You are on page 1of 5

Toponomic Quantum Computation

C. Chryssomalakos∗ and L. Hanotel†


Instituto de Ciencias Nucleares
Universidad Nacional Autónoma de México
PO Box 70-543, 04510, CDMX, México.

E. Guzmán-González‡
Departamento de Física
Universidad Autónoma Metropolitana-Iztapalapa
PO Box 55-534, 09340, CDMX, México.

E. Serrano-Ensástiga§
Centro de Nanociencias y Nanotecnología, Universidad Nacional Autónoma de México
arXiv:2202.01973v1 [quant-ph] 4 Feb 2022

PO Box 14, 22800, Ensenada, Baja California, México

Holonomic quantum computation makes use of non-abelian geometric phases, associated to the evolution of
a subspace of quantum states, to encode logical gates. We identify a special class of subspaces, for which a
sequence of rotations results in a non-abelian holonomy of a topological nature, so that it is invariant under
any 𝑆𝑂(3)-perturbation. Making use of a Majorana-like stellar representation for subspaces, we give explicit
examples of topological-holonomic (or toponomic) NOT and CNOT gates.

It is well known that cyclic evolution of a quantum state [𝜓] tioned above. To this end, we need to synthesize a number of
in the projective Hilbert space ℙ() ≡ ℙ of a physical sys- results, some of them quite new.
tem gives rise to a geometric phase, which is invariant under The first ingredient we will need is a formulation of the
time-reparametrizatons of the curve 𝐶(𝑡) traced in ℙ [1] (we Wilczek-Zee (WZ) effect, that parallels the treatment of the
denote by [𝜓] points in ℙ, i.e., equivalence classes of normal- abelian case in [8]. In that reference, given a curve [𝜓𝑡 ],
ized states |𝜓⟩ in the Hilbert space  that only differ by an 0 ≤ 𝑡 ≤ 𝑇 , traced by a state in ℙ, one chooses arbitrarily but
overall phase). The non-abelian generalization of this effect, smoothly a phase for each 𝑡, obtaining a lift |𝜓𝑡 ⟩ in . In terms
due to Wilczek and Zee [2], involves the cyclic evolution of of the latter, the geometric phase assigned to [𝜓𝑡 ] is given by
a 𝑘-dimensional degenerate subspace Π of the Hilbert space
, in which case the phase factor mentioned above gets re- 𝜑geo = 𝜑tot − 𝜑dyn
placed by a unitary 𝑘 × 𝑘 matrix 𝑈 which acts in Π — as in the 𝑇
abelian case, 𝑈 is reparametrization invariant. This latter fact = arg⟨𝜓(0)|𝜓(𝑇 )⟩ + 𝑖 𝑑𝑡 ⟨𝜓(𝑡)|𝜕𝑡 |𝜓(𝑡)⟩ , (2)
∫0
has prompted the suggestion of using such holonomies [3] in
the implementation of logical gates in quantum computing, as where the two terms on the right hand side are known as the
