On the classification of two-qubit group orbits and the use
of coarse-grained ‘shape’ as a superselection property
Thomas Hebdige
1
and David Jennings
1,2,3
1
Controlled Quantum Dynamics Theory Group, Imp erial College London, Prince Consort Road, London SW7 2BW, UK
2
Department of Physics, University of Oxford, Oxford, OX1 3PU, UK
3
School of Physics and Astronomy, University of Leeds, Leeds, LS2 9JT, UK.
January 30, 2019
Recently a complete set of entropic con-
ditions has been derived for the intercon-
version structure of states under quan-
tum operations that respect a specified
symmetry action, however the core struc-
ture of these conditions is still only par-
tially understood. Here we develop a
coarse-grained description with the aim
of shedding light on both the structure
and the complexity of this general prob-
lem. Specifically, we consider the degree
to which one can associate a basic ‘shape’
property to a quantum state or channel
that captures coarse-grained data either
for state interconversion or for the use of
a state within a simulation protocol. We
provide a complete solution for the two-
qubit case under the rotation group, give
analysis for the more general case and dis-
cuss possible extensions of the approach.
1 Introduction
Symmetry principles are ubiquitous in both the
classical and quantum realms. They constrain
dynamics, connect with conservation laws and
simplify computations of physical properties. In
crystallography the crystal structure is described
by a symmetry group and largely determines the
properties of the material [1].
Symmetry considerations have been extremely
successful in the field of quantum information
theory [24]. Moreover a substantial component
of this work has addressed the regime in which
symmetry principles disconnect from conserva-
tion laws [5], and which has found application in
the study of quantum features of thermodynam-
ics [68]. Quantum states that break the symme-
try can also be used to circumvent limitations on
Figure 1: Residual symmetry-types in two-qubit sys-
tems. We associate to each state a ‘shape’ that de-
scribes the residual symmetries of that state. As an ex-
ample, in the above figure we show the symmetry proper-
ties of the set of all mixtures of Bell states under SU(2).
This is determined solely by the triplet component of the
state. The only possibilities are for the state to have a
residual symmetry of (a) a cylinder, (b) a double-sided
rectangle, or (c) a sphere. This ‘shape’ can be used
to provide superselection rules when using the two-qubit
state to simulate quantum operations. Details are given
in section 4.
the precision of measurements imposed by conser-
vation laws, as described by the WAY theorem
[914], while metrology can be viewed as using
broken symmetries to distinguish different group
transformations [15]. Symmetry principles have
also provided new insights into quantum speed
limits and how quickly quantum operations can
be performed [16]. Finally, symmetry groups pro-
vide an invaluable toolkit in the context of quan-
tum computing [1720].
In this work we look at how general quantum
states ρ can be classified by groups that break a
Accepted in Quantum 2019-01-23, click title to verify 1
arXiv:1804.09967v2 [quant-ph] 29 Jan 2019
given symmetry constraint to an equal degree. A
concrete example makes this clearer: both a chair
and an arrow break rotational symmetry, however
an arrow has a residual (axial) symmetry group
to it, and so breaks rotations to a weaker degree
than the chair. For a fixed quantum system of
dimension d, and symmetry action G, we can ask
what residual symmetries does the system admit,
and how do these residual symmetries transform
under symmetric dynamics. We also address the
degree to which this aspect of a quantum system
constitutes a property in the ‘resource-theoretic’
sense.
The structure of the paper is as follows. In
the next section we provide notation and back-
ground motivation. In section 3 we provide the
coarse-graining scheme and some basic relations
involved. This demonstrates the complexity of
the topic, and highlights some subtleties if one
wishes to associate the scheme to a quantum re-
source. In section 4 we provide a complete classi-
fication for two-qubit systems (partly pictured in
Fig. 1 and more fully in Fig. 5), and which illus-
trates how quantum state spaces admit a coarse-
grained partial order.
2 Background and Motivation
In this paper we shall use the following nota-
tion. For any quantum system A we denote
by H
A
its associated Hilbert space, and B(H
A
)
the set of linear operators on H
A
. We assume
that a symmetry action of G is defined on H
A
through the unitary representation g 7→ U (g)
B(H
A
). The group action on a quantum state
ρ B(H
A
) is given by the adjoint action U
g
(ρ) :=
U(g) ρ U
(g).
For any quantum channel E : B(H
A
)
B(H
A
0
), the action of the group on E is U
g
(E) :=
U
g
E U
g
1
(where denotes the concatenation
of channels). This can be interpreted as moving
to a different frame, performing the channel and
returning to the original frame.
If U
g
(E) = E for all g G then E is called
a symmetric operation. This includes certain
preparation procedures, and the resulting states
are known as symmetric states because they are
invariant under the group action: U
g
(ρ) = ρ for
all g G.
2.1 General state transformations under a
symmetry constraint
A primary concern here is the study of general
quantum states or quantum operations under a
symmetry action. In particular we would like
to study the degree to which any given quan-
tum state breaks the symmetry (a ‘resource state’
[21, 22]). One way of comparing states in this
manner is to say that ρ σ if and only if
ρ σ = E(ρ) for some symmetric quantum op-
eration E, which defines a partial order on quan-
tum states [23, 24]. A measure of how much a
state breaks the symmetry is then a function M
that respects the partial order namely, if ρ σ
then M(ρ) M(σ) [21, 22]. A central ques-
tion is whether a complete set of measures ex-
ist that fully capture this ordering of states, and
thus asks when two quantum states can or can-
not be interconverted under the symmetry. Very
recently [7] an answer was given for this general
problem in terms of single-shot entropies of the
form H
min
(R|A) that quantify the amount of in-
formation in a system A about an external quan-
tum reference frame R. Thus it is the case that
ρ σ under the symmetry constraint if and only
if H
min
(R|A) decreases with respect to all refer-
ence frames R.
However this set of measures for the resource
theory is over-complete and highly complex to
use. Applying it to general systems requires so-
phisticated techniques in single-shot information
theory. Here we wish to tackle a more mod-
est goal, and instead of describing the complete
symmetry properties of a quantum state, wish
to coarse-grain states into groups that break the
symmetry in the same way. Notably, this itself
turns out to be a highly non-trivial problem and
so sheds light on the structure of the more general
set of measures given in [7].
2.2 Efficient simulation of quantum channels
One very grounded way of probing the degree
to which a quantum state breaks a symmetry
is to see how useful it is when we wish to per-
form tasks, such as induce a channel on another
system. More generally, if we have some target
quantum operation E on a system, we would like
to know what kind of resource states and interac-
tions are required to realise that operation. The
degree to which a resource σ can do this under
Accepted in Quantum 2019-01-23, click title to verify 2
symmetric dynamics therefore constitutes a mea-
sure of its properties.
In [25], the symmetry properties of quantum
channels were analysed, with a focus on the orbit
of a channel E, defined as
M(E) := {U
g
(E) : g G}. (1)
A similar definition applies to the orbit of states
under the group, so M(ρ) = {U
g
(ρ) : g G}.
Now the type of orbit one obtains provides a
natural context in which to understand both how
the channel breaks a symmetry constraint, and
moreover how one might simulate such a channel
using an auxiliary quantum system B prepared
in a non-symmetric state σ
B
.
It is clear that an auxiliary resource state may
be used with varying efficiency in the simulation
of a channel on a system S, however in [25] the
question of when a state σ
B
can be used to in-
duce a channel E such that its ability to simulate
the very same channel on subsequent systems is
undiminished. This was motivated by the dis-
covery of a “catalytic coherence” protocol [26] in
which a quantum state σ
B
is used as a phase ref-
erence. The state undergoes non-trivial changes
σ
B
σ
(1)
B
σ
(2)
b
· · · σ
(n)
B
yet for all n 1
its abilities as a phase reference never diminish.
In [25] the structure of such arbitrarily repeatable
protocols was spelled out using harmonic anal-
ysis. It was shown that if a channel E on S is
simulated using a quantum state σ
B
under an ar-
bitrarily repeatable protocol then there exists a
POVM measurement {M
k
} on B such that
E(ρ) =
X
k
tr(M
k
σ
k
(ρ) (2)
where {Φ
k
} are completely positive maps on the
primary system S. The POVM {M
k
} on the
auxiliary system depends explicitly on the tar-
get channel E. Crucially however, it can be un-
derstood as resolving the location x of E on its
unitary orbit M(E) to some scale that depends
on the dimension d of S. For example, in the case
of catalytic coherence, the POVM measurement
on B corresponds simply to resolving a point x
on a circle (the orbit of the channel E under phase
rotations) down to an angular scale
2π
d
.
The degree to which x is resolved on M(E)
clearly depends on the resource state σ
B
, and
corresponds to the fact that σ
B
is a quantum
reference frame. This in turn can be viewed as
inducing non-classical geometry on M(E) for
example, for G = SU(2) a channel can have
orbit M(E) that is a 2-sphere. The use of a
bounded reference system in this case means that
only finite resolution of points on M(E) is pos-
sible, however since the 2-sphere carries a phase
space structure the bounded reference case cor-
responds to the so-called “fuzzy sphere” model
from non-commutative geometry [27]. This dif-
fers from catalytic coherence in that we now have
complementarity in resolving the coordinates on
the sphere, and which ultimately arises because
the two orbits have quite different structure.
Thus the general efficient use of a quantum
state σ
B
to simulate a channel E on S can be
analysed in terms of two aspects:
1. (‘Shape’) The type of orbit M(σ) the state
has under the group action compared to
M(E), independent of metrical aspects.
2. (‘Geometry’) The ability of σ to encode clas-
sical coordinate data for x in M(E).
What is a necessary relation between M(E) and
M(σ)? It is clear that we need σ to break the
symmetry to a ‘larger’ degree than E, however in
this paper we would like to make this statement
more precise. Given the complexity of the general
problem, one motivation in this work is to start
with the above separation into ‘shape’ and ‘ge-
ometry’, and develop an organisational setting in
which we provide a coarse-graining over states of
the same ‘shape’, with the remaining task being
to analyse the ‘geometry’ aspect of the state.
2.3 Related topics in quantum state tomogra-
phy and quantum computation
Beyond the abstract problem of simulating an ar-
bitrary quantum channel using a resource state,
there are specific contexts of importance where
such a coarse-grained division naturally arises. In
[28], quantum state tomography was studied in
the context of prior information that restricts the
state to a lower-dimensional submanifold of the
state space. The analysis shows that the topolog-
ical genus of the manifold can be used to bound
the number of measurements needed to discrim-
inate states on the submanifold. The orbit of a
state (or channel) is one such submanifold, and
so the study of what kinds of orbits can exist and
Accepted in Quantum 2019-01-23, click title to verify 3
how they transform among themselves is of po-
tential relevance to quantum tomography under
symmetry constraints.
A wholly practical direction where such analy-
sis might be of relevance is in the computational
power of gate-sets in quantum computation, and
how such gate-sets interact with states that in-
crease the computational power of the gate-set
(such as magic states for the stabilizers [29]). For
example, a recent work provides a classification
of all Clifford gate-sets [17] in a lattice hierar-
chy, and so aspects of the present work might be
applicable in studying how noisy quantum states
increase the computational power of easily realis-
able gate-sets.
2.4 Core questions of the present work
In subsection 2.1 we highlighted that state inter-
conversion ρ σ under a symmetry constraint
is highly non-trivial, and while progress has been
made on this fundamental problem, it is not clear
how complex it is in general. Also, as discussed
in subsection 2.2 recent results on optimal proto-
cols show that the structure of this theory natu-
rally decomposes into a geometric component and
the particular kind of group orbit the quantum
channel has. In light of these considerations one
expects that the group orbit level of description
might provide a form of coarse-grained descrip-
tion that sheds light on the more complex parent
resource theory. By studying this coarse-grained
theory we can shed light on the parent theory, and
moreover extend the notion of a resource theory
from being about pre-orders on quantum states to
considering pre-orders on sub-sets of states. How-
ever, in order for this to work and be meaningful,
it is crucial that the coarse-grained theory be (a)
physically sensible and (b) relate naturally to the
parent theory. Therefore the core questions we
address in this work are the following:
1. Does a coarse-grained resource-theory frame-
work exist?
2. Is this theory consistent with its parent re-
source theory? Is it physically sensible?
3. How complex is the general structure of
symmetry-based resource theories?
In the next section we tackle the first two of
these questions, by showing how a natural par-
tition of the state space arises that is consistent
with the finer-grained resource theory, and de-
scribe its limitations and physical robustness. In
the process, and through exhaustive analysis of
the G = SU(2) case for two qubits, we shed light
on the crucial third question posed.
3 Coarse-graining states and channels
under a Symmetry Action
3.1 Simulating Operations Under Symmetry
Constraints
If one wishes to realise a quantum channel E via
a symmetric interaction with a state σ, the most
natural way to model this is as
E(ρ) = tr
B
V (ρ
A
σ
B
)V
(3)
where V is a covariant unitary, namely
[V, U
A
(g) U
B
(g)] = 0 for all g G. If such
a relation holds, we say that E can be simulated
using the state σ
B
.
A basic classification of states and channels
under a symmetry can be done by studying the
subgroup of residual symmetries for the state or
channel, called the isotropy subgroup or stabilizer
of E and defined as
Iso(E) := {g G : U
g
(E) = E}. (4)
It specifies the residual symmetry that the chan-
nel has under the group action. Symmetric chan-
nels are invariant under all group transforma-
tions, and simply have Iso(E) = G. This notion
also applies to states, where Iso(ρ) := {g G :
U
g
(ρ) = ρ}.
