This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Kraus-Like Decompositions

Jonathan Boretsky J. Boretsky: Department of Mathematics, Harvard University, Cambridge, MA 02138  and  Robert Lin R. Lin: Department of Mathematics, Harvard University, Cambridge, MA 02138; Department of Chemistry and Chemical Biology, Harvard University, Cambridge, MA 02138
Abstract.

We introduce a new decomposition of quantum channels acting on group algebras, which we term Kraus-like (operator) decompositions. We motivate this decomposition with a general nonexistence result for Kraus operator decompositions in this setting. Given a length function which is a class function on a finite group, we construct a corresponding Kraus-like decomposition. We prove that this Kraus-like decomposition is convex (meaning its coefficients are nonnegative and satisfy a sum rule) if and only if the length is conditionally negative definite. For a general finite group, we prove a stability condition which shows that the existence of a convex Kraus-like decomposition for all t>0t>0 small enough necessarily implies existence for all time t>0t>0. Using the stability condition, we show that for a general finite group, conditional negativity of the length function is equivalent to a set of semidefinite linear constraints on the length function. Our result implies that in the group algebra setting, a semigroup PtP_{t} induced by a length function which is a class function is a quantum channel for all t0t\geq 0 if and only if it possesses a convex Kraus-like decomposition for all t>0t>0.

1. Introduction

The notion of a completely positive, trace-preserving map (CPTP map), also called a quantum channel, is important in a variety of areas, including quantum information theory and the study of various inequalities for operator algebras. Quantum channels are often studied in particular mathematical settings, such as full matrix algebras or other particular kinds of CC^{*} algebras. The setting of full matrix algebras is of particular importance in quantum information theory via the study of finite-dimensional density matrices and their properties (with respect to norms, entropies, etc.). In such a situation, quantum channels are characterized by the well-known Kraus operator decomposition theorem [Sti55][Cho75][NC10].

In this article, we specialize to the particular case of group algebras where the underlying group is finite111The group algebra setting has previously been studied in quantum computation in Kitaev’s quantum double model for finite groups [Kit03] (see [CDH+20] for recent results in this direction).. Quantum channels in the group algebra setting possess similar desirable properties as those in the full matrix algebra setting, and thus are of independent interest. For example, the interpolation theory of Uhlmann [Uhl77] extends the monotonicity of the relative entropy under identity-preserving completely positive maps from the usual matrix algebra setting to arbitrary *-algebras, which contains group algebras as a special case. We prove a nonexistence result for a Kraus operator decomposition of the quantum channel in terms of Kraus operators lying within the group algebra. This motivates us to introduce operators which allow one to decompose certain quantum channels acting on the group algebra. The decomposition we introduce involves a sum over operators which come with real-valued coefficients. Much as the work of Choi [Cho75] shows that the existence of a Kraus operator decomposition in the matrix algebra case implies that the corresponding linear mapping is a quantum channel, we will obtain a result in the group algebra case showing that, for particular naturally arising operator semigroups PtP_{t}, if a decomposition of PtP_{t} in terms of the operators we introduced satisfies a positivity condition on the coefficients, then the elements of the semigroup are quantum channels for all t0t\geq 0. Hence, we term these operators Kraus-like operators and the corresponding decompositions Kraus-like operator decompositions. We further define convex Kraus-like operator decompositions to be those in which the coefficients in the sum decomposition are nonnegative and satisfy a particular sum rule, which we will discuss explicitly later.

By the work of [Sch38], it is known that the imposition of a conditionally negative-definite length on the finite group underlying the group algebra results in the induced semigroup PtP_{t} (specifically, take the semigroup of operators PtP_{t} defined by Ptλ(g)=etl(g)λ(g)P_{t}\lambda(g)=e^{-tl(g)}\lambda(g) where λ(g)\lambda(g) is a multiplier and ll is a scalar-valued function on the group, and extend by linearity to all of G\mathcal{L}G) being a quantum channel for all t0t\geq 0. By restricting to length functions which are class functions, and passing to the Kraus-like operator decomposition, we show that the condition of conditional negative-definiteness can be efficiently verified by checking a number of semidefinite linear constraints on the length function. We show that the latter is a necessary and sufficient condition for conditional negative-definiteness, and thus for the semigroup to be a quantum channel for all t0t\geq 0.

The choice of multipliers in our framework is canonical, as we induce the multipliers by the characters of the finite group upon which the group algebra is built. This enables us to use representation theory of finite groups to achieve our results.

In terms of our proof method, for a general finite group GG, in order to obtain our semi-definite linear constraints, we prove a stability condition which shows that if the semigroup associated with a length has a convex Kraus-like operator decomposition for all t>0t>0 small enough, then it has a convex Kraus-like operator decomposition for all time t>0t>0. Thus, to obtain global positivity of the coefficients, it suffices to check positivity near t=0+t=0^{+}, which reduces to a bound on the derivatives. This derivative bound is what yields the semi-definite linear constraints.

2. Definitions

The main objects we are working with in this paper are the left regular representation of a finite group GG, and a semigroup acting on the elements in the left regular representation. Later in this section, we will find it useful to restrict to semigroups which are induced by conditionally negative-definite length functions on GG, in a way to be precisely defined.

Definition 2.1 (Left Regular Representation).

Given a group GG, let G\mathcal{F}G be the vector space of complex-valued functions on GG. We denote by λ\lambda the left regular representation of GG, which acts on G\mathcal{F}G by: (λ(g)f)(h)=f(hg1)(\lambda(g)f)(h)=f(hg^{-1}) for gGg\in G and fGf\in\mathcal{F}G. Denote the \mathbb{C}-linear span of {λ(g)}gG\{\lambda(g)\}_{g\in G} by G\mathcal{L}G.

As it is a property we will use repeatedly, we emphasize that by definition of a representation, for each g,hg,h in GG, we have the equality λ(g)λ(h)=λ(gh)\lambda(g)\lambda(h)=\lambda(gh) of operators on G\mathcal{F}G.

Recall the standard inner product on G\mathcal{F}G, given by f,hG=gGf(g)¯h(g)\left<f,h\right>_{\mathcal{F}G}=\sum_{g\in G}\overline{f(g)}h(g) for f,hGf,h\in\mathcal{F}G. The space G\mathcal{L}G also comes with a natural inner product. Let δe\delta_{e} be the function on GG defined by δe(g)=δg=e\delta_{e}(g)=\delta_{g=e}. Then, we define x,yG=δe,xyδeG\langle x,y\rangle_{\mathcal{L}G}=\langle\delta_{e},x^{*}y\,\delta_{e}\rangle_{\mathcal{F}G}, where the * operation is defined on G\mathcal{L}G antilinearly with iαiλ(gi)iαi¯λ(gi1)\sum_{i}\alpha_{i}\lambda(g_{i})\mapsto\sum_{i}\bar{\alpha_{i}}\lambda(g_{i}^{-1}). It is straightforward to verify that this is in fact an inner product and thus, (G,,G)(\mathcal{L}G,\braket{\cdot,\cdot}_{\mathcal{L}G}) is a Hilbert space.

To define a semigroup on G\mathcal{L}G, one needs the notion of a length function on GG. Following [JPPP17], we restrict ourselves to conditionally negative-definite lengths.

Definition 2.2.

A length function l:Gl:G\rightarrow\mathbb{R} on a group GG with identity ee is a function which satisfies l(e)=0l(e)=0, l(g)=l(g1)l(g)=l(g^{-1}) and l(g)0l(g)\geq 0 for all gg in GG.

If l(g)>0l(g)>0 for all geg\neq e, then we will call ll a strict length function.

Definition 2.3 ([vdBF75]).

A length function l:Gl:G\rightarrow\mathbb{R} is said to be conditionally negative-definite if for any α1,αn\alpha_{1},\cdots\alpha_{n}\in\mathbb{C} satisfying i=1nαi=0\sum_{i=1}^{n}\alpha_{i}=0 and any g1,,gnGg_{1},\cdots,g_{n}\in G, one has

i,j=1nαiαj¯l(gi1gj)0.\sum_{i,j=1}^{n}\alpha_{i}\overline{\alpha_{j}}l(g_{i}^{-1}g_{j})\leq 0.

An important additional assumption we adopt in this work is the assumption that all length functions are class functions, meaning they are constant on conjugacy classes. Symbolically, this means l(g)=l(h1gh)l(g)=l(h^{-1}gh) for any g,hg,h in GG. This assumption will be greatly exploited in our results and we will derive new characterizations of these conditionally negative-definite class function lengths.

We consider the semigroup PtP_{t} of operators on G\mathcal{L}G induced by a length function l()l(\cdot) which is given on generators of G\mathcal{L}G by the action

Ptλ(g)=etl(g)λ(g),P_{t}\lambda(g)=e^{-t\,l(g)}\lambda(g), (1)

and extended linearly. Observe that indeed, P0P_{0} acts by the identity and Pt1Pt2=Pt1+t2P_{t_{1}}P_{t_{2}}=P_{t_{1}+t_{2}}.

Our goal is to characterize PtP_{t} from various perspectives. Our new perspectives on PtP_{t} will shed light on a known characterization of PtP_{t} [JPPP17] presented in the continuation of this section.

Definition 2.4.

A Hermitian function K:G×GK:G\times G\rightarrow\mathbb{C} is said to be a positive definite kernel if for any α1,,αn\alpha_{1},\cdots,\alpha_{n}\in\mathbb{C} and any g1,,gnGg_{1},\cdots,g_{n}\in G, we have

i,j=1nαiαj¯K(gi,gj)0.\sum_{i,j=1}^{n}\alpha_{i}\overline{\alpha_{j}}K(g_{i},g_{j})\geq 0.