it provides a certain level of immunity from noise effects [4, 5]. total and dynamical phase, respectively — it is easily seen that
Specifying the above to the case of a spin-𝑠 system, so that 𝜑geo does not depend on the particular lift |𝜓𝑡 ⟩ chosen, so it is
 = ℂ𝑁 , 𝑁 = 2𝑠 + 1, the simplest realization of a cyclic a property of the curve [𝜓𝑡 ] in ℙ.
evolution of a state in ℙ is through a sequence of rotations 𝑅𝑡 The above physical quantum states [𝜓] correspond to rays
in 𝑆𝑂(3) (with 𝑅0 being the identity operation), in , and the set of such rays is the projective space ℙ. Sim-
ilarly, 𝑘-dimensional subspaces in , of the type that ap-
[𝜓𝑡 ] = [𝐷(𝑠) (𝑅̃ 𝑡 )|𝜓0 ⟩] , (1)
pear in the WZ effect, form the Grassmannian Gr(𝑘, 𝑁) (note
as in the familiar example of a precessing spin, where 𝐷(𝑠) (𝑅̃ 𝑡 ) that Gr(1, 𝑁) = ℙ) and the cyclic evolution of such a sub-
is the spin-𝑠 irreducible representation of the lift 𝑅̃ 𝑡 of 𝑅𝑡 in space corresponds to a closed curve Π(𝑡), in Gr(𝑘, 𝑁), with
𝑆𝑈 (2) — said lift is unique, if one specifies that 𝑅̃ 0 = 𝐼 0 ≤ 𝑡 ≤ 𝑇 and Π(0) = Π(𝑇 ). Given an arbitrary or-
(rather than −𝐼). For a general state [𝜓0 ], cyclicity implies that thonormal basis {|𝜙𝑖 (𝑡)⟩}, 𝑖 = 1, … , 𝑘 in Π(𝑡), with |𝜙𝑖 (0)⟩
𝑅𝑡 must start and end at the identity of 𝑆𝑂(3), 𝑅0 = 𝑅𝑇 = 𝐼, not necessarily equal to |𝜙𝑖 (𝑇 )⟩ (since a given plane may be
but when [𝜓0 ] has a nontrivial discrete rotational symmetry spanned by many different bases), define the matrix 𝑄 with
group Γ = {𝑅(0) ≡ 𝐼, 𝑅(1) , … , 𝑅(𝑝) }, 𝑅𝑡 might also end at any entries 𝑄𝑖𝑗 = ⟨𝜙𝑖 (0)|𝜙𝑗 (𝑇 )⟩, and the Wilczek-Zee connection
of the symmetry rotations 𝑅(𝑖) ∈ Γ. In this latter case, it has (𝑡), with entries 𝑖𝑗 (𝑡) = ⟨𝜙𝑖 (𝑡)|𝜙̇ 𝑗 (𝑡)⟩, the dot denoting time
been shown in [6], that when |𝜓0 ⟩ is anticoherent [7], meaning derivative. The polar part of 𝑄 is given by 𝑃 = 𝑊 𝑉 † , where
its spin expectation value vanishes, the geometric phase ac- 𝑄 = 𝑊 𝐷𝑉 † is the singular value decomposition of 𝑄, with
quired is of a topological character, and is thus invariant under 𝑊 , 𝑉 unitary, and 𝐷 diagonal. Then to the curve Π(𝑡) we
arbitrary (not necessarily small) smooth perturbations. The assign, following [9] (see also [10]), the unitary holonomy
principal aim of the present work is a generalization of these
considerations to the case of the non-abelian holonomies men- 𝑈geo = 𝑃 𝐹 −1 , (3)
2