In many cases the quantum operation E has
some residual symmetry, which is characterised
an isotropy subgroup H of G. The possible
isotropy subgroups form an abstract structure
called a lattice [30]. The partial ordering of the
subgroups is defined by subset inclusion, namely
H
1
H
2
if H
1
H
2
for H
1
, H
2
G. In addition,
every pair of elements have a unique supremum
and unique infimum, which define binary opera-
tions called the meet and join [31]. For two sub-
groups H
1
and H
2
, their meet is denoted H
1
H
2
and defined as H
1
H
2
, while their join is denoted
H
1
H
2
, defined as the subgroup generated by
H
1
H
2
[30]. This structure is illustrated by a
Hasse diagram of the subgroups, as shown in Fig.
2.
Accepted in Quantum 2019-01-23, click title to verify 4
Figure 2: Coarse-grained classification of quantum states under a symmetry. The set of all density operators
is partitioned into sets C(H) with definite symmetries (seen on the left). The C(H) correspond to the subgroups
shown in the Hasse diagram in the middle. The Hasse diagram represents the subgroup lattice, with lines indicating
subgroup inclusion, i.e. H
5
H
4
H
3
etc. Note that for finite-dimensional systems many of the C(H) will be
empty. The subsets C(H) can be associated, up to diffeomorphisms, with group orbits (indicated on the right). The
left figure shows the action of group-averaging (over H
2
), going from σ to P
H
2
(σ) as described in section 3.4.
Since quantum operations can be combined in
a number of ways, we present some basic state-
ments that constrain the residual symmetry of
the resultant quantum operation. The proofs for
this section are provided in Appendix A.
Lemma 1. Given any two quantum channels E :
B(H
A
) B(H
A
0
) and F : B(H
B
) B(H
B
0
),
with unitary representations of a group G defined
on all input and output spaces. Then we have the
following:
1. Iso(EF) Iso(E)Iso(F) under the tensor
product group action U
g
(E F) = U
g
(E)
U
g
(F).
2. If B
0
= A then Iso(E F) Iso(E) Iso(F).
3. If A = B and A
0
= B
0
, and p some proba-
bility 0 p 1, then Iso(pE + (1 p)F)
Iso(E) Iso(F).
4. Iso(U
g
E U
g
) = g[Iso(E)]g
1
.
These results also apply for quantum states, once
we view the state as the state preparation map
1 ρ. We can now make precise a coarse-grained
(necessary, but far from sufficient) requirement
that a state σ allow the simulation of a channel
E under symmetric dynamics and show that this
bound is tight in general. The proof is provided
in Appendix A.
Theorem 1. If a system B in a state σ
B
can be
used to simulate a CPTP map E under symmetric
dynamics, then Iso(σ
B
) Iso(E). Moreover, if
E has isotropy group Iso(E) then there exists a
quantum system B and quantum state σ
B
with
Iso(σ
B
) = Iso(E) that allows the simulation of E.
This result gives the basic relation, in the coarse-
grained picture, for when a quantum state σ
B
can
be used to simulate a quantum channel and also
shows that this relation is a tight one. This is
necessary in order for the isotropy subgroup ap-
proach to be consistent with the parent resource
theory, where one is concerned with the minimal
resources needed to realise a particular quantum
channel.
A similar result can be given in the case of ap-
proximate simulation. For any target operation
E we can define the noisy version given by
E
= (1 ) E + D (5)
where D is the complete depolarization map ρ 7→
1
d
for all ρ. From Lemma 1 we see that Iso(E
) =
Iso(E) G = Iso(E). If σ allows the simulation
of E
for some 0 < < 1 then this corresponds
Accepted in Quantum 2019-01-23, click title to verify 5
to an approximate, ‘isotropic’ simulation of the
original map with noise parameter . Theorem 1
extends to this case in an obvious way.
3.2 Is ‘shape’ a resource-theoretic property?
While we normally associate measurable proper-
ties with either projective measurements or more
generally POVMs, there is an alternative way
that is more general again. Specifically, a prop-
erty is associated with a pre-order defined on
the set of all quantum states. This recent ap-
proach is called the resource-theory method, and
has found success in areas such as entanglement,
coherence, thermodynamics, and many other sce-
narios [4, 3234]. Quantum maps that respect the
pre-order are called ‘free operations’ and any real-
valued function on quantum states that respects
the pre-order is a measure of the property.
Given an ability to order the isotropy sub-
groups of quantum states and channels under the
group action, we can ask if a meaningful notion
of ‘shape’ can be defined for quantum systems,
along such a resource-theoretic line. This would
essentially characterise the asymmetry of states
and channels without reference to a measure.
For continuous Lie groups, we find that the or-
bit of a state or channel is always a homogeneous
space [35, 36], specified by both the group G and
the particular isotropy subgroup H of the chan-
nel. More precisely the orbit of the channel is
M(E) and coincides with the quotient G/H up to
diffeomorphisms, which we write M(E)
=
G/H
(and likewise for states).
1
This is what we call
the ‘shape’ of a state or channel.
A simple example can be given for a single
qubit state under an SU(2) symmetry, where the
possible types of group orbit are
M
1
2
[1 + r · σ]
=
(
SU(2)/U(1)
=
S
2
r 6= 0
SU(2)/SU (2)
=
{e} r = 0,
(6)
and these are illustrated in Fig. 3.
The lattice of isotropy subgroups gives us a way
of comparing the ‘shapes’ of group orbits under a
group action G, with M(E) M(F) iff Iso(E)
Iso(F) (up to isomorphism of the isotropy sub-
groups). Likewise for states, M(ρ) M(σ) iff
1
Technically we define the ‘shape’ of the group orbit
with respect to the conjugacy class of the isotropy sub-
group H. This ‘shape’ is also called the Orbit-Type in the
literature. See [37] for further details.
Figure 3: Basic orbits of states. Group orbits for a
single qubit under the action of SU(2) shown in the
Bloch sphere. There are only two possibilities in this
case: a sphere for non-zero Bloch vectors (left), or a
point for the symmetric maximally mixed state (right).
Iso(ρ) Iso(σ) (up to isomorphism). The con-
vention here is to match with resource-theoretic
measures of asymmetry [21, 23], so that one
‘shape’ is ‘bigger’ than another if it has a smaller
isotropy group. For example, a single-qubit state
with non-zero Bloch vector (M(ρ)
=
S
2
) has a
‘bigger’ group orbit than the maximally mixed
state (M(1/2)
=
{e}). This gives valuable infor-
mation regarding the resources required to per-
form a quantum channel in the presence of sym-
metry constraints.
The isotropy subgroup gives us a natural equiv-
alence relation between states that behave in
the same way under the group action. We can
then say that ρ
1
and ρ
2
are related ρ
1
ρ
2
if
we have that Iso(ρ
1
) = Iso(ρ
2
). This equiva-
lence relation partitions the state space into sets
C(H) := {ρ : Iso(ρ) = H}, and collects together
all states that break the symmetry in the same
manner. Taking the union of C(H) corresponding
to isomorphic H partitions the states and chan-
nels according to their ‘shape’.
Viewed as a mapping, C maps from the sub-
group lattice onto a partition of the state space,
and so we can simply allow the partition to in-
herit the lattice structure, and order subsets of
states as C(H
1
) C(H
2
) whenever H
1
H
2
.
Note however that the sets {C(H)} are not
convex in general. For example a qubit system
under G = SU(2), both |0i h0| and |1i h1| have
the same U(1) isotropy subgroup, but
1
2
(|0i h0| +
|1i h1|) =
1
2
1, which is symmetric. However,
C(G) is always a convex set. This follows be-
cause if ρ
1
, ρ
2
C(G) then Lemma 1 implies that
G Iso(p
1
ρ
1
+p
2
ρ
2
) Iso(ρ
1
)Iso(ρ
2
) = G, and
so any mixture has the same isotropy subgroup.
The SU(2) qubit example also highlights that
if C(H) 6= it does not imply that C(H
0
) 6=
Accepted in Quantum 2019-01-23, click title to verify 6
for all H
0
H, since there are subgroups of
SU(2) containing U(1) which do not appear as
the isotropy subgroup of a single qubit state. Not
all possible isotropy subgroups will be seen in a
given system, and for a finite dimensional system
many of the sets C(H) will be empty.
However in order to have a meaning-
ful resource-theory interpretation, the coarse-
grained ordering must be compatible with the set
of free operations namely the symmetric oper-
ations. It is easy to show that this is in fact the
case at the level individual quantum states.
Lemma 2. Under a symmetric operation E,
Iso(E(ρ)) Iso(ρ).
This shows that the coarse-grained ordering of
states is consistent with the set of free (symmet-
ric) operations, as expected. However it does not
imply that a functional mapping is defined on the
sets C(H); it is possible to have ρ
1
and ρ
2
in the
same set C(H), but each get sent to different sets
C(H
0
1
) and C(H
0
2
). This aspect means that inter-
preting the coarse-graining as defining a quantum
resource needs care.
3.3 The speck of dust argument tiny pertur-
bations to symmetric states & channels
Having outlined how the set of quantum states or
the set of quantum channels of a quantum system
is partitioned according to the symmetry action
one might raise the following concern: what hap-
pens if one perturbs a channel via some small
perturbation? How does this affect its isotropy
subgroup?
It is readily seen that one can make an arbitrar-
ily small perturbation to any quantum channel E
so as to break the symmetry in a “maximal” way.
Informally this amounts to the “speck of dust”
argument where, for example, a perfectly rota-
tionally symmetric ball can be made completely
asymmetric by sticking arbitrarily small pieces of
dust onto it (Fig. 4).
In more abstract terms, one can use the princi-
pal orbit-type theorem [37] for the space of quan-
tum channels of a system to see that there is
always a principal isotropy subset that forms a
dense subset in the set of all quantum channels.
Thus “most” quantum channels on a system will
break a symmetry to a maximal degree (the prin-
cipal isotropy).
Figure 4: The speck of dust argument. If we start
with a perfectly rotationally symmetric object (left im-
age) we can always add arbitrarily small perturbations
that completely break the rotational symmetry (right
image). However physically the fact that we have fi-
nite resolution means that quantum states and channels
should always be understood as being defined up to some
–smoothing scale. This does not affect the resource-
theoretic account and therefore the ‘shape’ is a robust
feature in the presence of such tiny symmetry-breaking
perturbations.
This argument of course applies equally well to
the partitioning of the quantum state space into
different subsets with fixed isotropies, however in
both cases this does not pose a problem for our
formalism for the following reasons. Firstly, in
resource theories we are fundamentally interested
in the most efficient use of resources and therefore
our task is to look for those quantum resource
states that break the symmetry not in a maximal
way, but in a minimal way.
Secondly, at the level of channels, the most
interesting and commonly addressed channels in
quantum information science are not of princi-
pal isotropy type namely they have non-trivial
isotropies. For example, consider quantum chan-
nels on a single qubit, and the irreducible group
action of G = SU(2). A typical quantum chan-
nel on the qubit will only have a small resid-
ual isotropy group Z
2
= {1, 1} (the principal
isotropy subgroup of the set of all qubit chan-
nels). However consider the following standard
single qubit channels:
1. The depolarizing channel: E(ρ) = + (1
p)(
1
2
), with 0 p 1.
2. A rotation about the ˆn axis: U
ˆn
(θ) =
exp[iˆn · σ ].
3. A projective measurement along the ˆn axis:
E(ρ) =
P
k
tr(Π
k
ρ
k
, where Π
0
= |ˆni hˆn|
and Π
1
= 1 Π
0
= |−ˆni h−ˆn|.
Accepted in Quantum 2019-01-23, click title to verify 7
4. Partial dephasing along the axis ˆn: E(ρ) =
+(1p)
P
k
tr(Π
k
ρ
k
, where Π
0
= |ˆni hˆn|
and Π
1
= 1 Π
0
= |−ˆni h−ˆn|.
5. A qubit state preparation channel: E(ρ) =
1
2
(1 + p ˆn · σ).
It is readily seen that the isotropy subgroup for
channel (1) is SU(2), while for channels (2–5) it
is a U (1) subgroup of SU(2). The orbits will have
‘shapes’ SU(2)/SU(2)
=
{e} and SU(2)/U (1)
=
S
2
respectively. None of these channels have prin-
cipal isotropy type { , }, however they are
clearly key quantum operations in quantum in-
formation science for which one might wish to
determine the minimal quantum resources neces-
sary to realise them.
Thirdly, the fact that something is “measure
zero” in some space does not imply that it is
physically irrelevant. For example, a 2-d plane
is a measure zero set in three spatial dimensions,
however this does not mean that anyonic physics
or the quantum hall effect is irrelevant. Or of
closer bearing on our present work, we can con-
sider the case of Noether’s theorem: clearly the
set of dynamics with rotational symmetry is again
“measure zero”, however this does not mean that
Noether’s theorem is irrelevant for physics. In
both cases while the restrictions are strictly un-
stable under small perturbations, the important
thing is that the features of interest are opera-
tionally robust under small perturbations.
The speck of dust subtlety makes itself appar-
ent in our setting if one tries to naively exploit
some metric on quantum states to quantify how
far apart the sets {C(H)} are from each other.
It is readily seen that the distance been any two
such sets is in fact zero.