Schoenberg’s theorem provides a characterization of positive definite kernels in terms of conditionally negative-definite lengths:

Theorem 2.5 (Schoenberg’s Theorem [vdBF75]).

Let GG be a group. A function l:Gl:G\rightarrow\mathbb{R} satisfying l(g)=l(g1)l(g)=l(g^{-1}) for all gGg\in G is conditionally negative definite if and only if the following conditions hold:

  1. (1)

    l(e)=0l(e)=0, for ee the identity of GG, and

  2. (2)

    The function G×GG\times G\rightarrow\mathbb{C} defined by (g,h)etl(gh1)(g,h)\mapsto e^{-t\,l(gh^{-1})} is positive definite.

In Proposition 2.7, we recall an equivalent characterization of conditionally negative-definite lengths in terms of the notion of complete positivity, which is important in quantum physics and quantum information theory. Recall that a linear map is positive if it maps positive elements to positive elements. Following [NS06],

Definition 2.6.

Let 𝒜\mathcal{A} and \mathcal{B} be CC^{*} algebras, and θ:𝒜\theta:\mathcal{A}\rightarrow\mathcal{B} be a linear map. The map θ\theta is called completely positive if

θid:𝒜Matn()Matn()\theta\otimes\text{id}:\mathcal{A}\otimes\text{Mat}_{n}(\mathbb{C})\rightarrow\mathcal{B}\otimes\text{Mat}_{n}(\mathbb{C}) (2)

is positive for any nn\in\mathbb{N}.

A consequence of Schoenberg’s theorem is that the semigroup PtP_{t} is completely positive if and only if l()l(\cdot) is conditionally negative-definite (stated, but not proved, in [JPPP17]). To be complete, we provide a proof:

Proposition 2.7.

PtP_{t} is completely positive if and only if ll is conditionally negative-definite.

Proof.

(\Leftarrow) By Schoenberg’s theorem, if ll is conditionally negative-definite, then (etl(g1h))g,hG\left(e^{-t\,l(g^{-1}h)}\right)_{g,h\in G} is a positive semi-definite matrix. From appendix A of [NS06], Pt:GGP_{t}:\mathcal{L}G\rightarrow\mathcal{L}G is completely positive if and only if

i,j=1nbiPt(aiaj)bj0\sum_{i,j=1}^{n}b_{i}^{*}P_{t}(a_{i}^{*}a_{j})b_{j}\geq 0 (3)

for any nn\in\mathbb{N}, a1,,an,b1,,bnGa_{1},\ldots,a_{n},b_{1},\ldots,b_{n}\in\mathcal{L}G. We show that the latter holds.

Fix nn. Take ai=gGai(g)λ(g)a_{i}=\sum_{g\in G}a_{i}(g)\lambda(g). Then

Pt(aiaj)=g,hGai(g)aj(h)etl(g1h)λ(g1h).P_{t}(a_{i}^{*}a_{j})=\sum_{g,h\in G}a_{i}(g)^{*}a_{j}(h)e^{-t\,l(g^{-1}h)}\lambda(g^{-1}h). (4)

So we want to show that

S:=i,j=1ng,hGbiai(g)aj(h)etl(g1h)λ(g1h)bjS:=\sum_{i,j=1}^{n}\sum_{g,h\in G}b_{i}^{*}a_{i}(g)^{*}a_{j}(h)e^{-t\,l(g^{-1}h)}\lambda(g^{-1}h)b_{j} (5)

is non-negative. We note that this can be rearranged by setting vg=i=1nai(g)λ(g)biv_{g}=\sum_{i=1}^{n}a_{i}(g)\lambda(g)b_{i}. So the sum becomes

g,hGvgetl(g1h)vh.\sum_{g,h\in G}v_{g}^{*}e^{-t\,l(g^{-1}h)}v_{h}. (6)

Note that since (etl(g1h))g,hG\left(e^{-t\,l(g^{-1}h)}\right)_{g,h\in G} is positive semi-definite, we have that

etl(g1h)=xGrx(t)(wx(t)wx(t))g,he^{-tl(g^{-1}h)}=\sum_{x\in G}r_{x}(t)\left(w_{x}(t)^{*}w_{x}(t)\right)_{g,h} (7)

for some matrices {wx(t)|xG}\{w_{x}(t)|x\in G\}, and nonnegative numbers {rx(t)|xG}\{r_{x}(t)|x\in G\}.

Thus,

S\displaystyle S =g,hGvgxGrx(t)(wx(t)wx(t))g,hvh\displaystyle=\sum_{g,h\in G}v_{g}^{*}\sum_{x\in G}r_{x}(t)\left(w_{x}(t)^{*}w_{x}(t)\right)_{g,h}v_{h} (8)
=xGrx(t)g,hGvg(wx(t)wx(t))g,hvh\displaystyle=\sum_{x\in G}r_{x}(t)\sum_{g,h\in G}v_{g}^{*}\left(w_{x}(t)^{*}w_{x}(t)\right)_{g,h}v_{h} (9)
=xGrx(t)g,hGvgmG(wx(t))g,m(wx(t))m,hvh\displaystyle=\sum_{x\in G}r_{x}(t)\sum_{g,h\in G}v_{g}^{*}\sum_{m\in G}\left(w_{x}(t)^{*}\right)_{g,m}\left(w_{x}(t)\right)_{m,h}v_{h} (10)

Set

qx,m(t)=gG(wx(t))m,gvgq_{x,m}(t)=\sum_{g\in G}\left(w_{x}(t)\right)_{m,g}v_{g} (11)

then

S=xrx(t)mGqx,m(t)qx,m(t).S=\sum_{x}r_{x}(t)\sum_{m\in G}q_{x,m}^{*}(t)q_{x,m}(t). (12)

Thus, SS is positive semi-definite, as desired.

(\Rightarrow) Conversely, suppose PtP_{t} is completely positive. Then, with notation as above, for all nn\in\mathbb{N} and a1,,an,b1,,bnGa_{1},\cdots,a_{n},b_{1},\cdots,b_{n}\in\mathcal{L}G,

Si,j=1nbiPt(aiaj)bj=g,hGvget(g1h)vh0,S\coloneqq\sum_{i,j=1}^{n}b_{i}^{*}P_{t}(a_{i}^{*}a_{j})b_{j}=\sum_{g,h\in G}v_{g}^{*}e^{-t\ell(g^{-1}h)}v_{h}\geq 0, (13)

where vgi=1nai(g)λ(g)biv_{g}\coloneqq\sum_{i=1}^{n}a_{i}(g)\lambda(g)b_{i}. Since this holds for any choice of nn, aia_{i} and bib_{i}, we can fix n=|G|n=|G|. Order the elements of GG so that G={g1,,gn}G=\{g_{1},\cdots,g_{n}\}. Choose bi=λ(gi1)b_{i}=\lambda(g_{i}^{-1}) and choose aia_{i} such that ai(gj)=cjδija_{i}(g_{j})=c_{j}\delta_{ij} for some cjc_{j}\in\mathbb{C}. Then, for each j=1,nj=1,\cdots n, vgj=cjλ(e)v_{g_{j}}=c_{j}\lambda(e). This reduces eq. 13 to

(i,j=1nci¯et(gi1gj)cj)λ(e)0,\left(\sum_{i,j=1}^{n}\overline{c_{i}}e^{-t\ell(g_{i}^{-1}g_{j})}c_{j}\right)\lambda(e)\geq 0,

for any choice of c1,cnc_{1},\cdots c_{n}\in\mathbb{C}. Now we claim that if Aλ(e)0A\lambda(e)\geq 0 with AA\in\mathbb{C}, then A0A\geq 0. This follows since λ(e)\lambda(e) acts as the identity, and so its spectrum is just 11, and so the spectrum of Aλ(e)A\lambda(e) is just AA. Thus, for any choice of cnc\in\mathbb{C}^{n}, we have

i,j=1nci¯et(gi1gj)cj0.\sum_{i,j=1}^{n}\overline{c_{i}}e^{-t\ell(g_{i}^{-1}g_{j})}c_{j}\geq 0.

By definition, this means that the matrix {et(gi1gj)}i,j=1n\{e^{-t\ell(g_{i}^{-1}g_{j})}\}_{i,j=1}^{n} is positive semi-definite and so, by Schoenberg’s theorem, \ell is a conditionally negative-definite length. ∎

Corollary 2.8.

PtP_{t} is a quantum channel.

Proof.

Since PtP_{t} is also trace-preserving, it follows by definition that PtP_{t} is a quantum channel [NC10]. ∎

3. Kraus-Like Operator Decompositions

In quantum information theory, Kraus operators are used in sum representations of quantum channels which describe the dynamics of the density matrix of a system [NC10]. In fact, in the matrix case, quantum channels can be characterized by the existence of a Kraus operator decomposition. Equivalently, one may describe a quantum channel as the result of tracing out a subsystem from a unitary operator acting on a composite system. Conversely, one may always “lift” a quantum channel on a density matrix to a corresponding unitary operator on a larger system, a process known as Stinespring dilation [Sti55].

One of the main questions which drives this work is whether the usual intuition for quantum channels on density matrices extends to those on group algebras. Namely, does one get Kraus operators? And what do they look like?

What we find is that the Kraus operators, if they exist, are certainly not generally elements of the group algebra. This is in direct contrast to the case of quantum channels acting on density matrices, where the Kraus operators are themselves matrices.

This nonexistence result motivates us to look for alternate decompositions of the quantum channel, in the spirit of the Kraus operator decomposition. The main idea is to relax the action of Kraus operators as EEE\cdot E^{\dagger} to some linear operator σ\sigma\cdot, where σ\sigma acts on G\mathcal{L}G. We call our new σ\sigma’s Kraus-like operators.