where Γ, however, two rotations 𝑅 and 𝑅′ that differ by a symmetry,


( 𝑇 ) 𝑅′ = 𝑅𝑅(𝑖) with 𝑅(𝑖) ∈ Γ, give identical results when applied
𝐹 = Pexp d𝑡(𝑡) (4) to Π (via their lift to 𝑆𝑈 (2)), and Π is then identified with
∫0 the quotient space 𝑆𝑂(3)∕Γ.
The above two ingredients may be brought together in the
is the path-ordered exponential of the WZ connection — this
following
is clearly a nonabelian version of the exponentiation of (2). It
can be shown that 𝑈geo transforms covariantly under changes Theorem 1. Let Π be a 1-anticoherent spin-𝑠 𝑘-plane,
in the choice of the basis {|𝜙𝑖 ⟩}, and that it reproduces the with non-trivial rotation symmetry group Γ = {𝑅0 =
standard WZ holonomy when applied to closed curves — it is 𝐼, 𝑅1 , … , 𝑅𝑝 } and denote by Π(𝑡) = 𝑅(𝑡)(Π) the curve in
worth remarking though that (3) may also be applied to open Gr(𝑘, 𝑁) obtained by a sequence of rotations 𝑅(𝑡) ∈ 𝑆𝑂(3) of
curves. Π, that starts, at 𝑡 = 0, at the identity, and ends, at 𝑡 = 𝑇 , on
The second ingredient we will need is a generalization of some symmetry rotation 𝑅𝑚 ∈ Γ, so that Π(𝑇 ) = Π(0). The
the concept of anticoherent spin states. Consider a hermitean WZ holonomy associated, via (3), to the above closed curve
matrix 𝐴, acting on , and the action of an element 𝑔 of 𝑆𝑈 (2) only depends on its homotopy class in Π , and is therefore in-
on it via conjugation, variant under continuous deformations that fix its endpoints.
𝐴 ↦ 𝐴′ = 𝐷(𝑠) (𝑔)𝐴𝐷(𝑠) (𝑔)−1 . (5) Proof. Let {|𝜙𝑖 ⟩}, 𝑖 = 1, … , 𝑘, be a basis in Π = Π(0),
then Π(𝑡) is generated, by definition, by the kets {|𝜙𝑖 (𝑡)⟩ =
A natural question to ask is whether this transformation mixes 𝐷(𝑠) (𝑅(𝑡))|𝜙𝑖 ⟩}, where, for 𝑅(𝑡) = 𝑅𝐦(𝑡) ,
all hermitean 𝑁 × 𝑁 matrices, or whether there are subspaces
of hermitean matrices that only transform among themselves. (𝑠)
𝐷(𝑠) (𝑅(𝑡)) = 𝑒−𝑖𝐦(𝑡)⋅𝐒 , (8)
The answer is that there is a 1-dimensional subspace (multiples
of the identity matrix) that is invariant under this action, then a the direction of 𝐦(𝑡) defining the rotation axis, while its mod-
(𝑠) (𝑠) (𝑠)
3-dimensional subspace, spanned by matrices 𝑇1,1 , 𝑇1,0 , 𝑇1,−1 , ulus, in the interval [0, 𝜋], the corresponding rotation an-
that only transform among themselves, then a 5-dimensional gle. Then we get for the time derivative |𝜙̇ 𝑗 (𝑡)⟩ = −𝑖𝐦′ (𝑡) ⋅
subspace, etc., up to a (4𝑠 + 1)-dimensional invariant subspace 𝐒(𝑠) |𝜙𝑗 (𝑡)⟩, where 𝐦′ (𝑡) is given by a complicated expression,
(𝑠)
— the matrices 𝑇𝓁𝑚 , 𝓁 = 0, 1, … , 2𝑠, −𝓁 ≤ 𝑚 ≤ 𝓁, are called the details of which are immaterial to the present argument,
(𝑠)
spin-𝑠 polarization tensors [11]. In particular, the triplet 𝑇1𝑚 , since we may already conclude that the WZ connection van-