Lemma 3. Let d(·, ·) be any metric on the space
of quantum states. In terms of this metric we
define
d(C(H
1
), C(H
2
)) := inf
σ
1
C(H
1
)
σ
2
C(H
2
)
d(σ
1
, σ
2
). (7)
Then d(C(H
1
), C(H
2
)) = 0 for all H
1
, H
2
G.
This also shows that any set C(H) is arbitrar-
ily close to the set C(G). However, this simply
means that any symmetry properties must always
be considered up to some finite resolution scale
arbitrarily small perturbations can always elimi-
nate residual symmetries, even though this state
is practically indistinguishable from the unper-
turbed one.
Therefore any membership of a quantum state
ρ to a subset should only be considered up to
some smoothing scale based on a distance mea-
sure d (such as from the L
1
norm). For each state
there is an -ball of nearby states, B
(ρ) = {σ :
d(ρ, σ) }. Rather than Iso(ρ), we should con-
sider Iso
(ρ) := max
σ∈B
(ρ)
Iso(σ). Intuitively,
if a state ρ is within a distance of another
state with more residual symmetries, we asso-
ciate those additional residual symmetries with ρ.
The smoothing scale selects the size of symmetry-
breaking perturbations that we wish to consider,
and will depend on the particular physical con-
text.
This reasoning applies equally to quantum
channels. For example, the depolarisation chan-
nel on a qubit is fully invariant under the G =
SU(2) and so no non-trivial resource state is
needed to simulate it. If one perturbs this channel
to E E
by some > 0 perturbation in the dia-
mond norm one can clearly break this isotropy to
the principal isotropy { , }, however it is also
clear (e.g. from the continuity of the Stinespring
dilation [38]) that E
can either be simulated ex-
actly with a resource state on B that -close to
be being symmetric, or simulated approximately
to the same threshold with a perfectly symmet-
ric state on B. Thus the use of isotropy groups
within the resource theory context is robust under
such small perturbations and the speck of dust
argument is moot. We illustrate this point ex-
plicitly in Section 4, where we demonstrate that
for the case of two qubits under SU(2) the notion
of different ‘shapes’ is perfectly robust under the
particular case of perturbations that allow a 4%
error rate.
3.4 Projecting out residual symmetries via
quantum operations
Given the structure of states under the above
partition, we can ask how easy it is to move
from one set C(H) to another. As discussed
one does not in general have quantum operations
that map any C(H) neatly into some other sub-
set. Instead it makes sense to consider the sets
ˆ
C(H) :=
S
W H
C(W ). These sets are quite nat-
ural to consider because Lemma 2 implies that
each
ˆ
C(H) is closed under symmetric operations,
and any symmetric operation provides a well-
Accepted in Quantum 2019-01-23, click title to verify 8
defined mapping of
ˆ
C(H)
ˆ
C(f (H)).
The quantum operation
P
H
(ρ) :=
Z
H
dh U
h
(ρ) (8)
is the average of the state ρ over a fixed subgroup
H weighted by the invariant Haar measure dh.
In the case of a finite subgroup of size |H|, the
integral
R
H
dh is replaced by the sum
1
|H|
P
hH
.
Lemma 4. The map P
H
has the following prop-
erties:
1. P
H
is the (orthogonal) projector onto
ˆ
C(H).
2. P
H
(ρ) = arg min
σ
ˆ
C(H)
S(ρ||σ), where
S(ρ||σ) is the relative entropy.
Therefore P
H
projects onto
ˆ
C(H), as illus-
trated in Fig 2. Dephasing [20], E(ρ) =
1
2π
R
2π
0
e
iθZ
ρe
iθZ
, can be viewed as P
U(1)
,
sending ρ to the nearest incoherent state while
preserving the diagonal terms of a density ma-
trix.
The mapping P
H
moves down chains of the
subgroup Hasse diagram, however this is not al-
ways a symmetric operation. The following tells
us when such a transformation can be performed
freely in the resource theory.
Lemma 5. Given a group action for G,
Iso(P
H
) = N
G
(H), where N
G
(H) = {g G :
gHg
1
= H} is the normalizer [39] of H in G,
and therefore if H is a normal subgroup of G
(H G) then P
H
is a symmetric operation.
This constrains the kind of resources needed to
move from one C(H) to another using P
H
, since
this projects onto a given
ˆ
C(H), although less
resource-hungry ways may exist to perform the
same transformation.
3.5 Discussion of the section results
In this section we have given an analysis of the
what happens when one classifies quantum states
or channels in terms of their residual symmetries.
We discussed how composition and mixing affect
the ordering, and how one can relate these fea-
tures to the issue of simulation (exact or approx-
imate) of a target quantum operation with some
resource state. We also saw that the ordering
has subtleties if one wishes to interpret it in a re-
source theoretic sense. The statements one makes
are also quite blunt, and a good example of this
is the fact that the sets {C(H)} are all arbitrarily
close to one another. However the relations still
carry non-trivial content, and for example Theo-
rem 1 can be viewed as a form of superselection
rule on quantum operations that tells us which
states are ruled out and which are not.
While this is conceptually neat and provides
high-level insight into the complexities of the full
classification of states (as described in [7]), it is
less clear how computationally useful or simple
these structures are in practice. To address this
point, in the next section we consider the impor-
tant case of a two-qubit system under an SU (2)
action. We find that already in just this sim-
ple scenario the hierarchy of states is quite com-
plex, however the example does provide insight
into what to expect in the more general case.
4 Illustrative Example: Two Qubits
Under SU(2) Symmetry Constraints
The goal of this section is to illustrate the clas-
sification of quantum states in a simple quan-
tum system that has sufficient structure, yet is
tractable to the point of being fully solvable.
The state space of a two-qubit system is 15 di-
mensional and is sufficiently non-trivial, more-
over there is a very natural group action to con-
sider, namely the tensor product representation
of SU(2). The orbits of 2-qubit states have been
studied in relation to thermodynamics and corre-
lations within these states [40, 41].
Any two qubit state ρ
AB
can be written
ρ
AB
=
1
4
1 1 + a · σ 1 + 1 b · σ
+
3
X
i,j=1
T
ij
σ
i
σ
j
, (9)
with local Bloch vectors a and b and correlation
terms determined by the correlation matrix T
ij
[42]. In general |a| 1 and |b| 1. Moreover,
we can put all 2-qubit states into diagonal form,
ρ
AB
=
1
4
1 1 + a · σ 1 + 1 b · σ
+
3
X
i=1
τ
i
c
i
· σ d
i
· σ
!
(10)
Accepted in Quantum 2019-01-23, click title to verify 9
where {c
i
} and {d
i
} are orthonormal bases
of R
3
. The coefficients τ
i
specify a posi-
tion (τ
1
, τ
2
, τ
3
) in a tetrahedron with vertices
(1, 1, 1), (1, 1, 1), (1, 1, 1) and (1, 1, 1).
When a = b = 0, then ρ
AB
is a quantum state
with maximally mixed marginals represented by
a point in the tetrahedron. In the case a or b
being non-zero, every quantum state ρ
AB
corre-
sponds to a point in the tetrahedron, however the
converse is not true: not all triples (a, b, t) cor-
respond to a valid quantum states.
Via local unitaries U
1
U
2
, these can be trans-
formed into the canonical form
˜ρ
AB
=
1
4
1 1 + a
0
· σ 1 + 1 b
0
· σ
+
3
X
i=1
τ
i
σ
i
σ
i
!
(11)
where a
0
·σ = U
1
(a·σ)U
1
and b
0
·σ = U
2
(b·σ)U
2
.
The states in this canonical form with maximally
mixed marginals are called T-states [42],
ρ
T
=
1
4
1 1 +
X
j
τ
j
σ
j
σ
j
. (12)
The set of T-states is the convex hull of the four
Bell states. The corners of the tetrahedron de-
fined by the coefficients {τ
i
} correspond to the
Bell states.
We now consider the symmetry properties of 2-
qubit states under SU (2) symmetry constraints,
assuming an SU(2) tensor product representa-
tion for the group action, i.e. U
g
(ρ) = U(g)
U(g) ρ U
(g) U
(g), where U (g) is a two di-
mensional irrep of SU(2).
Our description of the group orbit as the quo-
tient space G/Iso(ρ) starts with the group man-
ifold of G and eliminates the redundancy caused
by the symmetries of ρ. The group transfor-
mations in Iso(ρ) have trivial group action on
ρ, so we quotient out the equivalence classes
g Iso(ρ) := {gh
i
: h
i
Iso(ρ)}.
Certain groups have structure which permits
us to further simplify our description of the group
orbit manifold. Here we consider SU(2) symme-
try constraints, however this SU(2) action has a
subgroup Z
2
= 1} that is always a symmetry
of a quantum state, and which is associated with
SU(2) being the double cover of SO(3). More-
over this sub-group has a key property which al-
lows us to simplify our description of the SU(2)
isotropy groups to those of SO(3). We can there-
fore think in terms of SO(3) subgroups, which are
the familiar chiral point groups one encounters in
for example crystallography [1].
A subgroup N of G is a normal subgroup, de-
noted N G, if it satisfies gNg
1
= N for all
g G [1]. These partition G into equivalence
classes gN := {gn
i
: n
i
N}, which themselves
constitute elements of the quotient group G/N.
Every element of G is specified by the pairing of
the equivalence class gN and an element of the
normal subgroup N. Likewise, elements of H are
specified by the pairing of an element of H/N and
an element of N because N H. This means that
the group orbit manifold G/H can be described
by the quotient groups G/N and H/N, and we
can write M(ρ)
=
G/H
=
(G/N )/(H/N ). In
our present analysis Z
2
is a normal sub-group
of SU(2) common to all quantum states and so
M(ρ)
=
SO(3)/(H/Z
2
) where H is the isotropy
group of ρ in SU(2). We can therefore use the
more intuitive SO(3) subgroups, listed in Table
1.
Group Description
C
n
Cyclic group of order n
C
Symmetries of Cone
D
n
Symmetries of a regular n-sided Polygon
D
Symmetries of Cylinder
T Symmetries of Tetrahedron
O Symmetries of Cube/Octahedron
I Symmetries of Dodecahedron/Icosahedron
Table 1: Chiral point groups: subgroups of SO(3),
as detailed in [1]. All quantum states have isotropy sub-
groups under SU(2) that can be related to particular
chiral point group.
In the next section we look at the isotropy sub-
groups of the Bell states, and then go on to
classify the symmetry properties of the T-states.
Then we continue onto the wider class of 2-qubit
states with maximally mixed marginals and fi-
nally we complete the classification for general 2-
qubit states by considering non-zero local Bloch
vectors.
4.1 Bell States
The Bell states correspond to the extremal points
of the tetrahedron of T-states. Their isotropy
subgroups under the SU (2) group action can be
Accepted in Quantum 2019-01-23, click title to verify 10
calculated directly, as described in Appendix B.
The singlet Bell state ψ
is symmetric un-
der SU(2), therefore Iso(ψ
) = SU(2) and so
M(ψ
) = {e} as expected. The triplet Bell
states do break the SU(2) symmetry, with
Iso(φ
+
) = {(iZ)
α
e
iθY
: 0 θ < 2π, α = 0, 1, 2, 3}.
(13)
We will call this subgroup K
=
U(1) o Z
4
, and
it provides an example of a semi-direct product
group
2
. Similarly,
Iso(φ
) = {(iY)
α
e
iθX
: 0 θ < 2π, α = 0, 1, 2, 3}
(14)
and
Iso(ψ
+
) = {(iX)
α
e
iθZ
: 0 θ < 2π, α = 0, 1, 2, 3},
(15)
both of which are isomorphic to K
.
Given the triplet Bell states have isomorphic
isotropy subgroups, their group orbits will have
the same shape, M(φ
+
)
=
M(φ
)
=
M(ψ
+
)
=
SU(2)/K
. In the previous section we described
how normal subgroups simplify our description
of the group orbit, and we can use this here
to put the description in terms of SO(3) sub-
groups. There is a 2-to-1 homomorphism from
the quotient G/{1, 1}. Elimination of the Z
2
normal subgroup from the group orbit descrip-
tion gives SU (2)/K
=
SO(3)/D
, where D
is the isotropy subgroup of the cylinder.
4.2 T-States
From the Bell states, we can begin to map out the
possible shapes of group orbit in the tetrahedron
2
The semi-direct product [1] is constructed from two
groups, in this case Z
4
and U (1). Each element of the
semidirect product group can be thought of as a pairing
(h, n) where h H = Z
4
and n N = U(1). However the
multiplication law of the semidirect product group does
not treat the elements of this pairing independently, with
(h
1
, n
1
)(h
2
, n
2
) = (h
1
h
2
, f
h
2
(n
1
)n
2
) where f
h
: N N
is a automorphism on the second group specified by an
element of the first group. The semi-direct product o
is defined by the choice of automorphism f
h
. Note that
N / (N o H). If a trivial (identity) automorphism is cho-
sen, where f
h
(n) = n for any h and n, we have a direct
product group [1]. In the U (1) o Z
4
example, the group
multiplication law is
(iZ)
α
1
e
1
Y
(iZ)
α
2
e
2
Y
= (iZ)
α
1
+α
2
e
i[(1)
α
2
θ
1
+θ
2
]Y
.
of T-states. Consider the convex hull of triplet
Bell states to be the base of the tetrahedron, and
ψ
to be the peak. The vertical height above the
base of this tetrahedron indicates the proportion
of the singlet state in the convex mixture p
1
φ
+
+
p
2
φ
+ p
3
ψ
+
+ p
4
ψ
, where p
1
+ p
2
+ p
3
+ p
4
= 1.