For a suitable choice of σ\sigma’s, we can obtain an explicit condition on the decomposition of a semigroup PtP_{t} induced by a length function ll which is also a class function, which determines whether or not PtP_{t} is a quantum channel.

3.1. An algebraic obstruction result on Kraus operator decompositions

Consider the semigroup PtP_{t}, induced by a length ll, which acts on the Hilbert space =G\mathcal{H}=\mathcal{L}G. We recall that this is given as the linear extension of:

Ptλ(g)=etl(g)λ(g).P_{t}\lambda(g)=e^{-tl(g)}\lambda(g). (14)

A Kraus decomposition of PtP_{t} is a decomposition of PtP_{t} given by

Pt(x)=iEixEi,P_{t}(x)=\sum_{i}E_{i}xE_{i}^{\dagger}, (15)

where the elements EiE_{i} satisfy iEiEiI\sum_{i}E_{i}^{\dagger}E_{i}\leq I [NC10]. The EiE_{i} are called Kraus operators. For simplicity, we focus only on the case where iEiEi=I\sum_{i}E_{i}^{\dagger}E_{i}=I, corresponding to the case where one can dilate PtP_{t} to a unitary in the matrix case [NC10]. In the usual matrix algebra setting for Kraus operator decompositions, quantum information theory, xx would be a finite-dimensional density matrix and PtP_{t} would be a completely positive map from density matrices to density matrices. The elements EiE_{i} would be matrices. For the group algebra setting, it is not a priori obvious where the EiE_{i}’s would live, so we will make some natural assumptions.

Since we are working with group algebras, to employ Kraus operator decompositions, some choices must be made as to the proper identification of terms. The natural mapping, extending the setting of matrix algebras to direct sum of matrix algebras, is to take PtP_{t} to map G\mathcal{L}G into G\mathcal{L}G. While one could also embed G\mathcal{L}G into a matrix algebra by the regular representation, this is fairly unnatural from the point of respecting the symmetry of the group, as different irreducible representations ought not to interact from a physical perspective.

For a Kraus operator decomposition, since we work in G\mathcal{}G, it is further natural, or at least convenient, to assume that the EiE_{i}’s all lie in G\mathcal{L}G. Note that this is not the most general setting, since if we interpret G\mathcal{L}G as a vector space of dimension |G||G|, the dimension of (G)\mathcal{B}(\mathcal{L}G) is |G|2|G|^{2}, whereas the embedding of G\mathcal{L}G inside (G)\mathcal{B}(\mathcal{L}G) only has dimension |G||G|. So we are deliberately choosing to focus on a smaller space of possible Kraus operators. However, we show that with this simple, and perhaps most natural, choice, we run into issues.

The following proposition shows that a Kraus operator decomposition as described in the previous paragraph is not the right tool for the job at the hand, in the sense that Kraus operator decompositions will not be readily available in many semigroups of interest.

Proposition 3.1.

Any finite group GG whose group algebra G\mathcal{L}G has a non-zero element of the form egGagλ(g)\sum_{e\neq g\in G}a_{g}\lambda(g) in its center will not admit a Kraus operator decomposition for the operator PtP_{t} induced by a strict length function ll via equation 1, in terms of Kraus operators lying in G\mathcal{L}G.

In particular, the hypothesis of this proposition holds for the expression egGλ(g)G\sum_{e\neq g\in G}\lambda(g)\in\mathcal{L}G when GG is any nontrivial finite group.

Proof.

Let hh be an element satisfying the conditions of the proposition. Suppose PtP_{t} admits a Kraus decomposition of the form Pt(x)=iEixEiP_{t}(x)=\sum_{i}E_{i}xE_{i}^{\dagger}, where EiGE_{i}\in\mathcal{L}G. Since λ(e)\lambda(e) is the identity, it follows that

λ(e)=exp(l(e)t)λ(e)=Ptλ(e)=iEiEi.\lambda(e)=\exp(-l(e)t)\lambda(e)=P_{t}\lambda(e)=\sum_{i}E_{i}E_{i}^{\dagger}.

Now, consider Pt(h)P_{t}(h). Since hh is in the center, we have

Pt(h)=iEihEi=iEiEih=λ(e)h=h.P_{t}(h)=\sum_{i}E_{i}hE_{i}^{\dagger}=\sum_{i}E_{i}E_{i}^{\dagger}h=\lambda(e)h=h.

However, by assumption, h=geagλ(g)h=\sum_{g\neq e}a_{g}\lambda(g), so

Pt(h)=geel(g)tagλ(g)=h=geagλ(g).P_{t}(h)=\sum_{g\neq e}e^{-l(g)t}a_{g}\lambda(g)=h=\sum_{g\neq e}a_{g}\lambda(g).

By the basis property, this implies that (el(g)t1)ag=0(e^{-l(g)t}-1)a_{g}=0 for all geg\neq e and for all t0t\geq 0. Since for a length function ll, l(g)>0l(g)>0 for all geg\neq e, this implies that ag=0a_{g}=0 for all geg\neq e. Thus, h=0h=0.

3.1.1. Discussion of dilation theorems

Some remarks must be made with respect to the Stinespring dilation theorem [Sti55], and the associated Choi isomorphism theorem [Cho75]. Firstly, the Stinespring dilation theorem still applies in this context of group algebras, but the construction of a dilation is so general as not to yield anything resembling a Kraus operator decomposition. The much sharper construction of Choi applies in the case of a finite-dimensional Hilbert space. When we look at the group algebras, the theorem of Choi is hard to apply directly, for the following reason: Embedding the group algebra G\mathcal{L}G into a matrix algebra via the left regular representation is a sparse isometric embedding222To see that the embedding is isometric, we can first note that in the case of N\mathbb{Z}_{N}, one easily sees that all the elements in the left-regular representation (except the identity) have no fixed points, so ρ(g),ρ(h):=tr(ρ(g)ρ(h)=tr(ρ(gh1)=0\langle\rho(g),\rho(h)\rangle:=\text{tr}(\rho(g)\rho(h)^{\dagger}=\text{tr}(\rho(gh^{-1})=0 unless g=hg=h. For the general finite group case, an element ρ(g)\rho(g) has an element on the diagonal only if gh=hgh=h for some hh. But this implies that g=eg=e since all elements are invertible in a group. So no non-identity elements of the left-regular representation have elements on the diagonal. Thus, the embedding is isometric in general., since the former has a basis set of size |G||G| whereas the latter has a basis set of size |G|2|G|^{2}. Thus, our definition of PtP_{t} as a completely positive, trace-preserving map on the group algebra G\mathcal{L}G does not mean that PtP_{t} is a completely positive, trace-preserving map on the corresponding matrix algebra, because the action of PtP_{t} is not even specified for matrices outside of the left regular representation. What one would be looking for instead is some analogue of Choi’s isomorphism theorem, which applies to a map which is completely-positive and trace-preserving on a subalgebra of a matrix algebra. Such a kind of restriction theorem333We are inspired by Stein’s restriction conjecture in analysis to use this suggestive vocabulary. would be interesting in its own right. Of course, one would be working with objects which are not really quantum channels, but only behave like quantum channels when applied to a subalgebra of the matrix algebra (as justified by Corollary 2.8). The corresponding extension problem has been considered recently by [Wir22] on an extension result for quantum Markov semigroups (completely positive maps which form a semigroup) defined on a subalgebra to a quantum Markov semigroup over the full matrix algebra.

3.2. Kraus-like Operator Decompositions

3.2.1. What is a Kraus-like Operator Decomposition?

The above nonexistence result motivates us to look for a more general decomposition of a semigroup, which will exist even when a Kraus operator decomposition in terms of group algebra elements is not available. We introduce the notion of a Kraus-like operator decomposition, which replaces the Kraus form by a multiplier on the group algebra. Under several equivalent hypotheses on the length function, we will be able to show that the coefficients of the decomposition satisfy a sum rule and are positive, hence admitting a possible probabilistic interpretation. In this respect, our motivation is to establish something analogous to the mixed-unitary quantum channel for our Kraus-like operator decomposition.

We now present a prototype for what we consider to be a Kraus-like operator decomposition. Consider a decomposition of the operator semigroup PtP_{t} into a sum of isometries, rather than taking a sum of EixEiE_{i}xE_{i}^{\dagger} operators. Let us show that, at least in a particularly simple example, such a decomposition does in fact exist.

Example 3.2.

Let GG be an arbitrary group with the length l(g)=1δg=el(g)=1-\delta_{g=e} for gGg\in G. Let p=1et2p=\frac{1-e^{-t}}{2}. Define the isometry

σ(λ(g))={λ(g),geλ(e),g=e\sigma(\lambda(g))=\begin{cases}-\lambda(g),&g\neq e\\ \lambda(e),&g=e\end{cases} (16)

Then for any gGg\in G, one can easily verify that Pt(λ(g))=(1p)λ(g)+pσ(λ(g))P_{t}(\lambda(g))=(1-p)\lambda(g)+p\sigma(\lambda(g)). The λ(g)\lambda(g) span G\mathcal{L}G so this proves that Pt=(1p)𝕀+pσP_{t}=(1-p)\mathbb{I}+p\sigma as operators on G\mathcal{L}G.

Let us make a few observations about this decomposition. The coefficients of the two isometries in the decomposition are (1p)(1-p) and pp. We note that by the definition of pp, these are both non-negative numbers for t0t\geq 0. Moreover, they evidently sum to 11.

This is a basic example, but we already see that there is hope that our decomposition might have a probabilistic interpretation. Motivated by this example, we introduce the following notion as an analog to Kraus operator decompositions:

Definition 3.3.

