

4 will typically be present in solu-tion, and thus could be substituted for one of the nativeions. We refer to these ionic substitutions as “impuri-ties”. Here we consider the energetics of substituting thecentral calcium ion with either another divalent cationor with two monovalent cations. In making any com-parisons in energies, it is important to take into account

the hydration energies of both the ions to be substitutedand of the central calcium ion once removed from themolecule. In principle the hydration energy of a Posnermolecule in solution also needs to be taken into account.However, it is reasonable to assume that any changes inthis hydration energy upon impurity substitution of thecentral calcium ion (which is encased inside the molecule)will be small, and hence they are ignored here. The en-thalpy of hydration for a single ion, ∆H0

hyd, is definedas the energy change upon taking an ion from a gaseousstate to a dilute solution in water. It is always negativesince polar water molecules can considerably reduce theirenergy by rearranging around the ionic point charge.37 Itis notoriously difficult to calculate hydration enthalpiesfrom first principles;38,39 in this work we take these valuesfrom experiment.40

We will denote the enthalpy change due to the sub-stitution of the central calcium ion with either a singledivalent cation or of two monovalent cations as ∆HA,where A specifies the cation substituted. ∆HA can beexpressed as

∆HA =(

∆H0f [AnCa8(PO4)6] + ∆H0

hyd[Ca2+])

−(

∆H0f [Ca9(PO4)6] + n∆H0

hyd

[A(3−n)+]

) (1)

where n = 1 corresponds to a single substituted divalentcation, and n = 2 corresponds to a pair of substitutedmonovalent cations. Here ∆H0

f is the enthalpy of forma-tion of the Posner molecule with or without the impurity.

Our results for several selected impurities are presentedin Table I. We find a significant difference in the favorabil-ity of various ionic substitutions. This difference is largeenough to outweigh the hydration enthalpy of cationssuch as Li+ and Fe2+, making them highly favored asimpurities. Indeed, the trend is that ions with a strongertendency to hydrate have an even stronger tendency tosubstitute for Ca2+ in the Posner molecule. Increasedhydration enthalpy is outweighed by increased stabilityon the central site of the Posner molecule.

These results suggest that if significant concentrationsof lithium, iron, or magnesium are present when Posnermolecules are formed, they are likely to incorporate intothe Posner molecule structure. This could have a vari-ety of implications. In the context of calcium phosphatebiomineralization, the presence of impurity ions and thenature of their interactions with Posner molecules willhave important impacts on Posner-molecule-mediatedbone growth. Additionally, spinful nuclei incorporatedas impurities within the Posner molecule will have a sig-nificant effect on the phosphorus spin states.

D. Bonding

Aggregation of Posner molecules has been proposed asan intermediate step in biomineralization of amorphouscalcium phosphate, a precursor to hydroxyapatite (bone

5

A2+ ∆H0hyd[A2+] (eV) ∆HA2+ (eV)

Fe2+ −20.21 −1.26

Mg2+ −19.96 −1.24

Hg2+ −18.96 −0.23

Ca2+ −16.37 0.00

Pb2+ −15.39 0.51

A+ 2 × ∆H0hyd[A+] (eV) ∆HA+ (eV)

Li+ −10.78 −1.48

Na+ −8.42 −0.86

K+ −6.63 3.62

TABLE I. Energy shift upon substitution of the central Ca2+

ion in a Posner molecule with either a divalent cation or withtwo monovalent cations. Experimental values for hydrationenthalpies (∆H0

hyd) are also listed (Ref. 40). In the case ofmonovalent cations, the reported value is twice the hydrationenthalpy of a single ion.

mineral).3–6 We approach this by studying the simplestform of aggregation: pairwise bonding.

We consider bonding of Posner molecules with the S6

structure described in Sec. III A. A variety of bonding ori-entations for a pair of rigid molecules were tried; the mostfavorable is depicted in Fig. 4(a) and is similar to the rela-tive ionic positions in hydroxyapatite. The molecules aremirror images of one another, oriented so that Ca2+ ionsmeet PO3−

4 ions, and ions of like charges remain sep-arated. The distance between centers and the relativeorientation of the molecules were varied to find the min-imum energy configuration. The resulting configuration,depicted in Fig. 4(a), has a binding energy of 1.04 eV (ref-erenced to isolated molecules in vacuum). Starting fromthis configuration, the constituent ions were allowed torelax. This relaxation gains another 3.95 eV, for a totalbonding energy of 4.99 eV. The relaxed bonded configu-ration is shown in Fig. 4(b).