Since ψ
is symmetric under the SU (2) group
action,
M(p
1
φ
+
+ p
2
φ
+ p
3
ψ
+
+ p
4
ψ
)
= M(p
1
φ
+
+ p
2
φ
+ p
3
ψ
+
), (16)
therefore the possible types of group orbit for T-
states will all be exhibited in the convex hull of
the triplet states. These can be visualized as a
triangle of states, as shown in Fig. 5.
The vertices of this triangle are the triplet Bell
states, and each have a K
isotropy subgroup.
Appendix B also shows that the midpoints of
the edges of this triangle, corresponding to states
1
4
(1 1 + σ
i
σ
i
), also have K
isotropy sub-
groups. Therefore the diagonals of this triangle,
corresponding to states
ρ =
+
+ (1 p)
1
4
(1 1 + Y Y ) (17)
ρ =
+ (1 p)
1
4
(1 1 + X X) (18)
ρ =
+
+ (1 p)
1
4
(1 1 + Z Z) (19)
for 0 p < 1, have K
Iso(ρ). The states
not on these diagonals remain invariant under the
subgroup K
2
= 1, ±iX, ±iY, ±iZ}.
The possible isotropy subgroups exhibited by
the T-states can be seen with the projectors P
H
,
as detailed in Appendix C. These are:
Iso(ρ
T
) = SU(2) M(ρ)
=
{e}
Werner states along the central axis of the
T-state tetrahedron, where τ
1
= τ
2
= τ
3
.
These are the states
ρ = (1 p) ψ
+
p
3
(ψ
+
+ φ
+
+ φ
) (20)
for 0 p 1.
Iso(ρ
T
) = K
M(ρ)
=
SO(3)/D
T-states with τ
i
= τ
j
6= τ
k
. These states lie
on or above the diagonals of the triangle of
triplet Bell states, as seen in Fig.5, excluding
the Werner states on the central axis of the
tetrahedron.
Accepted in Quantum 2019-01-23, click title to verify 11
Figure 5: The partitioning of the 2-qubit state space. a) Tetrahedron of T-states coloured to indicate different
group orbit shapes: grey SU(2)/K
2
=
SO(3)/D
2
, light blue SO(3)/D
, black {e}. Slices through the
tetrahedron that are parallel to the plane containing the triplet Bell states exhibit the same structure, exhibited
in the triangle to the right. b) Tetrahedron mapping out the smoothed isotropy subgroups with smoothing scale
= 0.04. The bands along the diagonals indicate Iso
(ρ)
=
K
, while states in the central hexagon are assigned
Iso
(ρ)
=
SU(2). The remaining states retain their K
2
isotropy subgroup. c) Hasse diagram of the observed isotropy
subgroups (up to isomorphism) for two qubits. Arrows indicate subset inclusion. This is a sublattice of the full SU(2)
subgroup lattice.
Iso(ρ
T
) = K
2
M(ρ)
=
SO(3)/D
2
T-states with τ
1
6= τ
2
6= τ
3
. These are states
that do not lie on or above the diagonals of
the triangle formed by the triplet Bell states.
The smoothed isotropy subgroups are illustrated
in the lower part of Fig. 5, with a smoothing
scale of = 0.04 under the trace distance. Under
this smoothing, the horizontal slices through the
tetrahedron are no longer equivalent; near ψ
all
states are assigned SU(2) as their isotropy sub-
group, while at the base there remain many states
with Iso
(ρ) = K
2
.
4.3 Maximally Mixed Marginals
The set of 2-qubit states with maximally mixed
marginals is a wider class of states reached
from the T-states through local SU(2) unitaries
U(g
1
) U(g
2
). Appendix C details how the
isotropy subgroups are enumerated.
If these local unitaries are the same for each
qubit, we reach states of the form
˜ρ
T
= U(g) U(g) ρ
T
U
(g) U
(g) = U
g
(ρ
T
)
=
1
4
1 1 +
3
X
i=1
τ
i
c
i
· σ c
i
· σ
!
, (21)
where the vectors {c
i
} form an orthonormal basis
of R
3
.
These states behave similarly to the T-states,
but with isomorphic isotropy subgroups. This
can be seen from the final part of Lemma 1, where
Iso(U
g
ρ
T
U
g
1
) = g[Iso(ρ
T
)]g
1
. Explicitly,
instead of 1, ±iX, ±iY, ±iZ}, these rotated
T-states will be invariant under 1, ±i(c
1
·
σ), ±i(c
2
· σ), ±i(c
3
· σ)}
=
K
2
.
The possible types of group orbit for these
states are the same as for the T-states:
Iso(˜ρ
T
) = SU(2) M(˜ρ
T
)
=
{e}
The same symmetric states as the T-states,
unchanged by the unitary transformation
Accepted in Quantum 2019-01-23, click title to verify 12
U(g) U(g). These are states ˜ρ
T
with
τ
1
= τ
2
= τ
3
.
Iso(˜ρ
T
) = K
M(˜ρ
T
)
=
SO(3)/D
States ˜ρ
T
with τ
i
= τ
j
6= τ
k
. These are
reached via unitary transformations U(g)
U(g) from the T-states with K
isotropy
subgroups.
Iso(˜ρ
T
) = K
2
M(˜ρ
T
)
=
SO(3)/D
2
States ˜ρ
T
with τ
1
6= τ
2
6= τ
3
. These are
reached via unitary transformations U(g)
U(g) from the T-states with K
2
isotropy sub-
groups.
Further states with maximally mixed marginals
are reached from the T-states by different local
unitaries on each qubit. Suppose that these two
local unitaries have the same generator r · σ, such
that U(g
1
) U(g
2
) = e
1
r·σ
e
2
r·σ
. This gives
states of the form
ρ
M
= e
1
r·σ
e
2
r·σ
ρ
T
e
1
r·σ
e
2
r·σ
=
1
4
1 1 + τ
1
r · σ r · σ +
3
X
i=2
τ
i
c
i
· σ d
i
· σ
!
(22)
where {r, c
2
, c
3
} and {r, d
2
, d
3
} are two sets of
orthonormal bases for R
3
. If U(g
1
) = U(g
2
) in the
local unitaries, these sets of bases will be the same
and we retrieve ˜ρ
T
. However, in the case that
θ
1
6= θ
2
in the local unitaries, new possibilities
for the isotropy subgroup are introduced:
Iso(ρ
M
) = U(1) M(ρ
M
)
=
S
2
States ρ
M
with τ
2
= τ
3
. These are reached
from T-states with K
isotropy subgroups
via local unitaries e
1
r·σ
e
2
r·σ
where θ
1
6=
θ
2
.
Iso(ρ
M
) = Z
4
M(ρ
M
)
=
SO(3)/C
2
States ρ
M
with τ
2
6= τ
3
. These are reached
from T-states with K
2
isotropy subgroups
via local unitaries e
1
r·σ
e
2
r·σ
where θ
1
6=
θ
2
.
The states ˜ρ
T
are only invariant under a K
2
sub-
group because the correlation terms are put into
diagonal form by the same orthonormal bases
{c
i
} and {d
i
} on each qubit. For example, the
correlation terms of the T-states are put into
the canonical form by the standard cartesian or-
thonormal basis {
ˆ
x,
ˆ
y,
ˆ
z} on each qubit. When
the local unitaries are applied, this can break the
K
2
symmetry, and the local bases become ‘mis-
aligned’. However, when c
1
= d
1
= r there is
still a subgroup Z
4
= 1, ±i(r · σ)} that leaves
these states invariant.
The remaining 2-qubit states with maximally
mixed marginals are reached from the T-states
through local SU(2) unitaries that do not share
a generator, such that U(g
1
) U(g
2
) = e
1
r
1
·σ
e
2
r
2
·σ
where |r
1
| = |r
2
| = 1 and r
1
6= r
2
. These
states can be expressed in the form
˜ρ
M
= e
1
r
1
·σ
e
2
r
2
·σ
ρ
T
e
1
r
1
·σ
e
2
r
2
·σ
=
1
4
1 1 +
3
X
i=1
τ
i
c
i
· σ d
i
· σ
!
, (23)
where {c
i
} and {d
i
} are orthonormal bases of R
3
that do not share any elements, i.e. c
i
6= d
j
for
all i and j. These states exhibit another type of
group orbit:
Iso(ρ) = Z
2
M(ρ)
=
SO(3)
States ˜ρ
M
. The orthonormal bases {c
i
} and
{d
i
} do not share any elements, therefore the
only subgroup that leaves these states invari-
ant is Z
2
= 1}. These are in a sense max-
imally asymmetric states, as they have no
non-trivial residual symmetries.
4.4 Local Bloch Vectors
Finally, the introduction of local Bloch vectors
completes the set of 2-qubit states. In the states
with maximally mixed marginals, the alignment
of the orthonormal bases describing the corre-
lation terms played a key role in determining
the symmetry properties. The effect of the lo-
cal Bloch vectors depends on the how they relate
to each other, as well as how they relate to the
bases for the correlation terms.
There are three possible isotropy subgroups for
these states:
Iso(ρ) = Z
2
M(ρ)
=
SO(3)
This will occur for any states
ρ =
1
4
1 1 + a · σ 1 + 1 b · σ
+
3
X
i=1
τ
i
c
i
· σ d
i
· σ
!
(24)
Accepted in Quantum 2019-01-23, click title to verify 13
whose local Bloch vectors a · σ and b · σ are
not parallel or anti-parallel, such that a 6=
γb for any γ R.
This isotropy subgroup will also occur in
states where the basis in which the correla-
tion terms are diagonalised does not align
with the local Bloch vectors. These are
states of the form
ρ =
1
4
1 1 + a · σ 1 + 1 b · σ
+τ
1
r · σ r · σ +
3
X
i=2
τ
i
c
i
· σ d
i
· σ
!
(25)
where {r, c
2
, c
3
} and {r, d
2
, d
3
} are two sets
of orthonormal bases for R
3
and either a
or b 6= γr for any γ R. In this case,
the correlation terms would be invariant un-
der Z
4
= 1, ±ir · σ}, however the lo-
cal Bloch vectors prevent this. Similarly if
τ
2
= τ
3
in states of this form, then the lo-
cal Bloch vectors break a U(1) symmetry,
while if {c
i
} = {d
i
} then the local Bloch
vectors break a K
2
symmetry. Likewise if
both {c
i
} = {d
i
} and τ
2
= τ
3
, the corre-
lation terms are invariant under a K
sub-
group that is again broken by the local Bloch
vectors.
Iso(ρ) = Z
4
M(ρ)
=
SO(3)/C
2
States of the form
ρ =
1
4
1 1 + ar · σ 1 + b1 r · σ
+τ
1
r · σ r · σ +
3
X
i=2
τ
i
c
i
· σ d
i
· σ
!
(26)
where {r, c
2
, c
3
} and {r, d
2
, d
3
} are two sets
of orthonormal bases for R
3
, a, b R and
τ
2
6= τ
3
. The local Bloch vectors are invari-
ant under the same Z
4
subgroup that leaves
the correlation terms invariant. This also in-
cludes the case where c
i
= d
i
for i = 2, 3.
The correlation terms are invariant under a
K
2
subgroup, but the local Bloch vectors
break this to a Z
4
subgroup.
Similarly, states of the form
ρ =
1
4
1 1 + ac
1
· σ 1 + b1 c
1
· σ
+
3
X
i=1
τ
i
c
i
· σ c
i
· σ
!
(27)
where {c
1
, c
2
, c
3
} is an orthonormal basis for
R
3
and τ
1
= τ
2
6= τ
3
. The correlation terms
are invariant under a K
subgroup, but the
local Bloch vectors break the residual sym-
metry down to a Z
4
subgroup as the local
Bloch vectors are orthogonal to c
3
· σ, the
generator of the relevant U(1) subgroup.
Iso(ρ) = U(1) M(ρ)
=
S
2
These are states
ρ =
1
4
1 1 + ar · σ 1 + b1 r · σ
τ
1
r · σ r · σ + τ
3
X
i=2
c
i
· σ d
i
· σ
!
(28)
where {r, c
2
, c
3
} and {r, d
2
, d
3
} are two sets
of orthonormal bases for R
3
, a, b R. The
local Bloch vectors are invariant under the
same U (1) subgroup that leaves the correla-
tion terms invariant. This also includes the
cases when {c
i
} = {d
i
} and when τ
1
= τ,
where the correlation terms are invariant un-
der K
and SU(2) respectively.
This completes the classification of the isotropy
subgroups of 2-qubit states.
4.5 Some simple consequences of the classifi-
cation
This classification allows us to infer constraints
on the resources required to simulate asymmetric
channels under symmetry constraints. If we wish
to simulate an axial channel (i.e. M(E) = S
2
)
under SU(2) symmetry constraints using the T-
states, the resource state must have M(σ) S
2
.
T-states on the diagonals of the tetrahedron, with
M(ρ) SO(3)/D
, will not be able to per-
form such simulations; only the T-states with
τ
1
6= τ
2
6= τ
3
will achieve this task up to some
approximation as discussed earlier.
However, convex combinations of different T-
states can create more states that break the sym-
metry to a larger degree. For example, Iso(pφ
+
+
Accepted in Quantum 2019-01-23, click title to verify 14
(1p)φ
) = K
2
when p 6= 1/2, despite Iso(φ
+
)
=
Iso(φ
)
=
K
.