For a group GG and operator semigroup Pt:GGP_{t}:\mathcal{L}G\rightarrow\mathcal{L}G, a Kraus-like operator decomposition of PP is a decomposition

Pt=ipi(t)σiP_{t}=\sum_{i}p_{i}(t)\sigma_{i} (17)

where each operator σi:GG\sigma_{i}:\mathcal{L}G\rightarrow\mathcal{L}G is diagonal in the basis of left-multipliers λ(g)\lambda(g), and pi(t)p_{i}(t)’s are complex-valued functions.

If the pi(t)p_{i}(t)’s are all nonnegative, and satisfy a sum rule iαipi(t)=1\sum_{i}\alpha_{i}p_{i}(t)=1, for some positive αi\alpha_{i}’s independent of tt, then we say that we have a convex Kraus-like operator decomposition.

The natural next question is if this sort of decomposition can be generalized to the semigroups generated by other lengths. We will show that, indeed, it can. In fact, for a specific class of naturally arising σi\sigma_{i} operators, we will even be able to describe a condition which classifies precisely which lengths on a given group will yield semigroups with convex Kraus-like operator decompositions.

4. Character-Induced Kraus-Like Operator Decompositions

4.1. Kraus-like operator decompositions for finite abelian groups

Our next goal is to consider a general abelian group GG and think more broadly about when we have a convex Kraus-like operator decomposition.

Fix a finite abelian group GG of size nn. It is natural to consider maps which are induced by the characters of GG. Explicitly, if the characters of GG are denoted by {χi}i=1n\{\chi_{i}\}_{i=1}^{n}, then we consider the maps which act on generators by σi:λ(g)χi(g)λ(g)\sigma_{i}:\lambda(g)\mapsto\chi_{i}(g)\lambda(g) and extend by linearity. Note that since GG is abelian, all characters are simply one-dimensional representations. These are, in fact, isometries and the multiplicative structure they inherit from the fact that they are representations will prove useful later.

Let ll be any length on GG. Note that the characters of a group span the class functions on that group and, in the case of an abelian group where each element is its own conjugacy class, this means that the characters span the complex-valued functions on the group. Thus, we can write PtP_{t} as a sum Pt=k=1npkσkP_{t}=\sum_{k=1}^{n}p_{k}\sigma_{k} for appropriate pkp_{k}. By applying both sides of the previous equation to λ(g)\lambda(g) for each gGg\in G and comparing coefficients, one finds that the pkp_{k} must satisfy:

l=1nχl(g)pl=exp(tl(g)).\sum_{l=1}^{n}\chi_{l}(g)p_{l}=\exp(-tl(g)). (18)

Recall that part of our goal in all this is to understand if there is an interpretation of the pkp_{k} as probabilities. As we now show, this is equivalent to demanding that f(g)=exp(tl(g))f(g)=\exp(-tl(g)), as a function of gGg\in G, be of positive type. In other words, we must have that f(gh1)f(gh^{-1}) is a positive-definite kernel, when considered as a function of both gg and hh.

Proposition 4.1 (Bochner-like Theorem).

Let GG be a finite abelian group of size nn with characters {χl}l=1n\{\chi_{l}\}_{l=1}^{n}, and suppose f:Gf:G\rightarrow\mathbb{C}, and pnp\in\mathbb{C}^{n} are related by f(g)=l=1nplχl(g)f(g)=\sum_{l=1}^{n}p_{l}\chi_{l}(g) for all gGg\in G.

Then pp is a probability measure on GG, that is, pi0p_{i}\geq 0 for all ii and ipi=1\sum_{i}p_{i}=1, if and only if f(gh1)f(gh^{-1}), considered as a function of g,hGg,h\in G, is a positive definite kernel, and f(e)=1f(e)=1.

Proof.

Note that the positive definiteness of ff is equivalent to the non-negativity of the following expression, for any choice of φ:G\varphi:G\rightarrow\mathbb{C}:

gGhGf(gh1)φ(g)φ(h)=l=1nplgGhGχl(gh1)φ(g)φ(h)\displaystyle\sum_{g\in G}\sum_{h\in G}f(gh^{-1})\varphi(g)\varphi(h)^{*}=\sum_{l=1}^{n}p_{l}\sum_{g\in G}\sum_{h\in G}\chi_{l}(gh^{-1})\varphi(g)\varphi(h)^{*}
=\displaystyle= l=1npl(gGχl(g)φ(g))(hGχl(h)φ(h))=l=1npl|gGχl(g)φ(g)|2,\displaystyle\sum_{l=1}^{n}p_{l}\left(\sum_{g\in G}\chi_{l}(g)\varphi(g)\right)\left(\sum_{h\in G}\chi_{l}(h)\varphi(h)\right)^{*}=\sum_{l=1}^{n}p_{l}\left|\sum_{g\in G}\chi_{l}(g)\varphi(g)\right|^{2},

where the second equality follows from the general fact that χ(g1)=χ(g)\chi(g^{-1})=\chi(g)^{*} and also the fact that any representation of an abelian group is 11 dimensional, so the characters of such a group are multiplicative. Note that if each pk0p_{k}\geq 0, then this expression is surely non-negative for any choice of φ\varphi. On the other hand, if some pkp_{k} is not greater than or equal to 0, then set φ(g)=χk(g)\varphi(g)=\chi_{k}(g). By the orthogonality of characters, this will result in the entire expression failing to be greater than or equal to zero. Thus, we conclude that ff is positive definite if and only if each of the pl0p_{l}\geq 0 for all 1ln1\leq l\leq n. Additionally, the condition that l=0N1p(l)=1\sum_{l=0}^{N-1}p(l)=1 is equivalent to the condition that f(e)=1f(e)=1, completing our proof.

Our goal was to characterize the lengths ll on a finite abelian group GG for which we obtain a probability measure pp in the convex Kraus-like operator decomposition of the operator semigroup PtP_{t} induced by ll. Using Schoenberg’s theorem, we can easily obtain such a condition.

Proposition 4.2.

Let ll be any length on a finite abelian group GG and let {χl}l=1n\{\chi_{l}\}_{l=1}^{n} be the characters of GG. Suppose the semigroup PtP_{t} is defined by Pt:λ(g)et|g|λ(g)P_{t}:\lambda(g)\mapsto e^{-t|g|}\lambda(g), which extends linearly to all of G\mathcal{L}G. Let PtP_{t} satisfy Pt=k=1npkσkP_{t}=\sum_{k=1}^{n}p_{k}\sigma_{k}, where σk(λ(g))=χk(g)λ(g)\sigma_{k}(\lambda(g))=\chi_{k}(g)\lambda(g) for all gGg\in G. Then pp is a probability measure on GG if and only if ll is a conditionally negative-definite length.

Proof.

This follows by combining the previous proposition with Schoenberg’s theorem, which says that ll is conditionally negative-definite if and only if f(gh1)=exp(tl(gh1))f(gh^{-1})=\exp(-tl(gh^{-1})) is a positive-definite kernel in g,hGg,h\in G which satisfies f(e)=1f(e)=1. ∎

Thus, we have established a nice coherent story for finite abelian groups. We have characterized a large class of lengths on these groups which admit a convex Kraus-like operator decomposition. However, this is hardly satisfying. For one thing, it is not clear whether the relationship between a conditionally negative-definite length and a convex Kraus-like decomposition is an accident or has some more fundamental significance. Moreover, it would be valuable to move this analysis beyond abelian groups. Thus, we explore next the notion of Kraus-like operator decompositions for general finite groups.

4.2. Kraus-like decompositions for general finite groups

In the more general setting of finite groups, there are two basic questions we need to answer.

  1. (1)

    What is the appropriate generalization of the Kraus-like operator decomposition to a general finite group GG, based on the model we considered for N\mathbb{Z}_{N}?

  2. (2)

    What is the corresponding condition for the coefficients pip_{i} arising in the decomposition to be non-negative, or more specifically, for the pip_{i}’s to be a probability distribution (at least up to rescaling)?

The answer to the question (1) is that we can simply define multipliers induced by characters in the following way: Since we suppose that ll is a class function, PtP_{t} acts as a constant multiple of the identity on the left-multipliers λ(g)\lambda(g) for gCig\in C_{i}, for each conjugacy class CiC_{i}. Accordingly,

Pt=i(etl(Ci)1#Ci),P_{t}=\oplus_{i}\left(e^{-t\,l(C_{i})}\otimes 1_{\#C_{i}}\right), (19)

where l(Ci)l(C_{i}) is the unique value of the length function on the conjugacy class CiC_{i}.

We may also use the irreducible characters χ\chi of the group GG, which are themselves class functions, to induce maps with similar direct sum decompositions,

σχ:=i(χ(Ci)1#Ci).\sigma_{\chi}:=\oplus_{i}\left(\chi(C_{i})\otimes 1_{\#C_{i}}\right). (20)

From the similar forms of these operators, it seems reasonable to study relationships of the form

Pt=χpχσχ,P_{t}=\sum_{\chi}p_{\chi}\sigma_{\chi}, (21)

for complex numbers pχp_{\chi}. We call this expression a character-induced Kraus-like operator decomposition of the semigroup PtP_{t}.

Let {χ1,,χm}\{\chi_{1},\cdots,\chi_{m}\} be the irreducible characters of GG. We can quickly determine the coefficients pjp_{j}. By applying the map Pt=jpjσχjP_{t}=\sum_{j}p_{j}\sigma_{\chi_{j}} to λ(g)\lambda(g) for gCig\in C_{i} and setting the two sides equal to each other, one obtains, for each ii, the equation:

jpjχji=etl(Ci)\sum_{j}p_{j}\chi_{ji}=e^{-t\,l(C_{i})} (22)

where χji=χj(Ci)\chi_{ji}=\chi_{j}(C_{i}). We can consider these as entries of a matrix χ:=(χij)\chi:=(\chi_{ij}). Note that χ\chi is simply the character table of the group. With this, we can see eq. 22 as a matrix equation.