We note that these calculations do not take the pres-ence of solvent into account. We expect that this is a rea-sonable approximation when Posner molecules are closeenough to bond, since there is not enough space for sol-vent molecules to enter between the molecules and screenthe ionic interaction. At larger separation, the solventwill likely reduce the bonding tendency, suppressing thelong-distance tail of the attraction.

We have also explored the motion of two rigidmolecules in a bonded pair with respect to one another.Specifically, we consider “rolling without slipping” ro-tation, i.e., rotation by both molecules simultaneouslyin opposite directions, such that the mirror symmetryof the configuration is maintained. The energy land-scape for this rotation is mapped out by repeating thedistance-optimization procedure for a set of rotation an-gles, finding the optimum distance for each orientation.The saddle-point configuration (shown in Fig. 4(c)) is at arotation angle of φ = 45◦, and the rotation barrier is 0.33

b)

a)

c)

FIG. 4. Bonding of two Posner molecules. (a) Bonding con-figuration for rigid Posner molecules. The symmetry axes arealigned antiparallel perpendicular to the plane of the page;the two molecules are mirror images of one another. Wherethe Posner molecules meet, the Ca2+ ions and PO3−

4 ionsare in different planes perpendicular to the page, keeping aseparation between ions of like charge. This configurationwas found through a manual search of bonding distance andrelative orientation. The bonding energy (referenced to twoisolated Posner molecules) is 1.04 eV. (b) Bonded pair of Pos-ner molecules after relaxation starting from the configurationshown in (a). The bonding energy is 4.99 eV. The individualmolecules are significantly distorted. (c) Saddle-point config-uration in rotation of rigid Posner molecules with respect toone another: φ = 45◦.

eV. We expect the rotation barrier for rigid molecules is areasonable approximation in the early stages of the bond-ing process (before full relaxation).

6

E. Spin Interactions

1. Phosphorus Nuclear Spin Coupling

Coupling between phosphorus nuclear spins arises dueto two factors: magnetic dipole-dipole interaction and“indirect” spin-spin coupling.12 The rotational motion ofthe molecule tends to average out the dipole-dipole inter-action and the anisotropic part of the indirect coupling,so we only consider the isotropic part of the indirect spin-spin coupling. This coupling between nuclei i and j is asum of four terms:

Jij = JDSOij + JPSOij + JSDij + JFCij , (2)

which represent the diamagnetic spin-orbit (DSO), para-magnetic spin-orbit (PSO), spin-dipole (SD), and Fermicontact (FC) terms.12 Details of the calculation of nu-clear spin-spin couplings were given in Sec. II.

Adding together these four contributions leads to aneffective Heisenberg-like Hamiltonian which describes theinteractions between phosphorus nuclear spins:

H0 =∑i,j

Jij σi · σj . (3)

The nuclear spins of phosphorus are arranged in two equi-lateral triangles, one on top of the other, centered on themolecule’s symmetry axis. The S6 symmetry restrictsthe couplings Jij to three values: nearest-neighbor J1,second-nearest-neighbor J2, and third-nearest-neighborJ3, as shown in Fig. 5. We find J1 = 0.178 Hz, J2 = 0.145Hz, and J3 = −0.003 Hz.

The threefold rotational symmetry of the Posnermolecule ensures that the effective Hamiltonian sharesthis same symmetry. Eigenstates of this rotation have aneigenvalue of ei2πτ/3, where τ can take values of 0,±1.We call this quantum number τ the “pseudospin”. Eigen-states may be expressed using the notation

|ψ〉 = |E,S, Sz, τ〉 (4)

where E is the energy (in Hz), S =∑j σj is the total

spin, and Sz is the z component of total spin. A plot ofthe spectrum is shown in Fig. 6.