Group orbits also constrain the dynamics of
the T-states. For instance, Lemma 2 implies that
symmetric operations move states situated on a
diagonal of the tetrahedron only along that diag-
onal. The T-states are closed under symmetric
dynamics leaving would entail further symme-
tries being broken.
Also note that H
1
H
2
does not imply
there exists a symmetric channel from C(H
1
) into
C(H
2
). T-states on the diagonals of the base
of the tetrahedron cannot be reached from those
that lie off these diagonals under symmetric op-
erations [25], despite K
2
K
. Group orbits
impose the highest level constraints; further tech-
niques such as those of [25] then impose more de-
tailed constraints.
5 Outlook
Our analysis has shown that the basic prob-
lem of state interconversion under a symmetry
constraint has a rich and non-trivial structure.
We have established a high-level description of
the problem that is consistent with the resource-
theoretic picture, however the question whether
the framework can find practical use in the same
way that superselection rules simplify computa-
tion remains to be explored.
We have classified the set of all two-qubit states
under the SU(2) tensor product representation,
which may be of use in its own right for the study
of quantum correlations. However, beyond two-
qubits, the general problem can be extremely dif-
ficult for example even the finite subgroups of
SU(5) remain unclassified [17]. Despite this com-
plexity, we believe that the tools we have devel-
oped here provide useful insight into the struc-
ture of information processing under symmetry
constraints.
Other techniques for studying the effect of sym-
metry constraints on quantum operations con-
nect with the present analysis. As already men-
tioned, recent work on the harmonic analysis of
quantum channels suggests that the geometry of
group orbits provides an insight into the use of
finite quantum resources [25]. The present paper
has focused purely on the ‘shape’ of this orbit,
but the role of the induced geometry on the or-
bits is only partially understood. To extend such
an analysis would require adapting techniques in
metrology [15] and differential geometry [43]. It
would also be of interest to draw connections with
recent work [44] that extends classical coarse-
graining into the quantum regime with symme-
tries present.
Finally, these ideas may inform us about other
resource theories, using the symmetry constraints
directly (e.g. in thermodynamics). Along similar
lines, the techniques we have described may find
application in the context of realising universal
quantum computation via the combination of re-
source states with simple gate-sets, for example
Clifford operations and magic states.
6 Acknowledgements
We would like to thank Cristina Cirstoiu, Erick
Hinds Mingo, Zoe Holmes and Markus Frembs
for many useful discussions. TH is funded by
the EPSRC Centre for Doctoral Training in Con-
trolled Quantum Dynamics. DJ is supported by
the Royal Society.
References
[1] M. Tinkham. Group theory and quantum
mechanics. International series in pure and
applied physics. McGraw-Hill, 1964.
[2] Carlton M. Caves, Christopher A. Fuchs,
and Rüdiger Schack. Unknown quan-
tum states: The quantum de finetti
representation. Journal of Mathemati-
cal Physics, 43(9):4537–4559, 2002. DOI:
10.1063/1.1494475.
[3] R. Jozsa. Quantum factoring, discrete log-
arithms, and the hidden subgroup prob-
lem. Computing in Science Engineering, 3
(2):34–43, Mar 2001. ISSN 1521-9615. DOI:
10.1109/5992.909000.
[4] Ryszard Horodecki, Pawel Horodecki,
Michal Horodecki, and Karol Horodecki.
Quantum entanglement. Rev. Mod.
Phys., 81:865–942, Jun 2009. DOI:
10.1103/RevModPhys.81.865.
[5] I. Marvian and R. W. Spekkens. Ex-
tending Noether’s theorem by quantifying
the asymmetry of quantum states. Nature
Communications, 5:3821, May 2014. DOI:
10.1038/ncomms4821.
Accepted in Quantum 2019-01-23, click title to verify 15
[6] Matteo Lostaglio, Kamil Korzekwa, David
Jennings, and Terry Rudolph. Quan-
tum coherence, time-translation symmetry,
and thermodynamics. Phys. Rev. X, 5:
021001, Apr 2015. DOI: 10.1103/Phys-
RevX.5.021001.
[7] Gilad Gour, David Jennings, Francesco
Buscemi, Runyao Duan, and Iman Mar-
vian. Quantum majorization and a com-
plete set of entropic conditions for quantum
thermodynamics. Nature Communications,
9(1):5352, 2018. ISSN 2041-1723. DOI:
10.1038/s41467-018-06261-7.
[8] M. Lostaglio, D. Jennings, and T. Rudolph.
Description of quantum coherence in ther-
modynamic processes requires constraints
beyond free energy. Nature Commu-
nications, 6:6383, March 2015. DOI:
10.1038/ncomms7383.
[9] E. P. Wigner. Die messung quanten-
mechanischer operatoren. Zeitschrift für
Physik A Hadrons and nuclei, 133(1):101–
108, Sep 1952. ISSN 0939-7922. DOI:
10.1007/BF01948686.
[10] Huzihiro Araki and Mutsuo M. Yanase. Mea-
surement of quantum mechanical operators.
Phys. Rev., 120:622–626, Oct 1960. DOI:
10.1103/PhysRev.120.622.
[11] L. Loveridge and P. Busch. ‘measurement
of quantum mechanical operators’ revisited.
The European Physical Journal D, 62(2):
297–307, Apr 2011. ISSN 1434-6079. DOI:
10.1140/epjd/e2011-10714-3.
[12] Mehdi Ahmadi, David Jennings, and Terry
Rudolph. The wigner–araki–yanase the-
orem and the quantum resource theory
of asymmetry. New Journal of Physics,
15(1):013057, 2013. DOI: 10.1088/1367-
2630/15/1/013057.
[13] I. Marvian and R. W. Spekkens. An
information-theoretic account of the Wigner-
Araki-Yanase theorem. ArXiv e-prints, De-
cember 2012. URL https://arxiv.org/
abs/1212.3378.
[14] Miguel Navascués and Sandu Popescu.
How energy conservation limits our
measurements. Phys. Rev. Lett., 112:
140502, Apr 2014. DOI: 10.1103/Phys-
RevLett.112.140502.
[15] Dominik Šafránek, Mehdi Ahmadi, and
Ivette Fuentes. Quantum parameter estima-
tion with imperfect reference frames. New
Journal of Physics, 17(3):033012, 2015. DOI:
10.1088/1367-2630/17/3/033012.
[16] Iman Marvian, Robert W. Spekkens, and
Paolo Zanardi. Quantum speed limits, co-
herence, and asymmetry. Phys. Rev. A,
93:052331, May 2016. DOI: 10.1103/Phys-
RevA.93.052331.
[17] D. Grier and L. Schaeffer. The Classification
of Stabilizer Operations over Qubits. ArXiv
e-prints, March 2016. URL https://arxiv.
org/abs/1603.03999.
[18] D. Gottesman. Stabilizer codes and quan-
tum error correction. PhD thesis, California
Institute of Technology, 1997. URL https:
//arxiv.org/abs/quant-ph/9705052.
[19] D S França and A K Hashagen. Approx-
imate randomized benchmarking for finite
groups. Journal of Physics A: Mathemati-
cal and Theoretical, 51(39):395302, aug 2018.
DOI: 10.1088/1751-8121/aad6fa.
[20] M.A. Nielsen and I.L. Chuang. Quantum
Computation and Quantum Information:
10th Anniversary Edition. Cambridge Uni-
versity Press, 2010. ISBN 9781139495486.
DOI: 10.1017/CBO9780511976667.
[21] J. A. Vaccaro, F. Anselmi, H. M. Wise-
man, and K. Jacobs. Tradeoff between ex-
tractable mechanical work, accessible entan-
glement, and ability to act as a reference
system, under arbitrary superselection rules.
Phys. Rev. A, 77:032114, Mar 2008. DOI:
10.1103/PhysRevA.77.032114.
[22] Gilad Gour, Iman Marvian, and Robert W.
Spekkens. Measuring the quality of a quan-
tum reference frame: The relative entropy
of frameness. Phys. Rev. A, 80:012307, Jul
2009. DOI: 10.1103/PhysRevA.80.012307.
[23] Gilad Gour and Robert W Spekkens. The re-
source theory of quantum reference frames:
manipulations and monotones. New Jour-
nal of Physics, 10(3):033023, 2008. DOI:
10.1088/1367-2630/10/3/033023.
[24] Iman Marvian and Robert W Spekkens. The
theory of manipulations of pure state asym-
metry: I. basic tools, equivalence classes
and single copy transformations. New Jour-
nal of Physics, 15(3):033001, 2013. DOI:
10.1088/1367-2630/15/3/033001.
[25] C. Cirstoiu and D. Jennings. Irreversibility
and quantum information flow under global
Accepted in Quantum 2019-01-23, click title to verify 16
and local gauge symmetries. ArXiv e-prints,
Dec 2017. URL https://arxiv.org/abs/
1707.09826.
[26] Johan Åberg. Catalytic coherence. Phys.
Rev. Lett., 113:150402, Oct 2014. DOI:
10.1103/PhysRevLett.113.150402.
[27] J. Madore. An Introduction to Noncommu-
tative Differential Geometry and its Physical
Applications. London Mathematical Society
Lecture Note Series. Cambridge University
Press, 1995. ISBN 9780521467919. DOI:
10.1017/CBO9780511569357.
[28] Michael Kech, Peter Vrana, and Michael M
Wolf. The role of topology in quantum to-
mography. Journal of Physics A: Mathemat-
ical and Theoretical, 48(26):265303, 2015.
DOI: 10.1088/1751-8113/48/26/265303.
[29] Victor Veitch, S A Hamed Mousavian,
Daniel Gottesman, and Joseph Emerson.
The resource theory of stabilizer quantum
computation. New Journal of Physics,
16(1):013009, 2014. DOI: 10.1088/1367-
2630/16/1/013009.
[30] R. Schmidt. Subgroup Lattices of Groups.
De Gruyter Expositions in Mathematics.
De Gruyter, 1994. ISBN 9783110868647.
URL https://www.degruyter.com/view/
product/172483.
[31] S. Burris and H.P. Sankappanavar. A
course in universal algebra. Gradu-
ate texts in mathematics. Springer-
Verlag, 1981. ISBN 9780387905785.
URL http://www.math.uwaterloo.ca/
~snburris/htdocs/ualg.html.
[32] Bob Coecke, Tobias Fritz, and Robert W.
Spekkens. A mathematical theory of re-
sources. Information and Computation, 250
(Supplement C):59 86, 2016. ISSN 0890-
5401. DOI: 10.1016/j.ic.2016.02.008.
[33] Alexander Streltsov, Gerardo Adesso, and
Martin B. Plenio. Colloquium: Quantum co-
herence as a resource. Rev. Mod. Phys., 89:
041003, Oct 2017. DOI: 10.1103/RevMod-
Phys.89.041003.
[34] John Goold, Marcus Huber, Arnau Ri-
era, Lídia del Rio, and Paul Skrzypczyk.
The role of quantum information in ther-
modynamics—a topical review. Journal of
Physics A: Mathematical and Theoretical,
49(14):143001, 2016. DOI: 10.1088/1751-
8113/49/14/143001.
[35] I. Bengtsson and K. Życzkowski. Geome-
try of Quantum States: An Introduction to
Quantum Entanglement. Cambridge Univer-
sity Press, 2017. ISBN 9781107026254. DOI:
10.1017/CBO9780511535048.
[36] A. Arvanitoge¯orgos. An Introduction to Lie
Groups and the Geometry of Homogeneous
Spaces. Student mathematical library. Amer-
ican Mathematical Society, 2003. ISBN
9780821827789. DOI: 10.1090/stml/022.
[37] G.E. Bredon. Introduction to Compact
Transformation Groups. Pure and Applied
Mathematics. Elsevier Science, 1972. ISBN
9780080873596.
[38] Dennis Kretschmann, Dirk Schlingemann,
and Reinhard F. Werner. A continuity
theorem for stinespring’s dilation. Jour-
nal of Functional Analysis, 255(8):1889
1904, 2008. ISSN 0022-1236. DOI:
10.1016/j.jfa.2008.07.023.
[39] I.M. Isaacs. Algebra: A Graduate Course.
Graduate studies in mathematics. Amer-
ican Mathematical Society, 1994. ISBN
9780821847992.
[40] Sania Jevtic, David Jennings, and Terry
Rudolph. Quantum mutual information
along unitary orbits. Physical Review A,
85(5):052121, 2012. DOI: 10.1103/Phys-
RevA.85.052121.
[41] Sania Jevtic, David Jennings, and Terry
Rudolph. Maximally and minimally cor-
related states attainable within a closed
evolving system. Physical review letters,
108(11):110403, 2012. DOI: 10.1103/Phys-
RevLett.108.110403.
[42] Ryszard Horodecki and Michal Horodecki.
Information-theoretic aspects of inseparabil-
ity of mixed states. Phys. Rev. A, 54:
1838–1843, Sep 1996. DOI: 10.1103/Phys-
RevA.54.1838.
[43] S. Amari and H. Nagaoka. Methods of In-
formation Geometry. Translations of math-
ematical monographs. American Mathemat-
ical Society, 2007. ISBN 9780821843024.
[44] Oleg Kabernik. Quantum coarse graining,
symmetries, and reducibility of dynamics.
Phys. Rev. A, 97:052130, May 2018. DOI:
10.1103/PhysRevA.97.052130.