As a preliminary step, we obtain a sum rule for the pip_{i}’s:

Lemma 4.3 (Sum Rule for pip_{i}’s).
ipiχi(e)=1.\sum_{i}p_{i}\chi_{i}(e)=1. (23)
Proof.

This follows from equation 22 applied to the conjugacy class {e}\{e\}. ∎

We can solve for the pip_{i}’s by inverting the matrix equation 22. There is a trick to do this which is well known in the representation theory of groups: If one normalizes the χji\chi_{ji}’s, then one gets a unitary matrix. Namely,

χ^ji=χji#Ci#G\hat{\chi}_{ji}=\chi_{ji}\cdot\frac{\sqrt{\#C_{i}}}{\sqrt{\#G}} (24)

defines a unitary matrix. Using unitarity, we can solve for the pisp_{i}^{\prime}s by applying the adjoint of χ^\hat{\chi} to the matrix equation, and use well known properties of the character, to get that

pi(t)=j#Cj#Getl(Cj)(χ^)ji=j#Cj#Getl(Cj)χij.p_{i}(t)=\sum_{j}\frac{\sqrt{\#C_{j}}}{\sqrt{\#G}}e^{-t\,l(C_{j})}(\hat{\chi}^{\dagger})_{ji}=\sum_{j}\frac{\#C_{j}}{\#G}e^{-t\,l(C_{j})}\chi_{ij}^{*}. (25)

To answer question (2), we need to study the convexity of the Kraus-like decomposition. Since we already have a sum rule, we now need to find conditions which ensure that pi0p_{i}\geq 0 for 1in1\leq i\leq n.

Our following theorem gives the answer for question (2):

Theorem 4.4.

The character-induced Kraus-like decomposition of PtP_{t} under the class function length ll is convex if and only if

pi(t=0)=j#Cj#Gl(Cj)χij0p_{i}^{\prime}(t=0)=-\sum_{j}\frac{\#C_{j}}{\#G}l(C_{j})\chi_{ij}^{*}\geq 0 (26)

for all i2i\geq 2.

We split the proof of Theorem 4.4 into two parts, necessity and sufficiency.

Proposition 4.5 (Necessity).

If the character-induced Kraus-like decomposition of PtP_{t} is convex, then

pi(t=0)=j#Cj#Gl(Cj)χij0p_{i}^{\prime}(t=0)=-\sum_{j}\frac{\#C_{j}}{\#G}l(C_{j})\chi_{ij}^{*}\geq 0 (27)

for all i2i\geq 2.

Proof.

First observe that since χ1\chi_{1} is the character for the identity representation, χ1(Ci)=1=χ1i\chi_{1}(C_{i})=1=\chi_{1i} for all ii, and so by the well known orthogonality of characters,

pi(t=0)=χi,χ1=δi1p_{i}(t=0)=\langle\chi_{i},\chi_{1}\rangle=\delta_{i1} (28)

where f,g:=j#Cj#Gf(Cj)g(Cj)\langle f,g\rangle:=\sum_{j}\frac{\#C_{j}}{\#G}f(C_{j})^{*}g(C_{j}) for f,gf,g class functions on GG.

Thus, if pi(t)p_{i}(t) is always positive, it is necessary that pi(t=0)0p_{i}^{\prime}(t=0)\geq 0 for all i2i\geq 2. ∎

We will next show that this condition is actually sufficient for any finite group, and has a group-theoretical explanation. Before demonstrating the sufficiency of this condition in general, we show how one might go about it in a few specific cases. For the sake of our proofs, we state explicitly the nonnegative bounds for the length, even though it is explicit in the definition. We hope these examples highlight how quickly it becomes difficult to study the inequalities pi0p_{i}\geq 0 due to the interdependency of the pip_{i}.

4.2.1. Sufficiency of pi(0)0p_{i}^{\prime}(0)\geq 0 for S3S_{3}

The character table for S3S_{3} is given by:

χ=(111201111)\chi=\begin{pmatrix}1&1&1\\ 2&0&-1\\ 1&-1&1\end{pmatrix} (29)

with #C1=1\#C_{1}=1, #C2=3\#C_{2}=3, #C3=2\#C_{3}=2. This yields

χ^=16(132202132)\hat{\chi}=\frac{1}{\sqrt{6}}\begin{pmatrix}1&\sqrt{3}&\sqrt{2}\\ 2&0&-\sqrt{2}\\ 1&-\sqrt{3}&\sqrt{2}\end{pmatrix}

For convenience, denote l(Ci)l(C_{i}) by lil_{i}, and take l1=0l_{1}=0 so that the identity is of length 0. This yields the following equations for the pip_{i}:

p1(t)\displaystyle p_{1}(t) =16(1+3etl2+2etl3)\displaystyle=\frac{1}{6}\left(1+3e^{-tl_{2}}+2e^{-tl_{3}}\right) (30)
p2(t)\displaystyle p_{2}(t) =16(22etl3)\displaystyle=\frac{1}{6}\left(2-2e^{-tl_{3}}\right) (31)
p3(t)\displaystyle p_{3}(t) =16(13etl2+2etl3).\displaystyle=\frac{1}{6}\left(1-3e^{-tl_{2}}+2e^{-tl_{3}}\right). (32)

Note that p1p_{1}, p2p_{2} are clearly non-negative for all t0t\geq 0. It is also automatic that p1(0)0p_{1}^{\prime}(0)\geq 0 and p2(0)0p_{2}^{\prime}(0)\geq 0.

It is certainly necessary that 6p3(0)=3l22l306p_{3}^{\prime}(0)=3l_{2}-2l_{3}\geq 0. This will in fact turn out to be a sufficient condition.

Proposition 4.6.

In the set up for S3S_{3} described above, all the pip_{i} are non-negative for all t0t\geq 0 if and only if l223l30l_{2}\geq\frac{2}{3}l_{3}\geq 0.

Proof.

The only if direction is clear.

For the if direction, as already noted, for any t0t\geq 0, the non-negativity of p1p_{1} and p2p_{2} is independent of the choice of lil_{i}.

If l223l3l_{2}\geq\frac{2}{3}l_{3} we consider 2 cases

  1. (1)

    Suppose 23l3l2l3\frac{2}{3}l_{3}\leq l_{2}\leq l_{3}. Then

    6p3(t)=3l2etl22l3etl3(3l22l3)etl30.6p_{3}^{\prime}(t)=3l_{2}e^{-tl_{2}}-2l_{3}e^{-tl_{3}}\geq(3l_{2}-2l_{3})e^{-tl_{3}}\geq 0.

    Thus, p3(0)=0p_{3}(0)=0 and p3(t)p_{3}(t) is increasing, so p3(t)0p_{3}(t)\geq 0 for all t0t\geq 0.

  2. (2)

    Suppose l2l3l_{2}\geq l_{3}. Then 6p3(t)13tl2+2etl2=1etl206p_{3}(t)\geq 1-3^{-tl_{2}}+2e^{-tl_{2}}=1-e^{-tl_{2}}\geq 0 for all t0t\geq 0.

As an example of what we have just shown, we will evaluate two natural notions of length on the group S3S_{3}. Let ||=n# of cycles|\cdot|=n-\text{\# of cycles}. Then l1=33=0l_{1}=3-3=0, l2=32=1l_{2}=3-2=1 and l3=31=2l_{3}=3-1=2. Clearly, (l1,l2,l3)=(0,1,2)(l_{1},l_{2},l_{3})=(0,1,2) violates the conditions of the above theorem and so it does not yield a probabilistic interpretation of the pip_{i}.

On the other hand, if ||=n# of cycles|\cdot|=\sqrt{n-\text{\# of cycles}}, then (l1,l2,l3)=(0,1,2)(l_{1},l_{2},l_{3})=(0,1,\sqrt{2}) and we do indeed obtain such a probabilistic interpretation.

4.2.2. Sufficiency of pi(0)0p_{i}^{\prime}(0)\geq 0 for Q8Q_{8}

The character table for Q8Q_{8} is

χ=(1111111111111111111122000)\chi=\begin{pmatrix}1&1&1&1&1\\ 1&1&1&-1&-1\\ 1&1&-1&1&-1\\ 1&1&-1&-1&1\\ 2&-2&0&0&0\end{pmatrix} (33)

where the conjugacy classes associated to the columns, in order, have sizes 1,1,2,21,1,2,2 and 22. This means that

χ^=122(1122211222112221122222000.)\hat{\chi}=\frac{1}{2\sqrt{2}}\begin{pmatrix}1&1&\sqrt{2}&\sqrt{2}&\sqrt{2}\\ 1&1&\sqrt{2}&-\sqrt{2}&-\sqrt{2}\\ 1&1&-\sqrt{2}&\sqrt{2}&-\sqrt{2}\\ 1&1&-\sqrt{2}&-\sqrt{2}&\sqrt{2}\\ 2&-2&0&0&0.\end{pmatrix} (34)

Let li=l(Ci)l_{i}=l(C_{i}), where we always take l1=0l_{1}=0 (i.e the identity element has length 0). This yields the following expressions for the pip_{i}:

p1=\displaystyle p_{1}= 18(1+etl2+2etl3+2etl4+2etl5)\displaystyle\frac{1}{8}\left(1+e^{-tl_{2}}+2e^{-tl_{3}}+2e^{-tl_{4}}+2e^{-tl_{5}}\right) (35)
p2=\displaystyle p_{2}= 18(1+etl2+2etl32etl42etl5)\displaystyle\frac{1}{8}\left(1+e^{-tl_{2}}+2e^{-tl_{3}}-2e^{-tl_{4}}-2e^{-tl_{5}}\right) (36)
p3=\displaystyle p_{3}= 18(1+etl22etl3+2etl42etl5)\displaystyle\frac{1}{8}\left(1+e^{-tl_{2}}-2e^{-tl_{3}}+2e^{-tl_{4}}-2e^{-tl_{5}}\right) (37)
p4=\displaystyle p_{4}= 18(1+etl22etl32etl4+2etl5)\displaystyle\frac{1}{8}\left(1+e^{-tl_{2}}-2e^{-tl_{3}}-2e^{-tl_{4}}+2e^{-tl_{5}}\right) (38)
p5=\displaystyle p_{5}= 18(22etl2)\displaystyle\frac{1}{8}\left(2-2e^{-tl_{2}}\right) (39)