𝑚 = ±1, 0, are, up to normalization, the three 𝔰𝔲(2) gener- ishes for all times,
ators, 𝑆± , 𝑆𝑧 , in the spin-𝑠 representation. We define now a
𝑡-anticoherent spin-𝑠 𝑘-plane Π ∈ Gr(𝑘, 𝑁) by the require- 𝑖𝑗 (𝑡) = −𝑖⟨𝜙𝑖 (𝑡)|𝐦′ (𝑡) ⋅ 𝐒(𝑠) |𝜙𝑗 (𝑡)⟩ = 0 , (9)
ment that any basis {|𝜓𝑖 ⟩}, 𝑖 = 1, … , 𝑘 in Π satisfies
by virtue of the 1-anticoherence assumption about Π, and the
(𝑠) rotational invariance of the concept. Then the second factor
⟨𝜓𝑖 |𝑇𝓁𝑚 |𝜓𝑗 ⟩ = 0, (6)
in the r.h.s. of (3) is equal to the unit matrix, and 𝑈g = 𝑄𝑃 ,
for 1 ≤ 𝓁 ≤ 𝑡, −𝓁 ≤ 𝑚 ≤ 𝓁, 𝑖, 𝑗 = 1, … , 𝑘. Note that, just like which, evidently, is invariant under continuous deformations
for the 𝑘 = 1 definition, the above notion of 𝑡-anticoherence of the curve Π(𝑡) that fix its endpoints.
is rotationally invariant. We will actually only consider here
1-anticoherent planes, for which The above theorem forms the basis for the applications to
quantum computing we discuss later on. Before getting to
⟨𝜓𝑖 |𝐒(𝑠) |𝜓𝑗 ⟩ = 0 , (7) that part though, there is still one final ingredient missing that
needs to be incorporated in our discussion, and which is cap-
where 𝐒(𝑠) = (𝑆𝑥(𝑠) , 𝑆𝑦(𝑠) , 𝑆𝑧(𝑠) ) is the vector of the 𝔰𝔲(2) gen- tured in the following two related questions:
erators, in the spin-𝑠 representation — this particular case can
1. How do we identify closed curves in the Grassmannian?
be related to the concept of anticoherent subspaces in [12].
Note that 𝑆𝑈 (2) acts on a 𝑘-plane in  by transforming the 2. Given a spin-𝑠, 𝑘-plane in , generated by the kets
elements of any basis that generates it, i.e., if Π ∈ Gr(𝑘, 𝑁) is {|𝜙𝑖 ⟩}, 𝑖 = 1, … , 𝑘, how do we identify its possible
generated by {|𝜙𝑖 ⟩}, 𝑖 = 1, … , 𝑘, then the transformed plane rotational symmetries?
Π𝑔 , 𝑔 ∈ 𝑆𝑈 (2), is generated by {𝐷(𝑠) (𝑔)|𝜙𝑖 ⟩}, 𝑖 = 1, … , 𝑘.
Of crucial importance to our subsequent discussion is the orbit Both questions are nontrivial, because we have been specify-
Π of Π under this action, i.e., the set of all 𝑘-planes that can ing a plane giving a basis in it, and there are many bases gener-
be obtained by transforming Π. For a generic plane Π, with ating the same plane. Thus, for a curve Π(𝑡) in the Grassman-
𝑘 ≥ 2, i.e., one that has no rotational symmetries, the orbit Π nian, with the plane Π(𝑡) being generated by the kets {|𝜙𝑖 (𝑡)⟩},
is essentially a copy of 𝑆𝑂(3) since distinct 𝑆𝑈 (2) elements 𝑖 = 1, … , 𝑘, one may have |𝜙𝑖 (𝑇 )⟩ ≠ |𝜙𝑖 (0)⟩, and, yet,
give rise to distinct planes, except when they differ only by a Π(𝑇 ) = Π(0). Similarly, the basis kets rotated by a partic-
sign. In the presence of a discrete rotational symmetry group ular rotation 𝑅𝑚 might not coincide with the original ones,
3