[45] M. Keyl and R. F. Werner. Optimal cloning
of pure states, testing single clones. Journal
Accepted in Quantum 2019-01-23, click title to verify 17
of Mathematical Physics, 40(7):3283–3299,
1999. DOI: 10.1063/1.532887.
[46] Marvian Mashhad, Iman. Symmetry, Asym-
metry and Quantum Information. PhD the-
sis, 2012. URL http://hdl.handle.net/
10012/7088.
[47] A. Gilchrist, D. R. Terno, and C. J. Wood.
Vectorization of quantum operations and its
use. ArXiv e-prints, November 2009. URL
https://arxiv.org/abs/0911.2539.
Accepted in Quantum 2019-01-23, click title to verify 18
A Proofs
Lemma 1. Given any two quantum channels E : B(H
A
) B(H
A
0
) and F : B(H
B
) B(H
B
0
),
with unitary representations of a group G defined on all input and output spaces. Then we have the
following
1. Iso(E F) Iso(E) Iso(F) under the tensor product group action U
g
(E F) = U
g
(E) U
g
(F).
2. If B
0
= A then Iso(E F) Iso(E) Iso(F).
3. If A = B and A
0
= B
0
, and p some some probability 0 p 1, then Iso(pE + (1 p)F)
Iso(E) Iso(F).
4. Iso(U
g
E U
g
) = g[Iso(E)]g
1
.
Proof. For g Iso(E) Iso(F), the global group action gives U
g
(E F) = U
g
(E) U
g
(F) = E F,
therefore g Iso(E F). However E F may have further symmetries from taking the tensor product
(for example Iso(|0i h0| |0i h0|) Iso(|0i h0|)), therefore Iso(E F) Iso(E) Iso(F).
The group action on E F is given by U
g
(E F) = U
g
E F U
g
1
= U
g
E U
g
1
U
g
F U
g
1
=
U
g
(E)U
g
(F). Therefore if g Iso(E)Iso(F), then g Iso(E◦F), implying Iso(E◦F) Iso(E)Iso(F).
If g Iso(E) Iso(F), then U
g
(pE + (1 p)F) = p U
g
(E) + (1 p) U
g
(F) = pE + (1 p)F, implying
Iso(E) Iso(F) Iso(pE + (1 p)F).
For h Iso(U
g
E U
g
1
), we have U
h
U
g
E U
g
1
U
h
1
= U
g
E U
g
1
which simplifies to
U
(g
1
hg)
E U
(g
1
hg)
1
. Therefore g
1
hg Iso(E) or h g[Iso(E)]g
1
. When gh
0
g
1
g[Iso(E)]g
1
acts upon U
g
E U
g
1
, we have (U
g
U
h
0
U
g
1
)(U
g
E U
g
1
)(U
g
U
h
0−1
U
g
1
) = U
g
U
h
0
E U
h
0−1
U
g
1
= U
g
EU
g
1
, therefore gh
0
g
1
Iso(U
g
EU
g
1
). Therefore Iso(U
g
EU
g
1
) = g[Iso(E)]g
1
.
Theorem 1. If a system B in a state σ
B
can be used to simulate a CPTP map E under symmetric
dynamics, then Iso(σ
B
) Iso(E). Conversely, if E has isotropy group Iso(E) then there exists a
quantum system B and quantum state σ
B
that can be used to simulate E with Iso(σ
B
) = Iso(E).
Proof. Viewing the state σ
B
as a CPTP map, we can view the simulation as a composite channel
E = F σ
B
. Therefore Iso(E) = Iso(F σ
B
) Iso(F) Iso(σ
B
) = Iso(σ
B
) since F is symmetric.
Therefore Iso(E) Iso(σ
B
), so M(σ
B
) M(E).
Alternatively, from the Covariant Stinespring theorem [45, 46], we can write E(ρ) = F(ρ σ
B
) =
tr
BC
[V(ρ σ
B
γ
C
)V
], where γ
C
= |ηihη| is some pure state, [V, U
A
(g) U
B
(g) U
C
(g)] = 0 and
Iso(γ
C
) = G. Supposing g Iso(σ
B
),
E(ρ) = tr
BC
[V (ρ U
g
(σ
B
) U
g
(γ
C
))V
] = tr
BC
[V (1 U
g
U
g
)(ρ σ
B
γ
C
)V
]. (29)
V is a symmetric unitary, so V (1U
g
U
g
) = V (U
g
U
g
U
g
)(U
g
11) = (U
g
U
g
U
g
)V (U
g
11),
giving
E(ρ) = tr
BC
[U
g
(V (U
g
1
(ρ) σ
B
γ
C
)V
)] = U
g
(tr
BC
[V (U
g
1
(ρ) σ
B
γ
C
)V
])
= U
g
E U
g
1
(ρ). (30)
Therefore Iso(σ
B
) Iso(E), and M(σ
B
) M(E).
We now establish the second part of the theorem, and construct the protocol that realises E with
a quantum state that has the same isotropy subgroup as E. Firstly, we note that we can decompose
any E into irreducible process modes [25] as
E =
X
λ,k
α
λ
,k
(x
λ,k
(31)
where λ labels irreps, including multiplicities and k labels the basis vector of the irrep. As shown in
[25] the coefficients α
λ,k
(x) are non-normalised harmonic wavefunctions on the space M := G/Iso(E),
Accepted in Quantum 2019-01-23, click title to verify 19
which is diffeomorphic to the orbit of E under the group G, and where the point x corresponds to the
reference data needed to simulate E under the symmetry constraint.
For our reference system B, we take its Hilbert space to be L
2
(M, C), where σ
B
= |ψihψ| is a
reference state with wavefunction ψ(z) = δ(z x), understood in the distributional sense. The
protocol to realise E via σ
B
involves a measurement of the position z of B, which provides the
necessary reference frame data to realise E on the system A. Explicitly, we have
E =
X
λ,k
α
λ
,k
(x) Φ
λ,k
= tr
B
X
λ,k
|xi
B
hx| α
λ
,k
(x) Φ
λ,k
=
X
λ,k
Z
dz tr[M(z) σ
B
] α
λ
,k
(z) Φ
λ,k
(32)
where {dzM(z)} is the position measurement on M, which is covariant since G acts transitively on
M. Finally, the action of G on σ
B
is given by the action of G on the quotient space G/Iso(E) where
at each point we have an isotropy group Iso(E), and thus Iso(σ
B
) = Iso(E) as required.
Lemma 2. Under a symmetric operation E, Iso(E(ρ)) Iso(ρ).
Proof. We can view the state ρ as a CPTP map 1 ρ, therefore from Lemma 1, Iso(E(ρ)) =
Iso(E ρ) Iso(E) Iso(ρ). Since G H = H for any subgroup H G, then Iso(E) Iso(ρ) = Iso(ρ),
therefore Iso(E(ρ)) Iso(ρ), and thus the group orbits also obey M(E(ρ)) M(ρ).
Lemma 3. Let d(·, ·) be any metric on the space of quantum states. In terms of this metric we define
d(C(H
1
), C(H
2
)) := inf
σ
1
C(H
1
)
σ
2
C(H
2
)
d(σ
1
, σ
2
). (33)
Then d(C(H
1
), C(H
2
)) = 0 for all H
1
, H
2
G.
Proof. For a state ρ
1
C(H
1
), U
g
(ρ
1
+ (1 ) σ) = U
g
(ρ
1
) + (1 ) U
g
(σ) = U
g
(ρ
1
) + (1 ) σ
when σ C(G). Hence Iso(ρ
1
+ (1 )σ) = Iso(ρ
1
), and ρ
1
+ (1 )σ C(H
1
). Likewise,
for ρ
2
C(H
2
), the state ρ
2
+ (1 )σ (with the same σ C(G)) is also in C(H
2
). As 0,
d(ρ
1
+ (1 )σ, ρ
2
+ (1 )σ) 0, therefore d(C(H
1
), C(H
2
)) = 0 for all H
1
, H
2
G.
Lemma 4. The map P
H
has the following properties:
1. P
H
is the (orthogonal) projector onto
ˆ
C(H).
2. P
H
(ρ) = arg min
σ
ˆ
C(H)
S(ρ||σ), where S(ρ||σ) is the relative entropy.
Proof. For states ρ C(W ) for W H, the unitary U(h) for h H will leave the state unchanged.
Therefore P
H
(ρ) =
R
H
dh U
h
(ρ) =
R
H
dh ρ = ρ. Moreover, P
H
is a projector because P
H
P
H
=
R
H
dh
R
H
dh
0
U
h
◦U
h
0
=
R
H
dh
R
H
dh
0
U
hh
0
=
R
H
dh
00
U
h
00
= P
H
. Furthermore, P
H
(ρ) =
R
H
dh U
h
1
(ρ) =
R
H
dh U
h
(ρ) = P
H
(ρ), meaning P
H
is an orthogonal projector.
We follow the proof in [22]. By Klein’s Inequality, we have S(P
H
(ρ)||σ) 0, with equality iff
σ = P
H
(ρ), therefore min
σ
ˆ
C(H)
S(P
H
(ρ)||σ) = 0. From the definition of relative entropy, S(P
H
(ρ)) =
min
σ
ˆ
C(H)
[Tr(ρ log σ)] because log is analytic and σ
ˆ
C(H). Using the idempotency of P
H
,
S(ρ||P
H
(ρ)) = S(ρ) Tr(ρ log P
H
(ρ)) = S(ρ) + S(P
H
(ρ)) = S(ρ) + min
σ
ˆ
C(H)
[Tr(ρ log σ)] =
min
σ
ˆ
C(H)
S(ρ||σ).
Lemma 5. Given a group action for G, Iso(P
H
) = N
G
(H), where N
G
(H) = {g G : gHg
1
= H}
is the normalizer [39] of H in G, and therefore if H is a normal subgroup of G (H G) then P
H
is a
symmetric operation.
Accepted in Quantum 2019-01-23, click title to verify 20
Proof. The group action on P
H
is U
g
(P
H
) =
R
H
dh U
g
(U
h
) =
R
H
dh U
g
U
h
U
g
1
=
R
H
dh U
ghg
1
.
An element g G will therefore be a member of Iso(P
H
) iff gHg
1
= H. This is the definition of the
normalizer, so we have that Iso(P
H
) = N
G
(H). Averaging over a normal subgroup H G is therefore
symmetric, because N
G
(H) = G by the definition of a normal subgroup.
B Bell States and Single Component T-States
The extremal points of the T-states are the Bell states, which can be expressed as the vectorisation
[47] of Pauli matrices:
φ
+
E
|vec(1)i (34)
φ
|vec(Z)i (35)
ψ
+
E
|vec(X)i (36)
ψ
|vec(Y)i . (37)
Elements of the isotropy subgroup of a Bell state therefore satisfy U
g
(|vec(σ
i
)i hvec(σ
i
)|) =
|vec(σ
i
)i hvec(σ
i
)|, hence
U(g) U(g) |vec(σ
i
)i = e
|vec(σ
i
)i (38)
where φ R is some arbitrary phase. From the identity A B |vec(M)i =
vec(AMB
T
)
E
[47], the
residual symmetries satisfy
U(g)σ
i
U
T
(g) = e
σ
i
. (39)
We now consider each of the Bell states in turn.
B.1 Iso(φ
+
)
=
K
φ
+
|vec(1)i is invariant under group transformations satisfying U(g)U
T
(g) = e
1. SU(2) group
elements take the form U (g) = exp([aX +bY +cZ]) with a
2
+b
2
+c
2
= 1 (a, b, c R) and 0 ϕ < 2π,
therefore
e
(aX+bY +cZ)
e
(aXbY +cZ)
= e
1 = e
(aX+bY +cZ)
= e
e
(aX+bY cZ)
. (40)
Expansion of the matrix exponentials gives
cos ϕ1 + i sin ϕ(aX + bY + cZ) = e
[cos ϕ1 + i sin ϕ(aX + bY cZ)] (41)
which provides the conditions
cos ϕ(1 e
) = 0 (42)
ia sin ϕ(1 + e
) = 0 (43)
ib sin ϕ(1 e
) = 0 (44)
ic sin ϕ(1 + e
) = 0. (45)
The problem simplifies because φ R is restricted in the values it may take. Suppose cos ϕ 6= 0, then
cos ϕ(1 e
) = 0 implies φ = 0 (modulo 2π). Now suppose cos ϕ = 0, ensuring that sin ϕ 6= 0. Since
a
2
+ b
2
+ c
2
= 1, at least one of a, b or c is non-zero, therefore either 1 + e
= 0 or 1 e
= 0 in order
for the conditions to hold. Together, these imply that φ = 0 or π (modulo 2π).
When φ = 0, the conditions reduce to
2ia sin ϕ = 0 (46)
2ic sin ϕ = 0. (47)
Accepted in Quantum 2019-01-23, click title to verify 21
One possible solution is a = c = 0, leaving ϕ unconstrained and therefore b = ±1 (without loss of
generality choose b = 1, since b = 1 gives ‘rotations’ of the opposite handedness). These solutions
take the form U(g) = e
iϕY
. The equations are also satisfied when sin ϕ = 0, so when ϕ = 0, π, which
give group transformations U (g) = ±1.