It is clear that p1p_{1} and p5p_{5} are positive for all t0t\geq 0 and also that both p1(0)p_{1}^{\prime}(0) and p5(0)p_{5}^{\prime}(0) are non-negative. Note that when t=0t=0, we have p2=p3=p4=0p_{2}=p_{3}=p_{4}=0. Let us focus momentarily on p2p_{2}. To make p2(0)0p_{2}^{\prime}(0)\geq 0, we must have l22l3+2l4+2l50-l_{2}-2l_{3}+2l_{4}+2l_{5}\geq 0. The analogous computations for p3p_{3} and p4p_{4} show that p2(0),p3(0)p_{2}^{\prime}(0),p_{3}^{\prime}(0) and p4(0)p_{4}^{\prime}(0) are non-negative if and only if l2min{2l4+2l52l3,2l3+2l52l4,2l3+2l42l5}l_{2}\leq\min\{2l_{4}+2l_{5}-2l_{3},2l_{3}+2l_{5}-2l_{4},2l_{3}+2l_{4}-2l_{5}\}. In fact, this condition turns out to be essentially sufficient.

Proposition 4.7.

In the set up for S3S_{3} described above, all the pip_{i} are non-negative for all t0t\geq 0 if and only if 0l2min{2l4+2l52l3,2l3+2l52l4,2l3+2l42l5}0\leq l_{2}\leq\min\{2l_{4}+2l_{5}-2l_{3},2l_{3}+2l_{5}-2l_{4},2l_{3}+2l_{4}-2l_{5}\} and all the lengths are non-negative.

Proof.

The only if direction is clear. For the if direction, as explained above, the non-negativity of p1p_{1} and p5p_{5} for t0t\geq 0 is independent of the choice of the lil_{i}.

Note that l2l_{2}, l3l_{3} and l4l_{4} are symmetric in the equations for the pip_{i} in the sense that by relabeling the conjugacy classes, we can always assume WLOG that l3l4l5l_{3}\leq l_{4}\leq l_{5}. In this case, p4p3p_{4}\leq p_{3} and p4p2p_{4}\leq p_{2}, so it suffices to show that for any t0t\geq 0,

8p4(t)=1+etl2+2(etl5etl3etl4)0.8p_{4}(t)=1+e^{-tl_{2}}+2(e^{-tl_{5}}-e^{-tl_{3}}-e^{-tl_{4}})\geq 0.

To this end, let a=etl3a=e^{-tl_{3}}, b=etl4b=e^{-tl_{4}}, c=etl5c=e^{-tl_{5}}. By assumption, etl2(abc)2e^{-tl_{2}}\geq\left(\frac{ab}{c}\right)^{2}. Thus it suffices to show that 1+(abc)2+2(cab)01+\left(\frac{ab}{c}\right)^{2}+2(c-a-b)\geq 0 if 0cba10\leq c\leq b\leq a\leq 1.

First, we make a change of variable, setting x=cb1x=\frac{c}{b}\leq 1. The desired inequality now reads 2(a+b(1x))1+a2x22(a+b(1-x))\leq 1+\frac{a^{2}}{x^{2}} for 0ba10\leq b\leq a\leq 1. Clearly, it suffices to check b=ab=a, since the left-hand-side is monotonically increasing in bb. So we only need to show that

2a(2x)1+a2x2.2a(2-x)\leq 1+\frac{a^{2}}{x^{2}}.

It is easy to see that the right-hand-side is at least 2ax\frac{2a}{x} by the inequality of arithmetic and geometric means (AM-GM). Thus, our desired inequality is reduced to 42x2x4-2x\leq\frac{2}{x}. This is true since 1x+x2\frac{1}{x}+x\geq 2, by AM-GM once more. ∎

Extensions of these methods allow one to compute that for S4S_{4}, it is sufficient to have pj(0)0p_{j}^{\prime}(0)\geq 0 for all j2j\geq 2 in order to guarantee the non-negativity of all the pip_{i}’s. The computations for S4S_{4} introduce new tools in addition to those used in the cases already presented, which may be useful for similar computations in larger groups. That being said, the proof for S4S_{4} is significantly longer than the proof for the other groups, due to there being essentially three independent variables as opposed to two. The main additional technique involved in the proof for S4S_{4} is a method to reduce the number of variables by Fourier-Motzkin elimination, which is an iterative approach. Due to the length of this computation, we relegate it to Appendix A.

Since the number of cases one needs to consider grows rapidly with the number of variable lengths involved, even with the algorithmic reduction method used for S4S_{4}, it is unlikely that the method is useful for any but low-dimensional groups. This leaves us looking for a more powerful, more general approach, which we take up next.

4.2.3. Sufficiency of pi(0)0p_{i}^{\prime}(0)\geq 0 for General Finite Groups

Following the character-based method introduced in [Lin21], we now prove that the condition

pi(t=0)=j#Cj#Gl(Cj)χij0p_{i}^{\prime}(t=0)=-\sum_{j}\frac{\#C_{j}}{\#G}l(C_{j})\chi_{ij}^{*}\geq 0 (40)

for all i2i\geq 2 is actually sufficient to guarantee that pj(t)0p_{j}(t)\geq 0 for all 1jm1\leq j\leq m and for all t0t\geq 0, where mm is the number of conjugacy classes of GG, and equivalently the number of distinct irreducible representations of GG.

Let {χr}r=1m\{\chi_{r}\}_{r=1}^{m} be the irreducible characters of GG, and for each rr, define σrλ(g)=χr(g)λ(g)\sigma_{r}\lambda(g)=\chi_{r}(g)\lambda(g). Further assume that ll is a class function. Then, since characters span the class functions, PtP_{t} admits a decomposition as Pt=rpr(t)σrP_{t}=\sum_{r}p_{r}(t)\sigma_{r}.

Theorem 4.8.

If there exists ϵ>0\epsilon>0 such that for all 1rm1\leq r\leq m, pr(t)0p_{r}(t)\geq 0 for all 0tϵ0\leq t\leq\epsilon, then for any 1sm1\leq s\leq m, ps(t)0p_{s}(t)\geq 0 for all t0t\geq 0.

Proof.

For any t>0t>0, take nn large such that t/n<ϵt/n<\epsilon. Then, Pt=Pn(tn)=(Ptn)n=(i=1mpi(t/n)σi)nP_{t}=P_{n\left(\frac{t}{n}\right)}=\left(P_{\frac{t}{n}}\right)^{n}=\left(\sum_{i=1}^{m}p_{i}(t/n)\sigma_{i}\right)^{n} since PtP_{t} is a semigroup. For 1rm1\leq r\leq m, let ρr\rho_{r} denote the irreducible representation with character χr\chi_{r}. The tensor product of irreducible representations of a finite group can be completely reduced, and the multiplicity of the irreducible representation ρc\rho_{c} in ρaρb\rho_{a}\otimes\rho_{b}, nabcn_{ab}^{c}, is always non-negative. By extension, we can write σa1σa2σan=i=1mna1a2aniσi\sigma_{a_{1}}\sigma_{a_{2}}\cdots\sigma_{a_{n}}=\sum_{i=1}^{m}n_{a_{1}a_{2}\ldots a_{n}}^{i}\sigma_{i}, where na1a2anin_{a_{1}a_{2}\ldots a_{n}}^{i} is the multiplicity of the irreducible representation ρi\rho_{i} in ρa1ρa2ρan\rho_{a_{1}}\otimes\rho_{a_{2}}\otimes\cdots\otimes\rho_{a_{n}}. Thus,

Pt=a1,,anbpa1(t/n)pa2(t/n)pan(t/n)na1a2anbσb.P_{t}=\sum_{a_{1},\cdots,a_{n}}\sum_{b}p_{a_{1}}(t/n)p_{a_{2}}(t/n)\cdots p_{a_{n}}(t/n)n_{a_{1}a_{2}\ldots a_{n}}^{b}\sigma_{b}. (41)

Note that the coefficients in the above expression are all non-negative. Thus, for any 1bm1\leq b\leq m, we find that pb(t)p_{b}(t), the coefficient of σb\sigma_{b}, is nonnegative for all t0t\geq 0. ∎

Corollary 4.9.

If there exists ϵ>0\epsilon>0 such that for all 1<rm1<r\leq m, pr(t)0p_{r}(t)\geq 0 for all tϵt\leq\epsilon, then for any 1sm1\leq s\leq m, ps(t)0p_{s}(t)\geq 0 for all t0t\geq 0.

Proof.

Since pi(t=0)=δi1p_{i}(t=0)=\delta_{i1} and the pip_{i} are continuous, there is certainly a neighborhood of 0+0^{+} where p1(t)>0p_{1}(t)>0. The conclusion then follows since all the hypotheses of Thm 4.8 are satisfied. ∎

Corollary 4.10.

If the pi(t=0)p_{i}^{\prime}(t=0)’s are positive for all i2i\geq 2, then for any 1sm1\leq s\leq m, ps(t)0p_{s}(t)\geq 0 for all t0t\geq 0.

Proof.