𝐷(𝑠) (𝑅𝑚 )|𝜙𝑖 ⟩ ≠ |𝜙𝑖 ⟩, but the plane they generate after the ro-
tation might be identical with the original one, 𝑅𝑚 (Π) = Π.
Clearly, to be able to describe efficiently loops in Gr(𝑘, 𝑁),
we need an intrinsic characterization of a 𝑘-plane, that brings
to the forefront its transformation properties under rotations.
For the case 𝑘 = 1, i.e., when dealing with physical states in
ℙ, there is such a characterization, due to Majorana [13] (see
also [14, 15]). It is well known that spin-1/2 states can be rep-
resented as points on the Bloch sphere, essentially via their
spin expectation value,
[𝜓] ↦ 𝑛̂ = 2⟨𝜓|𝐒|𝜓⟩ . (10)
It is also a fact that the quantum states of a system of 2𝑠 qubits,
that are symmetric under permutations of the qubits, are in 1-
to-1 correspondence with spin-𝑠√states, e.g., the 2-qubit sym-
metric state (| + −⟩ + | − +⟩)∕ 2 corresponds to the spin-1 FIG. 1. Majorana constellation of the spin-2 state |𝜒⟩ of Eq. (12),
state |𝑠 = 1, 𝑚 = 0⟩ — a similar correspondence holds for any forming a regular tetrahedron with one vertex on the 𝑥-axis.
spin-𝑠 state |𝜓⟩,
|𝜓⟩ ↦ |𝜓1 ⟩ ∨ |𝜓2 ⟩ ∨ … ∨ |𝜓2𝑠 ⟩ , (11) in [16], albeit in a more complicated form. A general spin-𝑠
where the |𝜓𝑖 ⟩ are qubit states and the ∨-product denotes a 𝑘-plane can be uniquely represented by a family of constel-
symmetrized tensor product, e.g., |𝜓1 ⟩ ∨ |𝜓2 ⟩ ≡ |𝜓1 ⟩ ⊗ lations, of various spins, the values of which depend on both
√ 𝑠 and 𝑘, each of which is weighted by a complex number. If
|𝜓2 ⟩ + |𝜓2 ⟩ ⊗ |𝜓1 ⟩)∕ 2. Of course, a 1-to-1 correspon-
one arbitrarily orders the constellations, one may treat these
dence can always be established, in an infinity of ways, be-
complex weights as a pseudo-spinor, and assign to it a “spec-
tween vector spaces of equal dimension — the special prop-
tator” constellation, so that the above 𝑘-plane representation
erty of the particular correspondence mentioned above is that
is purely visual. There is a way of doing this such that un-
it commutes with the action of rotations, namely, given a spin-
der rotations, the Majorana constellations rotate rigidly, while
𝑠 state |𝜓⟩ = (𝜓1 , … , 𝜓𝑁 ), one may rotate it by 𝑅 ∈ 𝑆𝑂(3) by
the spectator constellation remains invariant. Thus, any rota-
left-multiplying it with 𝐷(𝑠) (𝑅), or, one may represent |𝜓⟩ as
tional symmetries of the 𝑘-plane are conveniently encoded in
in (11), transform each of the qubits with the spin-1/2 matrix
its multiconstellation.
𝐷(1∕2) (𝑅), and map the result back into a spin-𝑠 state with the
As a concrete example of the above general procedure, con-
inverse mapping — the results of the two sequences of opera-
sider the spin-2 2-plane ΠNOT ∈ Gr(2, 5), generated by
tions are the same. Since each state in the r.h.s. of (11) can be
represented by a point on the unit sphere, |𝜓⟩ itself can be rep- 1 √
resented by an unordered (because of the symmetrization) set |𝜓1 ⟩ = √ (1, 0, 0, 2, 0) (13)
of 2𝑠 points (stars) on the sphere, its Majorana constellation. 3
1 √
When |𝜓⟩ is transformed by 𝐷(𝑠) (𝑅) in , its constellation |𝜓2 ⟩ = √ (0, − 2, 0, 0, 1) . (14)
rigidly rotates by 𝑅 in physical space. To appreciate the enor- 3
mous advantage furnished by this representation of rays in ,
consider the spin-2 state The Majorana constellations of these states are both regular
(√ √ √ √ √ √ √ ) tetrahedra, antipodal to each other, with one vertex on the 𝑧-
3−2 6 3+ 6 1 3− 6 4+ 2 axis (see image on the left in Fig. 2). Following the procedure
|𝜒⟩ = , , √ , , √ , detailed in [16], we find that the vector space inside which
12 6 2 2 6 4 6
(12) the spin-2 2-planes are naturally embedded (this is the Plücker
and ponder whether it has any rotational symmetries, i.e., embedding, see, e.g., p. 43 of [17], or p. 110 of [18]) is 10
whether there exists any rotation 𝑅 such that 𝐷(2) (𝑅)|𝜒⟩ = dimensional, and splits into spin-3 and spin-1 subspaces. For
𝑒𝑖𝛼 |𝜒⟩, 𝛼 ∈ ℝ, so that [𝜒] ∈ ℙ is invariant under 𝑅 — ad- ΠNOT itself we find the components
( ))
mittedly not an easy question. But then, taking a look at the 1 ( √ √ √ )(
corresponding Majorana constellation, shown in Fig. 1, one ΠNOT = − 2, 0, 0, 5, 0, 0, 2 0, 0, 0 , (15)
3
realizes (e.g., by verifying that all stars are equidistant) that it
forms a regular tetrahedron, and one immediately concludes where we have grouped together the components of each spin
that [𝜒] is invariant, e.g., under a rotation about the 𝑥-axis by multiplet. Since the spin-1 component vanishes in this case,
2𝜋∕3. the multiconstellation of ΠNOT only has a spin-3 constella-
Could it be that a similar stellar representation exists for tion, shown on the right in Fig. 2 — this is a regular octahe-
spin-𝑠 𝑘-planes? Indeed it does, as it has been recently shown dron, which fully represents the rotational symmetries of the
4