2. Pseudospin and Rotations

The transformation properties of the nuclear spinstates of a molecule under a symmetry transforma-tion dictate the allowed values of the rotational angu-lar momentum quantum number L. This effect is mostdramatic and well-studied in molecular hydrogen (H2).Parahydrogen, in which the protons form a spin singlet,is restricted to even values of L by the requirement thatthe wavefunction be antisymmetric under exchange of theprotons. Orthohydrogen (with a proton spin triplet) issimilarly restricted to odd values of L. These spin isomers

J2 J2

J1 J1

J3

σ1

σ2

σ3

σ4

σ5

σ6

FIG. 5. Schematic illustration of the configuration of thephosphorus nuclear spins, labeled 1 through 6, with odd spinson the top layer and even on the bottom. Solid lines indicatethe isotropic spin-spin couplings J1, J2, and J3

●●

● ●

■■■■ ■■ ■■ ■

■ ■■ ■■ ■■■■■■■■ ■■ ■■ ■◆ ◆◆ ◆◆ ◆◆ ◆◆ ◆◆◆◆◆◆◆ ◆◆ ◆◆ ◆◆ ◆◆ ◆

▲▲▲▲▲▲▲

-1 0 1

-0.4

-0.2

0.0

0.2

0.4

τ

E(Hz)

Stot=0

Stot=1

Stot=2

Stot=3

FIG. 6. Eigenstates of H, with energy in Hz on the verticalaxis and pseudospin quantum number τ on the horizontalaxis. The shape and color of the points indicates total spinas shown on the plot. Each point with spin S has 2S + 1degeneracy in Sz.

7

have different thermodynamic, scattering, and chemicalproperties.41

Likewise, for a Posner molecule with threefold rota-tional symmetry, the Fermi statistics of the 31P nucleidictate the allowed rotational angular momentum aboutthe symmetry axis. With three-fold rotational symmetry,the full wavefunction must be unchanged by a rotationby 2π/3, since such a rotation is equivalent to an evennumber of fermion swapping operations. The pseudospinτ therefore constrains the angular momentum, L, to sat-isfy: L+ τ = 0 mod 3.

We propose that this restriction may be important inthe case of two Posner molecules (a, b) binding together.Indeed, the recently proposed Quantum Dynamical Se-lection rule,42 when generalized to the binding of twoPosner molecules, predicts that chemical bonding im-plements a projective measurement onto a state withτa + τb = 0 — essentially a result of the requirementthat binding two Posner molecules stops any relative ro-tational motion. If the pair of Posner molecules subse-quently unbinds, this constraint is predicted to be main-tained, leaving the two molecules “pseudospin entan-gled”. Thus, pair binding/unbinding of Posner moleculesmay provide a mechanism to quantum entangle nuclearspin states in multiple Posner molecules, a necessary pre-condition for the quantum brain concept.11

3. Decoherence

A decoherence time for the spins in a Posner moleculeis the NMR spin-lattice relaxation time T1. The primarymechanism for decoherence is entanglement with exter-nal nuclear spins of protons in water molecules externalto the Posner molecule. The T1 due to dipole-dipole in-teractions between an external spin M (e.g. a proton)and the phosphorus spins (indexed by I) is given by themodified Solomon-Bloembergen equation,43

1

T1=∑I

2

15M(M + 1)C2

DD,I

[τc

1 + (ωM − ωI)2τ2c(5)

+3τc

1 + ω2Iτ

2c

+6τc

1 + (ωM + ωI)2τ2c

],

where CDD,M is the dipole-dipole coupling strength (inHz) between spin M and I, τc is the rotational correla-tion time, and ωM,I are the Larmor frequencies of thespins. We take M = 1/2, and the correlation time tobe given by the thermal rotation frequency 1/τc of thePosner molecule. With a moment of inertia 1.22× 10−43

kg m2, 1/τc = 2.6 × 1011 Hz at 300 K. This is firmly inthe regime ωτc � 1. In this limit, Eq. (5) reduces to

1

T1=∑I

C2DD,Iτc . (6)

As an illustration giving a rough estimate of this co-

herence time, we consider a proton (perhaps associatedwith a water molecule in the solvent) as the external spin,located 7 A from the center of the Posner molecule alongthe symmetry axis (3.5 A from the apical Ca2+). Thisscenario gives T1 = 1.8 × 106 s = 21 days. Since dif-ferent pseudospin sectors couple differently to environ-mental degrees of freedom, the decoherence times for thepseudospin quantum number may be even longer. Thelong-lived spin states in the Posner molecule could pro-vide a platform for liquid-state NMR quantum compu-tation, and are also key to the “quantum brain” conceptset forth in Ref. 11.