The continuity of the pip_{i}’s, combined with the positivity of the derivative, guarantees that there is a neighborhood of 0+0^{+} in which pi(t)>0p_{i}(t)>0 for all i2i\geq 2. So Corollary 4.9 can be applied. ∎

Now we wish to strengthen Corollary 4.10 so that it suffices that all pi(t=0)p_{i}^{\prime}(t=0)’s are nonnegative for all i2i\geq 2. To do so, it suffices to show that the set of lengths satisfying pi(t)0p_{i}(t)\geq 0 for all t0t\geq 0 is a closed set. This simply follows from the fact that the arbitrary intersection of closed sets is closed, applied to the set of lengths which satisfies {pi(t0)0}\{p_{i}(t_{0})\geq 0\} for t0[0,)t_{0}\in[0,\infty), in the Euclidean topology. We offer a different argument in the next section, which uses structural features from the condition of conditional negativity. This corollary, together with Proposition 4.5, completes the proof of Theorem 4.4.

5. Conditional Negativity Revisited

In the previous section, it was shown that for the abelian finite groups, the notion of conditional negativity of a length function was equivalent with positing the existence of a convex Kraus-like decomposition. At the end of section 3.1, in particular, we used the need to understand this relationship on a more fundamental level to motivate our study of Kraus-like decompositions for general finite groups. In this section, we now give the characterization of the relationship between conditional negativity and the existence of a character-induced convex Kraus-like decomposition in the full finite group case.

In the context of character-induced Kraus-like decompositions, the object of study is the decomposition of G-circulant matrices [Mor98] A=(Aij)A=(A_{ij}) with entries given by

Aij=f(gigj1)=r an irrepprχr(gigj1)A_{ij}=f(g_{i}g_{j}^{-1})=\sum_{r\text{ an irrep}}p_{r}\chi_{r}(g_{i}g_{j}^{-1}) (42)

where ff is a class function on GG. Such a decomposition exists since the irreducible characters form a basis of the set of class functions on GG.

We wish to show the following theorem:

Theorem 5.1 (Decomposition Theorem).

A G-circulant matrix is positive semidefinite if and only if the prp_{r}’s arising in the decomposition are all nonnegative.

By Schoenberg’s theorem, if we can show this, then we obtain the following theorem:

Theorem 5.2.

Suppose ll is a class function on GG satisfying l(e)=0l(e)=0. Then the corresponding character-induced Kraus-like decomposition is convex (i.e. the prp_{r}’s are all nonnegative and satisfy a sum rule) if and only if ll is a conditionally negative-definite length.

Proof of Theorem 5.1.

Define AA to be the matrix with entries f(gigj1)f(g_{i}g_{j}^{-1}). We first prove the if direction. Observe that if the prp_{r}’s are all non-negative, then AA is self-adjoint since

Aij=r an irrepprχr(gigj1)A_{ij}=\sum_{r\text{ an irrep}}p_{r}\chi_{r}(g_{i}g_{j}^{-1}) (43)

and

Aji=r an irrepprχr(gjgi1)=r an irrepprχr(gjgi1)A_{ji}^{*}=\sum_{r\text{ an irrep}}p_{r}\chi_{r}(g_{j}g_{i}^{-1})^{*}=\sum_{r\text{ an irrep}}p_{r}\chi_{r}(g_{j}g_{i}^{-1})^{*} (44)

and we can write

χr(gjgi1)\displaystyle\chi_{r}(g_{j}g_{i}^{-1})^{*} =(n,mr(gj)n,mr(gi)m,n1)\displaystyle=(\sum_{n,m}r(g_{j})_{n,m}r(g_{i})^{-1}_{m,n})^{*} (45)
=(n,mr(gj)n,mr(gi)m,n1)\displaystyle=(\sum_{n,m}r(g_{j})_{n,m}^{*}{r(g_{i})^{-1}_{m,n}}^{*}) (46)
=m,nr(gj)m,n(r(gi)1)n,m\displaystyle=\sum_{m,n}r(g_{j})^{\dagger}_{m,n}(r(g_{i})^{-1})^{\dagger}_{n,m} (47)
=m,nr(gj1)m,nr(gi)n,m\displaystyle=\sum_{m,n}r(g_{j}^{-1})_{m,n}r(g_{i})_{n,m} (48)
=χr(gigj1)\displaystyle=\chi_{r}(g_{i}g_{j}^{-1}) (49)

where we have used the fact that the irreducible representations of finite groups are unitary.

Since AA is self-adjoint, it has a full spectrum. We want to show that the spectrum is nonnegative. It suffices to show that the matrices (χr(gigj1))i,j(\chi_{r}(g_{i}g_{j}^{-1}))_{i,j}, which by the above are self-adjoint, are in fact (up to normalization) orthogonal projections, in particular, that they satisfy

j=1|G|χr(gigj1)χs(gjgk1)=δr,s|G|χr(e)χr(gigk1).\sum_{j=1}^{|G|}\chi_{r}(g_{i}g_{j}^{-1})\chi_{s}(g_{j}g_{k}^{-1})=\delta_{r,s}\frac{|G|}{\chi_{r}(e)}\chi_{r}(g_{i}g_{k}^{-1}). (50)

This would imply that the eigenvalues of AA are simply given by prp_{r}’s up to positive rescaling (with additional multiplicities to account for the size of the subspace with the same eigenvalue).

We prove this result using the idempotent method. Namely, it is known (see e.g., [Jan66]) that within the group algebra of a finite group, one has the idempotents

er=χr(e)|G|gGχr(g1)λ(g)e_{r}=\frac{\chi_{r}(e)}{|G|}\sum_{g\in G}\chi_{r}(g^{-1})\lambda(g) (51)

where λ(g)\lambda(g) is the left-regular representation of the finite group GG. These idempotents satisfy

eres=δr,ser.e_{r}e_{s}=\delta_{r,s}e_{r}. (52)

Comparing coefficients of λ(g)\lambda(g) in erere_{r}e_{r} and ere_{r}, one obtains that

χr(e)|G|hGχr(hg1)χr(h1)=χr(g1)\frac{\chi_{r}(e)}{|G|}\sum_{h\in G}\chi_{r}(hg^{-1})\chi_{r}(h^{-1})=\chi_{r}(g^{-1}) (53)

for all gGg\in G. Setting g1=gigk1g^{-1}=g_{i}g_{k}^{-1}, and rewriting the dummy element hh as h=gjgi1h=g_{j}g_{i}^{-1} and summing over gjg_{j}, equation 53 becomes

χr(e)|G|gjGχr(gjgk1)χr(gigj1)=χr(gigk1)\frac{\chi_{r}(e)}{|G|}\sum_{g_{j}\in G}\chi_{r}(g_{j}g_{k}^{-1})\chi_{r}(g_{i}g_{j}^{-1})=\chi_{r}(g_{i}g_{k}^{-1}) (54)

and so we have shown that equation 50 is true for r=sr=s. When rsr\neq s, one has that eres=0e_{r}e_{s}=0, so the coefficient of each λ(g)\lambda(g) must vanish, yielding

χr(e)|G|gjGχr(gjgk1)χs(gigj1)=0.\frac{\chi_{r}(e)}{|G|}\sum_{g_{j}\in G}\chi_{r}(g_{j}g_{k}^{-1})\chi_{s}(g_{i}g_{j}^{-1})=0. (55)

This completes the proof of equation 50.

For the only if direction, assume AA is positive semi-definite. Applying AA to an eigenvector vv of (χr(gigj1))i,j(\chi_{r}(g_{i}g_{j}^{-1}))_{i,j} yields prvp_{r}v up to a positive rescaling factor (since (χr(gigj1))i,j(\chi_{r}(g_{i}g_{j}^{-1}))_{i,j} is a projection up to positive rescaling; we will fix the overall constant in the next section). So prp_{r} must be nonnegative. Since this must be true for any irrep rr, the prp_{r}’s must all be nonnegative.

Bearing in mind that positive constant multiples of conditionally negative-definite lengths are conditionally negative-definite, we may summarize the theorems of this section as follows:

Theorem 5.3.

Fix a group G={g1=e,,gn}G=\{g_{1}=e,\cdots,g_{n}\} and let MM be a GG-circulant matrix such that M1,1=1M_{1,1}=1. Since the χr\chi_{r} form a basis for class functions, there exists a decomposition Mij=irrepsprχr(gigj1)M_{ij}=\sum_{irreps}p_{r}\chi_{r}(g_{i}g_{j}^{-1}). The following are equivalent:

  • MM is positive semi-definite.

  • Each prp_{r} is non-negative.

  • The length defined by l(gi)=ln(Mi1)l(g_{i})=-\ln(M_{i1}) is conditionally negative-definite.

Thus, if we take prp_{r} to be the pr(t)p_{r}(t) induced by the conditionally negative-definite length ll via our Kraus-like decomposition, then pr(t)0p_{r}(t)\geq 0 for all t0t\geq 0. Conversely, if pr(t)0p_{r}(t)\geq 0, then one obtains canonically a conditionally negative-definite length defined by l(gi)=1tln(Mi1(t))l(g_{i})=-\frac{1}{t}\ln(M_{i1}(t)). Hence, the set of class function lengths such that PtP_{t} has a convex character-induced Kraus-like decomposition is precisely the set of conditionally negative-definite class function lengths.

Based on this equivalence, we offer a different proof that the set of definite linear constraints we obtained to show the nonnegativity of all pi(t)p_{i}(t)’s for all t0t\geq 0 in Corollary 4.10 can be improved to semidefinite linear constraints.