FIG. 2. Left: Majorana constellations of the states |𝜓1 ⟩ (red), |𝜓2 ⟩ FIG. 3. Majorana constellations of the states |𝜓𝑎 ⟩ (red), |𝜓𝑏 ⟩ (blue),
(blue) — see Eqs. (13), (14). Right: Majorana-like spin-3 constella- which form an alternative basis of ΠNOT .
tion of the plane ΠNOT , generated by |𝜓1 ⟩, |𝜓2 ⟩ — it forms a regular
octahedron. The spin-1 component of ΠNOT is zero, hence, its spec-
tator constellation consists of a single star at the north pole (yellow).

2-plane. Accordingly, the pseudo-spinor of relative weights is


(1, 0), and the spectator constellation, interpreting the pseudo-
spinor as a spin-1/2 state, consists of a single star at the north
pole.
It is easily checked that ΠNOT is 1-anticoherent (as defined
in (6)), so our theorem 1 may be used. We give an example of
a cyclic evolution, involving a symmetry rotation, and produc- FIG. 4. Principal constellation of the spin-5 4-plane in (16)-(19). The
north and south pole, shown in blue (darker gray in print), are quadru-
ing (as WZ holonomy) a toponomic logical NOT gate. Con-
ply occupied.
sider, to that effect, a sequence of rotations 𝑅(𝑡) = 𝑅𝜋𝑡𝐲̂ ap-
plied to ΠNOT , that starts, at 𝑡 = 0, at the identity, and ends, at
𝑡 = 1, on the symmetry rotation 𝑅(1) = 𝑅(0,𝜋,0) , i.e., around generated by the states
the 𝑦-axis by 𝜋. The WZ holonomy for this curve is computed,
from (3), to be 𝜎𝑥 , i.e., the logical gate NOT, and this result 1 √
|𝜓1 ⟩ = (|5, 5⟩ + 2𝑖|5, 0⟩ + |5, −5⟩) , (16)
is invariant under continuous, however large, perturbations of 2
𝑅(𝑡) — accordingly, this realization of the NOT gate is totally 1 √
|𝜓2 ⟩ = (|5, 5⟩ − 2𝑖|5, 0⟩ + |5, −5⟩) , (17)
immune to noise. 2
1 √ √
It is worth emphasizing that we were able to identify the |𝜓3 ⟩ = √ ( 2|5, 3⟩ + 3𝑖|5, −2⟩) , (18)
above symmetry rotation of ΠNOT by inspection of its mul- 5
1 √ √
ticonstellation. It is true that the particular basis |𝜓1,2 ⟩ used |𝜓4 ⟩ = √ ( 3𝑖|5, 2⟩ + 2|5, −3⟩) , (19)
already makes this symmetry obvious, but it should be kept in 5
mind that there is no guarantee that each of the elements of
an arbitrary basis of a plane Π possesses the rotational sym- which can be shown to be 1-anticoherent. An analysis simi-
metries of Π. For example, consider the alternative basis of lar to that presented above allows the identification of its ro-
ΠNOT tational symmetries — space limitations do not allow us to
give the full multiconstellation, we only show the principal
√ (i.e., highest-spin) constellation in Fig. 4. We find that the
𝜋 𝜋 1 1 √ 6 sequence of rotations 𝑅1 (𝑡) = 𝑅𝜋𝑡𝐱̂ ends, at 𝑡 = 1, on the
|𝜓𝑎 ⟩ = cos |𝜓1 ⟩ + sin |𝜓2 ⟩ = √ ( √ , − 3, 0, 1, ) symmetry rotation 𝑅(𝜋,0,0) of ΠCNOT , with the corresponding
3 3 6 2 2
√ holonomy 𝑈1 being given on the left in the equation below,
𝜋 𝜋 1 6 √ 1
|𝜓𝑏 ⟩ = cos |𝜓1 ⟩ − sin |𝜓2 ⟩ = √ ( , 1, 0, 3, − √ ) , ⎛ 1 0 0 0⎞ ⎛ 1 0 0 ⎞ 0
6 6 6 2 2 ⎜ 0 1 0 0⎟ ⎜ 0 1 0 ⎟ 0
𝑈1 = − ⎜
1⎟
, 𝑈2 = ⎜ ⎟
0 0 0 0 0 𝑒𝑖4𝜋∕5 0
⎜ ⎟ ⎜ ⎟
with the corresponding constellations shown in Fig. 3 — nei- ⎝ 0 0 1 0⎠ ⎝ 0 0 0 ⎠
𝑒−𝑖4𝜋∕5
(20)
ther of the basis states has the symmetry 𝑅(0,𝜋,0) of ΠNOT .
i.e., it is a CNOT gate, with an extra overall sign, which is,
As a second example, consider the spin-5 4-plane ΠCNOT , again, totally immune to noise, while 𝑈2 on the right above is
5