In the φ = π case, the conditions become
2 cos ϕ = 0 (48)
2ib sin ϕ = 0. (49)
If b 6= 0, then cos ϕ = sin ϕ = 0, which is not possible. Therefore the only solutions for these conditions
is b = 0 and cos ϕ = 0, which gives ϕ = π/2 or 3π/2. In this case a and c are only constrained by
a
2
+ c
2
= 1. The ϕ = π/2 solution corresponds to the SU(2) transformations
U(g) = i(aX + cZ) = (iZ)(c1 + iaY ) = (iZ)(cos θ1 + i sin θY ) = (iZ)e
iθY
, (50)
since a
2
+ c
2
= 1, so we parametrise in terms of cos θ and sin θ. Similarly U (g) = (iZ)e
iθY
for the
ϕ = 3π/2 case. This shows that
Iso(φ
+
) = {(iZ)
α
e
iθY
: 0 θ < 2π, α = 0, 1, 2, 3}
=
K
. (51)
B.2 Iso(φ
) = K
Analogous arguments hold for |φ
i |vec(Z)i. The condition for membership of Iso(φ
) is
e
(aX+bY +cZ)
Ze
(aXbY +cZ)
= e
Z (52)
which simplifies to
e
(aX+bY +cZ)
= e
e
(aXbY cZ)
. (53)
This gives the conditions
cos ϕ(1 e
) = 0 (54)
ia sin ϕ(1 e
) = 0 (55)
ib sin ϕ(1 + e
) = 0 (56)
ic sin ϕ(1 + e
) = 0, (57)
solved in the same way as the previous example. The isotropy subgroup is
Iso(φ
) = {(iY)
α
e
iθX
: 0 θ < 2π, α = 0, 1, 2, 3}
=
K
. (58)
B.3 Iso(ψ
+
) = K
ψ
+
|vec(X)i gives the condition
e
(aX+bY +cZ)
= e
e
(aXbY +cZ)
. (59)
This is satisfied when
cos ϕ(1 e
) = 0 (60)
ia sin ϕ(1 + e
) = 0 (61)
ib sin ϕ(1 + e
) = 0 (62)
ic sin ϕ(1 e
) = 0, (63)
which gives the isotropy subgroup as
Iso(ψ
+
) = {(iX)
α
e
iθZ
: 0 θ < 2π, α = 0, 1, 2, 3}
=
K
. (64)
Accepted in Quantum 2019-01-23, click title to verify 22
B.4 Iso(ψ
) = SU(2)
The singlet Bell state |ψ
i |vec(Y)i behaves differently to the other Bell states. The isotropy
subgroup condition is
e
(aX+bY +cZ)
= e
e
(aX+bY +cZ)
, (65)
which is satisfied for all elements of SU(2), therefore Iso(ψ
) = SU(2).
B.5 Iso(
1
4
(1 1 + τ
i
σ
i
σ
i
) = K
For states, ρ =
1
4
(1 1 + τ
i
σ
i
σ
i
), the 1 1 term can be neglected because Iso(1 1 + ρ
0
) = Iso(ρ
0
).
For ρ =
1
4
(1 1 + τ
1
X X), an SU (2) group element belongs to Iso(ρ) when
U
g
(X X) = U
g
(X) U
g
(X) = X X, (66)
satisfied when
e
(aX+bY +cZ)
Xe
(aX+bY +cZ)
= ±X, (67)
or equivalently
e
(aX+bY +cZ)
= ±e
(aXbY cZ)
. (68)
This is the same condition as for ψ
+
, therefore
Iso(X X) = Iso(ψ
+
) = {(iY)
α
e
iθX
: 0 θ < 2π, α = 0, 1, 2, 3}
=
K
. (69)
Likewise, elements of Iso(Y Y) satisfy
e
(aX+bY +cZ)
= ±e
(aX+bY cZ)
, (70)
the same as for φ
+
, therefore
Iso(Y Y) = {(iZ)
α
e
iθY
: 0 θ < 2π, α = 0, 1, 2, 3}
=
K
, (71)
while the condition for membership of Iso(Z Z) is
e
(aX+bY +cZ)
= ±e
(aX+bY cZ)
, (72)
showing that
Iso(Z Z) = {(iX)
α
e
iθZ
: 0 θ < 2π, α = 0, 1, 2, 3}
=
K
. (73)
C 2-Qubit Isotropy Subgroups
The isotropy subgroups of 2-qubit states under a tensor product SU(2) group action can be found
with projection operators P
H
. These do not indicate which subgroups of SU (2) will appear as isotropy
subgroups, however the following lemma restricts the possibilities.
Lemma 6. For a 2-qubit state ρ transforming under an SU(2) tensor product representation, any
cyclic subgroup of the isotropy subgroup Iso(ρ) must be either Z
2
, Z
4
or U(1).
Proof. A group element belongs to Iso(ρ) when U
g
(ρ) = ρ, or in vectorised form [47],
U(g) U(g) U
(g) U
(g) |vec(ρ)i = |vec(ρ)i . (74)
For ρ to be invariant under the group transformation U
g
, it requires |vec(ρ)i to be in ker[U(g)U(g)
U
(g) U
(g) 1].
Accepted in Quantum 2019-01-23, click title to verify 23
All groups contain cyclic subgroups from repeated application of a single generator, because g G
implies that g
n
G. For any cyclic subgroup of Iso(ρ), there is some basis in which its representation
is diagonal, i.e. U (g) = e
|0i h0| + e
|1i h1|. Therefore
U(g) U(g) U
(g) U
(g) 1
= (e
2
1)(|1i h1| + |2i h2| + |7i h7| + |11i h11|) + (e
4
1) |3i h3| + (e
4
1) |12i h12|
+ (e
2
1)(|4i h4| + |8i h8| + |13i h13| + |14i h14|), (75)
where we have assumed for simplicity that U(g) is diagonal in the computational basis (with for
example |2i = |0010i). Vectors in span(|0i , |5i , |6i , |9i , |10i , |15i) are invariant under this U(1)
subgroup. Vectors in span(|1i , |2i , |4i , |7i , |8i , |11i , |13i , |14i) are also in ker[U(g) U(g) U
(g)
U
(g) 1] for θ = 0, π, and are therefore invariant under a Z
2
subgroup. Finally, the vectors of
span(|3i , |12i) are in ker[U(g) U(g) U
(g) U
(g) 1] when θ = 0, π/2, π, 3π/2, and hence
invariant under a Z
4
subgroup.
This holds for any cyclic subgroup of SU(2), therefore any cyclic subgroup of Iso(ρ) must be Z
2
, Z
2
or U(1).
This plays a similar role to the crystallographic restriction theorem in crystallography [1], and
reduces the number of subgroups to check. For example, any 2-qubit state invariant under a Z
6
must
be invariant under the U(1) subgroup with the same generator. This allows us to enumerate the
possible isotropy subgroups for 2-qubit states. It also suggests that an N-qubit state will have only
cyclic subgroups Z
2
, Z
4
, . . . Z
2N
, U(1) within Iso(ρ).
The channels P
H
provide a way to find the sets C(H). Consider two subgroups H
1
and H
2
, with
H
1
H
2
and there are no subgroups H that could be the isotropy subgroup of some state such that
H
1
H H
2
. If a state ρ satisfies P
H
1
(ρ) = ρ but P
H
2
(ρ) 6= ρ, then Iso(ρ) = H
1
, and ρ C(H
1
).
We now use this technique to identify the isotropy subgroups of all 2-qubit states.
C.1 Z
2
Iso(ρ)
All 2-qubit states are symmetric under Z
2
= 1}, confirmed by
P
Z
2
(ρ) =
1
2
[ U
1
(ρ) + U
1
(ρ) ] = ρ (76)
for any 2-qubit state ρ.
C.2 Z
4
Iso(ρ)
From Lemma 6, if a 2-qubit state has Z
3
Iso(ρ), then it also has U(1) Iso(ρ). The next subgroup
to check is Z
4
= 1, ±ir · σ}. For illustrative purposes, let us consider the particular Z
4
subgroup
1, ±iZ}. For a general 2-qubit state ρ,
P
Z
4
(ρ) =
1
4
1 1 + P
Z
4
(a · σ 1) + P
Z
4
(1 b · σ) +
3
X
i,j=1
T
ij
P
Z
4
(σ
i
σ
j
)
(77)
=
1
4
1 1 + a
z
Z 1 + b
z
1 Z + T
33
Z Z +
2
X
i,j=1
T
ij
σ
i
σ
j
. (78)
Now consider the more general Z
4
= 1, ±ir · σ} subgroup. The useful identities
(r · σ)(v · σ) = (r · v)1 + i(r × v) · σ (79)
(r · σ)(v · σ)(r · σ) = 2(r · v)(r · σ) v · σ (80)
Accepted in Quantum 2019-01-23, click title to verify 24
allow us to calculate P
Z
4
(ρ). The local Bloch vectors give
P
Z
4
(a · σ 1) =
1
2
[a · σ 1 + (r · σ)(a · σ)(r · σ) 1] (81)
=
1
2
[a · σ 1 + (2(r · a)(r · σ) a · σ) 1] (82)
= (r · a)(r · σ) 1 (83)
and similarly P
Z
4
(1 b · σ) = (r · b)(1 r · σ). The correlation terms may be written
3
X
i,j=1
T
ij
σ
i
σ
j
=
3
X
i,j=1
T
0
ij
c
i
· σ c
j
· σ, (84)
where {c
i
} is an orthonormal basis for R
3
and we choose c
1
= r. It is clear that P
Z
4
(c
1
· σ c
1
· σ) =
P
Z
4
(r · σ r · σ) = r · σ r · σ. The identity
(r · σ)(u · σ)(r · σ) (r · σ)(v · σ)(r · σ) = 4(r · u)(r · v)r · σ r · σ 2(r · u)r · σ v · σ
2(r · v)u · σ r · σ + u · σ v · σ. (85)
allows us to show that when j 6= 1,
P
Z
4
(c
1
· σ c
j
· σ) = c
1
· σ c
j
· σ + 2(r · c
1
)(r · c
j
)r · σ r · σ (r · c
1
)r · σ c
j
· σ (r · c
j
)c
1
· σ r · σ
= r · σ c
j
· σ r · σ c
j
· σ = 0 (86)
because r · c
j
= c
1
· c
j
= 0. The same holds for the c
j
· σ c
1
· σ terms with j 6= 1, which also vanish
when P
Z
4
is applied. Finally, when i, j 6= 1,
P
Z
4
(c
i
· σ c
j
· σ) = c
i
· σ c
j
· σ + 2(r · c
i
)(r · c
j
)r · σ r · σ (r · c
i
)r · σ c
j
· σ (r · c
j
)c
i
· σ r · σ
= c
i
· σ c
j
· σ. (87)
Therefore
P
Z
4
(ρ) =
1
4
1 1 + (r · a)(r · σ) 1 + (r · b)(1 r · σ) + T
0
11
r · σ r · σ +
3
X
i,j=2
T
0
ij
c
i
· σ c
j
· σ
=
1
4
"
1 1 + (r · a)(r · σ) 1 + (r · b)(1 r · σ) + τ
1
r · σ r · σ +
3
X
i=2
τ
i
c
i
· σ d
i
· σ
#
,
(88)
where {r, c
i
} and {r, d
i
} are orthonormal bases of R
3
.
C.3 Iso(ρ) = Z
2
The image of P
Z
4
forms a proper subset of the image of P
Z
2
, therefore there exist 2-qubit states with
Iso(ρ) = Z
2
, and these arise in several ways:
States of the form
ρ =
1
4
"
1 1 + a · σ 1 + 1 b · σ +
3
X
i=1
τ
i
c
i
· σ d
i
· σ
#
, (89)
where {c
i
} and {d
i
} are orthonormal bases of R
3
and c
i
6= d
i
for i = 1, 2, 3.
Accepted in Quantum 2019-01-23, click title to verify 25
States of the form
ρ =
1
4
"
1 1 + a · σ 1 + 1 b · σ + τ
1
r · σ r · σ +
3
X
i=2
τ
i
c
i
· σ d
i
· σ
#
, (90)
where {r, c
i
} and {r, d
i
} are orthonormal bases of R
3
but a and/or b are not parallel or anti-
parallel to r, i.e. there are no γ
1
, γ
2
R such that both a = γ
1
r and b = γ
2
r.
States of the form
ρ =
1
4
1 1 + a · σ 1 + 1 b · σ +
3
X
i,j=1
T
ij
σ
i
σ
j
, (91)
where a 6= γb for any γ R, so the local Bloch vectors are neither parallel nor anti-parallel to
each other.