IV. CONCLUSIONS

We have explored the structure, symmetry, and spec-troscopic fingerprint of the Posner molecule, Ca9(PO4)6.We have shown that Posner molecules are stable in vac-uum, and identified S6 symmetry as the prototypicalsymmetry. The calculated vibrational spectrum of thePosner molecule may serve as a spectroscopic finger-print, assisting with the experimental identification ofthe Posner molecule either in vitro or in vivo. Impuritycations can take the place of a central calcium, with im-plications for both phosphorus spin properties and bonegrowth. We find that pairwise Posner molecule bondingis an important process, suggesting avenues for researchin bone growth. Finally, we have shown that the Posnermolecule is a promising host for nuclear spins maximallyprotected from environmental decoherence, with possibleimplications in liquid-state NMR quantum computationand medical imaging. We have identified the pseudospinquantum number τ which could encode long-lived coher-ent quantum information in the Posner molecule and mayprovide a mechanism for entangling the molecule’s ro-tational degrees of freedom with its nuclear spin. Thismechanism is central to the Posner molecule’s role as aneural qubit in the quantum brain concept.

ACKNOWLEDGMENTS

We thank Daniel Ish, Jim Swift, and Leo Radzihovskyfor fruitful discussions. M.P.A.F. is grateful to theHeising-Simons Foundation for support, to the NationalScience Foundation for support under Grant No. DMR-14-04230, and to the Caltech Institute of Quantum In-formation and Matter, an NSF Physics Frontiers Centerwith support of the Gordon and Betty Moore Founda-tion. Computational resources were provided by the Ex-treme Science and Engineering Discovery Environment(XSEDE), which is supported by National Science Foun-dation grant number ACI-1548562, and by the Centerfor Scientific Computing from the CNSI, MRL: an NSFMRSEC (DMR-1720256) and NSF CNS-0960316.

8

1 A. S. Posner and F. Betts, Acc. Chem. Res 8, 273 (1975).2 K. Onuma, , and A. Ito, Chem. Mater. 10, 3346 (1998).3 X. Yin and M. J. Stott, J. Chem. Phys. 118 (2003).4 L.-W. Du, S. Bian, B.-D. Gou, Y. Jiang, J. Huang, Y.-X.

Gao, Y.-D. Zhao, W. Wen, T.-L. Zhang, and K. Wang,Cryst. Growth Des 13, 3103 (2013).

5 A. Dey, P. H. H. Bomans, F. A. Muller, J. Will, P. M.Frederik, G. de With, and N. A. J. M. Sommerdijk, Nat.Mater. 9, 1010 (2010).

6 L. Wang, S. Li, E. Ruiz-Agudo, C. V. Putnis, and A. Put-nis, Cryst. Eng. Comm. 14, 6252 (2012).

7 G. Treboux, P. Layrolle, N. Kanzaki, K. Onuma, andA. Ito, J. Phys. Chem. A 104, 5111 (2000).

8 N. Kanzaki, G. Treboux, K. Onuma, S. Tsutsumi, andA. Ito, Biomaterials 22, 2921 (2001).

9 D. G. Nicholls and S. Chalmers, J. Bioenerg. Biomembr.36, 277 (2004).

10 Carol Weingarten and Tobias Fromme, private communi-cations (2017).

11 M. P. A. Fisher, Annals of Physics 362, 593 (2015).12 M. Jaszunski, A. Rizzo, and K. Ruud, in Handbook

of Computational Chemistry , edited by J. Leszczynski(Springer Netherlands, 2014) pp. 361–441.

13 J. N. Mundy, Solid State: Nuclear Methods (AcademicPress, 1984) Chap. 6.2.1.

14 L. M. K. Vandersypen, M. Steffen, G. Breyta, C. S. Yan-noni, M. H. Sherwood, and I. L. Chuang, Nature 414, 883(2001).

15 A. Zamani, G. R. Omrani, and M. M. Nasab, Bone 44,331 (2009).

16 L. Toxqui and M. P. Vaquero, Nutrients 7, 2324 (2015).17 S. Castiglioni, A. Cazzaniga, W. Albisetti, and J. A. M.

Maier, Nutrients 5, 3022 (2013).18 P. E. Blochl, Phys. Rev. B 50, 17953 (1994).19 G. Kresse and J. Furthmuller, Phys. Rev. B 54, 11169

(1996).20 J. Heyd, G. E. Scuseria, and M. Ernzerhof, J. Chem. Phys.

118, 8207 (2003).21 J. Paier, M. Marsman, K. Hummer, G. Kresse, I. C. Ger-

ber, and J. G. Angyan, J. Chem. Phys. 125, 249901(2006).