the holonomy associated to the sequence of rotations 𝑅2 (𝑡) = [3] B. Simon, “Holonomy, the quantum adiabatic theorem, and
𝑅2𝜋𝑡̂𝐳∕5 , 0 ≤ 𝑡 ≤ 1, with 𝑅2 (1) another symmetry rotation of Berry’s phase,” Phys. Rev. Lett., vol. 51, pp. 2167–2170, 1983.
ΠCNOT . [4] P. Zanardi and M. Rasetti, “Holonomic quantum computation,”
Phys. Lett. A, vol. 264, pp. 94–99, 1999.
Concluding this short letter we point out the necessity to [5] P. Solinas, P. Zanardi, and N. Zanghí, “Robustness of non-
refine our search for anticoherent planes in the spin-𝑠 Hilbert abelian holonomic quantum gates against parametric noise,”
space. The ones presented above, as well as several others that Phys. Rev. A, vol. 70, p. 042316, 2004.
we are aware of, were found with a variety of ad hoc methods [6] P. Aguilar, C. Chryssomalakos, E. Guzmán-González, L. Han-
otel, and E. Serrano-Ensástiga, “When geometric phases turn
that fail to convey a satisfactory geometrical picture of their
topological,” J. Phys. A: Math. Theor., vol. 53, no. 6, p. 065301,
locus inside the corresponding Grassmannian — we are cur- 2020. arXiv:1903.05022.
rently pursuing such an understanding in the hope of realizing [7] J. Zimba, “Anticoherent spin states via the Majorana representa-
additional noise-tolerant logical gates. tion,” Electronic Journal of Theoretical Physics, vol. 3, no. 10,
pp. 143–156, 2006.
[8] N. Mukunda and R. Simon, “Quantum kinematic approach
to the geometric phase: I. General formalism,” Ann. Phys.,
vol. 228, pp. 205–268, 1993.
[9] D. Kult, J. Åberg, and E. Sjöqvist, “Noncyclic geometric
ACKNOWLEDGEMENTS changes of quantum states,” Phys. Rev. A, vol. 74, p. 022106,
2006.
[10] E. Sjöqvist, “Geometric phases in quantum information,” In-
CC, LH, and ESE would like to acknowledge partial finan- tern. J. Quant. Chem., vol. 115, pp. 1311–1326, 2015.
cial support from the DGAPA PAPIIT project IN111920 of [11] D. Varshalovich, A. Moskalev, and V. Khersonskii, Quantum
UNAM. ESE would also like to thank DGAPA UNAM for a Theory of Angular Momentum. World Scientific, 1988.
postdoctoral fellowship. [12] R. Pereira and C. Paul-Paddlock, “Anticoherent subspaces,”
Jour. Math. Phys., vol. 58, p. 062107, 2017.
[13] E. Majorana, “Atomi orientati in campo magnetico variabile,”
Nuovo Cimento, vol. 9, pp. 43–50, 1932.
[14] I. Bengtsson and K. Życzkowski, Geometry of Quantum States

chryss@nucleares.unam.mx (2nd Ed.). Cambridge University Press, 2017.

hanotel@correo.nucleares.unam.mx [15] C. Chryssomalakos, E. Guzmán-González, and E. Serrano-

edgar.guzman@correo.nucleares.unam.mx Ensástiga, “Geometry of spin coherent states,” J. Phys. A: Math.
§
edensastiga@ens.cnyn.unam.mx Theor., vol. 51, no. 16, p. 165202, 2018. arXiv:1710.11326.
[1] M. V. Berry, “Quantal phase factors accompanying adiabatic [16] C. Chryssomalakos, E. Guzmán-González, L. Hanotel, and
changes,” Proceedings of the Royal Society of London A: Mathe- E. Serrano-Ensástiga, “Stellar representation of multipartite an-
matical, Physical and Engineering Sciences, vol. 392, no. 1802, tisymmetric states,” Commun. Math. Phys., vol. 381, pp. 735–
pp. 45–57, 1984. 764, 2021. arXiv:1909.02592.
[2] F. Wilczek and A. Zee, “Appearance of gauge structure in simple [17] I. R. Shafarevich, Basic Algebraic Geometry 1. Springer, 2013.
dynamical systems,” Phys. Rev. Lett., vol. 52, pp. 2111–2114, [18] N. Jacobson, Finite-Dimensional Division Algebras Over
Jun 1984. Fields. Springer, 2010.

You might also like