C.4 U(1) Iso(ρ)
The next subgroup to consider is U (1). We apply P
U(1)
to states P
Z
4
(ρ), because if Z
4
Iso(ρ) then
U(1) Iso(ρ). For illustrative purposes, consider the U(1) subgroup {e
iθZ
: 0 θ < 2π}, which
contains Z
4
= 1, ±iZ}. The states invariant under this U(1) subgroup are
P
U(1)
(ρ) = P
U(1)
P
Z
4
(ρ) (92)
=
1
4
1 1 + a
z
Z 1 + b
z
1 Z + T
33
Z Z +
2
X
i,j=1
T
ij
P
U(1)
(σ
i
σ
j
)
(93)
because the terms Z 1, 1 Z and Z Z are certainly invariant under the U(1) subgroup generated
by Z. For the remaining correlation terms,
P
U(1)
(X X) =
1
2π
Z
2π
0
e
iθZ
Xe
iθZ
e
iθZ
Xe
iθZ
(94)
=
1
2π
Z
2π
0
cos
2
2θ X X + sin
2
2θ Y Y cos 2θ sin 2θ (X Y + Y X) (95)
=
1
2
(X X + Y Y ) = P
U(1)
(Y Y ) (96)
and similarly P
U(1)
(X Y ) = −P
U(1)
(Y X) =
1
2
(X Y Y X). Together these give
P
U(1)
(ρ) =
1
4
1 1 + a
z
Z 1 + b
z
1 Z + T
33
Z Z +
1
2
(T
11
+ T
22
)(X X + Y Y )
+
1
2
(T
12
T
21
)(X Y Y X)
(97)
Now consider the more general U (1) = {e
ir·σ
: 0 θ < 2π}. This acts on a Bloch vector as
e
r·σ
(u · σ)e
r·σ
= cos 2θ u · σ + sin 2θ (u × r) · σ + 2 sin
2
θ (r · u) r · σ, (98)
therefore
P
U(1)
(a · σ 1) =
1
2π
Z
2π
0
e
r·σ
(a · σ)e
r·σ
1 (99)
=
1
2π
Z
2π
0
cos 2θ a · σ 1 + sin 2θ (a × r) · σ 1 + 2 sin
2
θ (r · a) r · σ 1
= (r · a) r · σ 1 (100)
Accepted in Quantum 2019-01-23, click title to verify 26
and similarly P
U(1)
(1 b · σ) = (r · b) 1 r · σ. The group action on the correlation terms is
e
r·σ
(u · σ)e
r·σ
e
r·σ
(v · σ)e
r·σ
= cos
2
2θ u · σ v · σ + 4 sin
4
θ(r · u)(r · v)r · σ r · σ + 2 cos 2θ sin 2θ [u · σ (v × r) · σ
+ (u × r) · σ v · σ] + 2 cos 2θ sin
2
θ[(r · v)u · σ r · σ + (r · u)r · σ v · σ]
+ 2 sin 2θ sin
2
θ [(r · v)(u × r) · σ r · σ + (r · u)r · σ (v × r) · σ]
+ sin
2
2θ(u × r) · σ (v × r) · σ. (101)
This gives
P
U(1)
(c
i
· σ c
j
· σ) =
1
2π
Z
2π
0
e
r·σ
(c
i
· σ)e
r·σ
e
r·σ
(c
j
· σ)e
r·σ
(102)
=
1
2
[ c
i
· σ c
j
· σ (r · c
j
)c
i
· σ r · σ (r · c
i
)r · σ c
j
· σ
+ (c
i
× r) · σ (c
j
× r) · σ + 3(r · c
i
)(r · c
j
)r · σ r · σ ] (103)
=
1
2
[ c
i
· σ c
j
· σ δ
1j
c
i
· σ c
1
· σ δ
1i
c
1
· σ c
j
· σ
+ (c
i
× c
1
) · σ (c
j
× c
1
) · σ + 3 δ
1i
δ
1j
c
1
· σ c
1
· σ], (104)
therefore
P
U(1)
(ρ) =
1
4
1 1 + (r · a) r · σ 1 + (r · b) 1 r · σ + T
0
11
r · σ r · σ
+
1
2
(T
0
22
+ T
0
33
) [c
2
· σ c
2
· σ + c
3
· σ c
3
· σ]
+
1
2
(T
0
23
T
0
32
) [c
2
· σ c
3
· σ c
3
· σ c
2
· σ]
(105)
=
1
4
"
1 1 + (r · a) r · σ 1 + (r · b) 1 r · σ + τ
1
r · σ r · σ + τ
3
X
i=2
c
i
· σ d
i
· σ
#
(106)
where {r, c
i
} and {r, d
i
} are orthonormal bases of R
3
.
C.5 K
2
Iso(ρ)
Consider one such subgroup, K
2
= 1, ±iX, ±iY, ±iZ}. For this subgroup,
P
K
2
(ρ) =
1
4
1 1 + P
K
2
(a · σ 1) + P
K
2
(1 b · σ) +
3
X
i,j=1
T
ij
P
K
2
(σ
i
σ
j
)
. (107)
Local Bloch vectors are eliminated, since P
K
2
(X 1) = P
K
2
(Y 1) = P
K
2
(Z 1) = 0, and likewise
for the other local Bloch vector. For the correlation terms,
P
K
2
(X X) =
1
4
[X X + XXX XXX + Y XY Y XY + ZXZ ZXZ] = X X, (108)
and similarly P
K
2
(Y Y ) = Y Y and P
K
2
(Z Z) = Z Z. The remaining correlation terms vanish
since
P
K
2
(X Y ) =
1
4
[X Y + XXX XY X + Y XY Y Y Y + ZXZ ZY Z] (109)
=
1
4
[X Y X Y X Y + X Y ] = 0 (110)
Accepted in Quantum 2019-01-23, click title to verify 27
and similarly for other σ
i
σ
j
terms with i 6= j. Therefore
P
K
2
(ρ) =
1
4
"
1 1 +
3
X
i=1
T
ii
σ
i
σ
i
#
, (111)
which are the T-states. This proves that all T-states ρ
T
have K
2
Iso(ρ
T
).
Consider the more general K
2
subgroup, 1, ±ir
1
· σ, ±ir
2
· σ, ±ir
3
· σ} where {r
1
, r
2
, r
3
} forms an
orthonormal basis of R
3
. We use that same basis for the correlation terms, such that 2-qubit states
can be expressed
ρ =
1
4
1 1 + a · σ 1 + 1 b · σ +
3
X
i,j=1
T
0
ij
r
i
· σ r
j
· σ
. (112)
We first consider the action of P
K
2
on the local Bloch vectors, a · σ =
P
3
i=1
(r
i
· a) r
i
· σ,
P
K
2
(a · σ 1) =
1
4
a · σ 1 +
3
X
i=1
(r
i
· σ)(a · σ)(r
i
· σ) 1
(113)
=
1
4
a · σ 1 +
3
X
i=1
(2(r
i
· a)r
i
· σ a · σ) 1
(114)
=
1
2
"
3
X
i=1
(r
i
· a) r
i
· σ 1 a · σ 1
#
= 0, (115)
and similarly P
K
2
(1 b · σ) = 0. The correlation terms give
P
K
2
(r
i
· σ r
j
· σ) =
1
4
"
r
i
· σ r
j
· σ +
3
X
k=1
(r
k
· σ)(r
i
· σ)(r
k
· σ) (r
k
· σ)(r
j
· σ)(r
k
· σ)
#
(116)
=
1
4
"
r
i
· σ r
j
· σ +
3
X
k=1
4(r
k
· r
i
)(r
k
· r
j
)r
k
· σ r
k
· σ + r
i
· σ r
j
· σ
2(r
k
· r
i
)r
k
· σ r
j
· σ 2(r
k
· r
j
)r
i
· σ r
k
· σ

=
1
4
"
3
X
k=1
(4δ
ik
δ
jk
r
k
· σ r
k
· σ 2δ
ik
r
k
· σ r
j
· σ 2δ
jk
r
i
· σ r
k
· σ) + 4r
i
· σ r
j
· σ
#
= δ
ij
r
j
· σ r
j
· σ (117)
Therefore
P
K
2
(ρ) =
1
4
"
1 1 +
3
X
i=1
τ
i
r
i
· σ r
i
· σ
#
(118)
where {r
i
} is an orthonormal basis for R
3
.
C.6 Iso(ρ) = Z
4
We have now calculated P
H
(ρ) for the two subgroups directly above Z
4
on the subgroup lattice that
are possible isotropy subgroups. This identifies the 2-qubit states in C(Z
4
). They are:
States of the form
ρ =
1
4
1 1 + τ
1
r · σ r · σ +
3
X
i=2
τ
i
c
i
· σ d
i
· σ
(119)
where τ
2
6= τ
3
and c
i
6= d
i
for i = 2, 3.
Accepted in Quantum 2019-01-23, click title to verify 28
States of the form
ρ =
1
4
"
1 1 + ar · σ 1 + b1 r · σ + τ
1
r · σ r · σ +
3
X
i=2
τ
i
c
i
· σ d
i
· σ
#
(120)
where τ
2
6= τ
3
and c
i
6= d
i
for i = 2, 3, and a, b R.
The states
ρ =
1
4
"
1 1 + ac
1
· σ 1 + b1 c
1
· σ +
3
X
i=1
τ
i
c
i
· σ c
i
· σ
#
(121)
where a, b R. This gives a Z
4
isotropy subgroup when τ
1
= τ
2
6= τ
3
.
C.7 K
Iso(ρ)
The next subgroup to consider is K
, which contains the K
2
subgroup. Therefore we need only apply
P
K
to states P
K
2
(ρ). We can also say the same about states P
U(1)
(ρ), however the K
2
subgroup
projects onto a simpler subset of states.
As an illustrative example, consider the particular subgroup {(iX)
α
e
iθZ
: 0 θ < 2π, α = 0, 1, 2, 3}.
This contains the K
2
subgroup 1, ±iX, ±iY, ±iZ}, so we need only apply P
K
to the T-states. The
CPTP map P
K
can be split up as P
K
= P
U(1)
P
Z
4
, therefore
P
K
(ρ
T
) = P
U(1)
(ρ
T
) =
1
4
1 1 +
1
2
(τ
1
+ τ
2
) (X X + Y Y ) + τ
3
Z Z
(122)
More generally, we consider
K
= {(i r
j
· σ)
α
e
r
k
·σ
: 0 θ < 2π, α = 0, 1, 2, 3} (123)
where {r
i
} is an orthonormal basis of R
3
and j 6= k. The projector P
K
on a general 2-qubit state
gives
P
K
(ρ) = P
U(1)
P
K
2
(ρ) = P
U(1)
"
1
4
1 1 +
3
X
i=1
τ
i
r
i
· σ r
i
· σ
!#
(124)
=
1
4
1 1 +
1
2
X
i6=k
τ
i
X
i6=k
r
i
· σ r
i
· σ
+ τ
k
r
k
· σ r
k
· σ
(125)
C.8 Iso(ρ) = K
2
These are the T-states and the states with maximally mixed marginals that can be reached from the
T-states through rigid SU(2) rotations:
ρ =
1
4
1 1 +
X
i
τ
i
c
i
· σ c
i
· σ
, (126)
with τ
1
6= τ
2
6= τ
3
. There cannot be any local Bloch vectors.
C.9 Iso(ρ) = U(1)
These are states of the form
ρ =
1
4
"
1 1 + τ
1
r · σ r · σ + τ
3
X
i=2
c
i
· σ d
i
· σ
#
(127)
Accepted in Quantum 2019-01-23, click title to verify 29
where c
i
6= d
i
for i = 2, 3. Local Bloch vectors are permitted so long as they are either aligned or
anti-aligned with the shared Bloch vector in the correlation terms:
ρ =
1
4
"
1 1 + ar · σ 1 + b1 r · σ + τ
1
r · σ r · σ + τ
3
X
i=2
c
i
· σ d
i
· σ
#
(128)
where c
i
6= d
i
for i = 2, 3 and a, b R.
C.10 Iso(ρ) = SU(2)
Since G = SU(2), when we apply P
SU (2)
to a general 2-qubit state we get the form of a general 2-qubit
symmetric state. Since K
SU(2),
P
SU (2)
(ρ) = P
SU (2)
P
K
(ρ) =
1
4
1 1 +
3
X
i=1
τ
i
P
SU (2)
[c
i
· σ c
i
· σ]
!
. (129)
A general SU (2) group transformation is U (g) = e
r·σ
. We parameterise r =
(sin φ cos θ, sin φ sin θ, cos φ). Considering each of the terms individually,
P
SU (2)
(c
i
· σ c
i
· σ)
=
1
2π
2
Z
2π
0
Z
π
0
sin φ
Z
π
0
sin
2
ϕ
e
r·σ
c
i
· σe
r·σ
e
r·σ
c
i
· σe
r·σ
(130)
=
1
2π
2
Z
2π
0
Z
π
0
sin φ
Z
π
0
sin
2
ϕ cos
2
2ϕ c
i
· σ c
i
· σ
+ 2 sin
2
ϕ cos 2ϕ sin 2ϕ [c
i
· σ (c
i
× r) · σ + (c
i
× r) · σ c
i
· σ]
+ 2 cos 2ϕ sin
4
ϕ (r · c
i
) [c
i
· σ r · σ + r · σ c
i
· σ]
+ 2 sin 2ϕ sin
4
ϕ (r · c
i
) [(c
i
× r) · σ r · σ + r · σ (c
i
× r) · σ]
+ 4 sin
6
ϕ (r · c
i
)
2
r · σ r · σ + sin
2
ϕ sin
2
2ϕ (c
i
× r) · σ (c
i
× r) · σ (131)
=
1
8π
Z
2π
0
Z
π
0
sin φ
c
i
· σ c
i
· σ 2(r · c
i
) [c
i
· σ r · σ + r · σ c
i
· σ]
+ (c
i
× r) · σ (c
i
× r) · σ + 5(r · c
i
)
2
r · σ r · σ
(132)
Performing the integrals over θ and φ gives
P
SU (2)
(ρ) =
1
4
1 1 + τ
3
X
i=1
σ
i
σ
i
. (133)
C.11 Iso(ρ) = K
These are states in the image of P
K
but not in that of P
SU (2)
. They have the form
ρ =
1
4
"
1 1 +
X
i
τ
i
c
i
· σ c
i
· σ
#
(134)
where τ
1
= τ
2
6= τ
3
i.e. two of the τ
i
are the same, but not equal to the third.
Accepted in Quantum 2019-01-23, click title to verify 30