22 P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car,C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni,I. Dabo, A. D. Corso, S. de Gironcoli, S. Fabris, G. Fratesi,R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj,M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri,R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto,C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen,A. Smogunov, P. Umari, and R. M. Wentzcovitch, J. Phys.Condens. Matter 21, 395502 (2009).

23 D. Vanderbilt, Phys. Rev. B 41, 7892 (1990).24 J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev.

Lett. 77, 3865 (1996).25 Dalton, a molecular electronic structure program, Release

Dalton2015.1 (2015), see http://daltonprogram.org.

26 K. Aidas, C. Angeli, K. L. Bak, V. Bakken, R. Bast, L. Bo-man, O. Christiansen, R. Cimiraglia, S. Coriani, P. Dahle,E. K. Dalskov, U. Ekstrom, T. Enevoldsen, J. J. Erik-sen, P. Ettenhuber, B. Fernandez, L. Ferrighi, H. Fliegl,L. Frediani, K. Hald, A. Halkier, C. Hattig, H. Heiberg,T. Helgaker, A. C. Hennum, H. Hettema, E. Hjertenæs,S. Høst, I.-M. Høyvik, M. F. Iozzi, B. Jansık, H. J. Aa.Jensen, D. Jonsson, P. Jørgensen, J. Kauczor, S. Kirpekar,T. Kjærgaard, W. Klopper, S. Knecht, R. Kobayashi,H. Koch, J. Kongsted, A. Krapp, K. Kristensen, A. Lig-abue, O. B. Lutnæs, J. I. Melo, K. V. Mikkelsen, R. H.Myhre, C. Neiss, C. B. Nielsen, P. Norman, J. Olsen,J. M. H. Olsen, A. Osted, M. J. Packer, F. Pawlowski,T. B. Pedersen, P. F. Provasi, S. Reine, Z. Rinkevicius,T. A. Ruden, K. Ruud, V. V. Rybkin, P. Sa lek, C. C. M.Samson, A. S. de Meras, T. Saue, S. P. A. Sauer, B. Schim-melpfennig, K. Sneskov, A. H. Steindal, K. O. Sylvester-Hvid, P. R. Taylor, A. M. Teale, E. I. Tellgren, D. P. Tew,A. J. Thorvaldsen, L. Thøgersen, O. Vahtras, M. A. Wat-son, D. J. D. Wilson, M. Ziolkowski, and H. Agren, WIREsComput. Mol. Sci. 4, 269 (2015).

27 A. D. Becke, J. Chem. Phys. 98 (1993).28 T. Helgaker, M. Watson, and N. C. Handy, J. Chem. Phys.

113, 9402 (2000).29 T. Helgaker, M. Jaszuski, and M. Pecul, Prog. Nucl. Mag.

Res. Sp. 53, 249 (2008).30 P. Hariharan and J. Pople, Theor. chim. acta 28, 213

(1973).31 V. A. Rassolov, J. A. Pople, M. A. Ratner, and T. L.

Windus, J. Chem. Phys. 109 (1998).32 F. Jensen, J. Chem. Theory Comput. 2, 1360 (2006).33 K. Momma and F. Izumi, J. Appl. Crystallogr. 44, 1272

(2011).34 F. Grases, M. Zelenkov, and O. Shnel, Urolithiasis 42, 9

(2014).35 E. B. Wilson, Molecular vibrations; the theory of infrared

and Raman vibrational spectra (McGraw-Hill, 1955).36 D. Porezag and M. R. Pederson, Phys. Rev. B 54, 7830

(1996).37 G. Wulfsberg, Inorganic Chemistry (University Science

Books, 2000).38 M. Soniat, D. M. Rogers, and S. B. Rempe, J. Chem.

Theory Comput. 11, 2958 (2015).39 A. P. Gaiduk and G. Galli, J. Phys. Chem. Lett. 8, 1496

(2017).40 D. W. Smith, J. Chem. Educ. 54, 540 (1977).41 P. Atkins and J. de Paula, Physical Chemistry, 8th Edition

(W.H.Freeman, 2006).42 M. P. A. Fisher and L. Radzihovsky, arXiv:1707.05320

[physics.chem-ph] (2017).43 R. van Eldik, Advances in Inorganic Chemistry, Vol. 57

(Academic Press, 2005).


Recommended