The idea is to interpret that the length function ll as a point in a finite-dimensional space. Following Corollary 4.10, one knows that the pre-image of the positive orthant OO under the map Φ(l)=(Φ2(l),Φ3(l),,Φk(l))\Phi(l)=(\Phi_{2}(l),\Phi_{3}(l),\cdots,\Phi_{k}(l)), where Φi(l)j#Cj#Gl(Cj)χij\Phi_{i}(l)\coloneqq-\sum_{j}\frac{\#C_{j}}{\#G}l(C_{j})\chi_{ij}^{*} from (l(C2),,,l(Ck))(l(C_{2}),\cdots,,l(C_{k})) to k1\mathbb{R}^{k-1}, where kk is the number of conjugacy classes (recall that l(C1)=0l(C_{1})=0), is contained in the set of conditionally negative-definite lengths. Note that Φ\Phi is a map to k1k-1 dimensional space and does not have a coordinate Φ1(l)\Phi_{1}(l), even though Φ1(l)\Phi_{1}(l) is defined.

We now prove two technical lemmas to support our proof that the definite constraints can be replaced by semi-definite constraints:

Lemma 5.4.

Let GG be a group with |G|=n|G|=n and let AnA\subset\mathbb{C}^{n} be the set of conditionally negative-definite lengths on GG, where we identify a function f:Gf:G\rightarrow\mathbb{C} with the vector (f(g1),,f(gn))n(f(g_{1}),\cdots,f(g_{n}))\in\mathbb{C}^{n}. Then AA is closed.

Proof.

We show that ACA^{C} is open. Let fACf\in A^{C}. Then there exists some {αi}i=1n\{\alpha_{i}\}_{i=1}^{n} such that i=1nαi=0\sum_{i=1}^{n}\alpha_{i}=0 and

i,j=1nαiαj¯f(gi1gj)>0.\sum_{i,j=1}^{n}\alpha_{i}\overline{\alpha_{j}}f(g_{i}^{-1}g_{j})>0.

This is an open condition, and so if we obtain a new length fϵf_{\epsilon} by changing ff by an ϵ\epsilon amount in each coordinate, for ϵ\epsilon sufficiently small, it will continue to hold with the same {αi}i=1n\{\alpha_{i}\}_{i=1}^{n}. Thus, fϵACf_{\epsilon}\in A^{C} as well. This proves that ACA^{C} is open and so also proves that AA is closed. ∎

Lemma 5.5.

The map Φ\Phi is invertible.

Proof.

Note that in our proof, the condition eq. 23, which says i=1kpiχi(e)=1\sum_{i=1}^{k}p_{i}\chi_{i}(e)=1, translates to a linear condition on the Φi(l)\Phi_{i}(l) which involves all the Φi(l)\Phi_{i}(l) and in particular Φ1(l)\Phi_{1}(l). Thus, given Φ(l)=(Φ2(l),,Φk(l))\Phi(l)=(\Phi_{2}(l),\ldots,\Phi_{k}(l)), we can determine Φ1(l)\Phi_{1}(l).

Observe that eq. 40 gives an explicit expression for {Φi(l)}i=1k\{\Phi_{i}(l)\}_{i=1}^{k}. Note that it is immediately clear from character theory that χij{\chi_{ij}^{*}} is invertible. It then follows that the map from ll to (Φ1(l),Φ2(l),,Φk(l))(\Phi_{1}(l),\Phi_{2}(l),\ldots,\Phi_{k}(l)) is invertible as well. Due to the addition of Φ1\Phi_{1}, note that this is not the map Φ\Phi.

Putting this all together, it follows that given x=(Φ2(l),,Φk(l))x=(\Phi_{2}(l),\ldots,\Phi_{k}(l)), we can determine Φ1(l)\Phi_{1}(l) and then uniquely recover l=Φ1(x)l=\Phi^{-1}(x). ∎

Proposition 5.6.

Fix l(C1)=0l(C_{1})=0, and take l(Ci)l(C_{i}) to be variable for i=2,,ki=2,\ldots,k. If the pi(t=0)p_{i}^{\prime}(t=0)’s are nonnegative for all i2i\geq 2, then for any 1sk1\leq s\leq k, ps(t)0p_{s}(t)\geq 0 for all t0t\geq 0.

Proof.

Since Φ\Phi is invertible and linear, Φ1\Phi^{-1} is continuous. Thus, Φ1(O¯)Φ1(O)¯\Phi^{-1}(\overline{O})\subset\overline{\Phi^{-1}(O)}, where OO is the set of points with coordinates xi>0x_{i}>0 for all i=1,2,,k1i=1,2,\ldots,k-1. The latter is a subset of the set of conditionally negative-definite lengths (since this set is closed). Hence, any length whose image lies in O¯\bar{O} is conditionally negative-definite. Thus, we have strengthened the constraints from definite linear constraints (defined by OO) to semidefinite linear constraints (defined by O¯\bar{O}). ∎

5.1. Values and Multiplicities of Eigenvalues

We further show that the rank of (χr(gigj1))i,j(\chi_{r}(g_{i}g_{j}^{-1}))_{i,j} is given by χr(e)2\chi_{r}(e)^{2}. To prove this, first note that χr(gigj1)=χr(gjgi1)\chi_{r}(g_{i}g_{j}^{-1})=\chi_{r}(g_{j}g_{i}^{-1})^{*}, so (χr(gigj1))gi,gj(\chi_{r}(g_{i}g_{j}^{-1}))_{g_{i},g_{j}} is Hermitian, and is fully diagonalizable. Furthermore, by rescaling equation 50, we get that

j=1|G|(χr(e)|G|χr(gigj1))(χr(e)|G|χs(gjgk1))=δr,sχr(e)|G|χr(gigk1),\sum_{j=1}^{|G|}\left(\frac{\chi_{r}(e)}{|G|}\chi_{r}(g_{i}g_{j}^{-1})\right)\left(\frac{\chi_{r}(e)}{|G|}\chi_{s}(g_{j}g_{k}^{-1})\right)=\delta_{r,s}\frac{\chi_{r}(e)}{|G|}\chi_{r}(g_{i}g_{k}^{-1}), (56)

and so the χr(e)|G|(χr(gigj1))gi,gj\frac{\chi_{r}(e)}{|G|}(\chi_{r}(g_{i}g_{j}^{-1}))_{g_{i},g_{j}} is are orthogonal projection matrices, with eigenvalues 0 or 11. Thus, to compute its rank, we just need its trace, which is |G|χr(e)|G|(χr(e))=(χr(e))2|G|\cdot\frac{\chi_{r}(e)}{|G|}(\chi_{r}(e))=(\chi_{r}(e))^{2}. It follows that the multiplicity of the eigenvalue 11 in the projection matrix corresponding to irrep rr is (χr(e))2(\chi_{r}(e))^{2}.

Rewriting the decomposition of AA given by eq. 42 in terms of the projections, one has that

f(gigj1)=r an irrep(|G|χr(e)pr)(χr(e)|G|χr(gigj1))f(g_{i}g_{j}^{-1})=\sum_{r\text{ an irrep}}\left(\frac{|G|}{\chi_{r}(e)}p_{r}\right)\left(\frac{\chi_{r}(e)}{|G|}\chi_{r}(g_{i}g_{j}^{-1})\right) (57)

and so the eigenvalues are given by (|G|χr(e)pr)\left(\frac{|G|}{\chi_{r}(e)}p_{r}\right). Since these are orthogonal projections, we thus have shown that the multiplicity of each eigenvalue |G|χr(e)pr\frac{|G|}{\chi_{r}(e)}p_{r} in the matrix given by eq. 42 is given by (χr(e))2(\chi_{r}(e))^{2}.

6. Conclusion

In this article, we have presented a new way to approach semigroups acting on group algebras, namely, Kraus-like decompositions. By modifying the action of possible operators which can be used in the sum decomposition of a semigroup, our approach allows us to introduce operator decompositions in problems where they were not readily available before. In particular, we are able to explore quantum channels induced by length functions in the context of group algebras. For length functions that are additionally class functions, we obtain several equivalent necessary and sufficient conditions on the length function for the semigroup PtP_{t} to be a quantum channel for all time t0t\geq 0.

Specifically, for character-induced Kraus-like decompositions, we have proven that a set of semidefinite linear constraints is necessary and sufficient to guarantee the positivity of the coefficients appearing in the decomposition for all t0t\geq 0. We also proved that this same set of semidefinite linear constraints suffices to characterize the conditionally negative-definite lengths, a class of length functions that are of independent interest. Using Schoenberg’s theorem, we further relate this to conditions on the complete positivity of semigroups PtP_{t} induced by lengths which are class functions. These constraints follow from the fact that for a semigroup PtP_{t}, the property of admitting a Kraus-like decomposition can be checked globally using a more local condition. Specifically, if one can show that PtP_{t} admits a Kraus-like decomposition for t>0t>0 sufficiently small, then the same can be concluded for all t>0t>0.

The coefficients of our pip_{i}’s are defined canonically. Moreover, they are nonnegative and satisfy a sum rule. Thus, the coefficients may have physical meaning (under appropriate rescaling) as probabilities. We plan to explore these physical connections in a follow-up paper.

Acknowledgments

R.L. acknowledges the research support of ARO grant W911NF-20-1-0082 through the MURI project “Toward Mathematical Intelligence and Certifiable Automated Reasoning: From Theoretical Foundations to Experimental Realization.” J.B. acknowledges the support of the Natural Sciences and Engineering Research Council of Canada (NSERC), [funding reference number 557353-2021]. J. B. a été financé par le Conseil de recherches en sciences naturelles et en génie du Canada (CRSNG), [numéro de référence 557353-2021].

The authors thank Radakrishnan Balu for helpful discussions at an intermediate stage of their work and helpful comments on the manuscript. The authors thank Arthur Jaffe for overall guidance, support, and feedback. They also thank Eric Carlen for bringing to their attention the complementary results of [Wir22], which give structural characterizations of extensions of completely positive maps.

Appendix A Proof of Stability Condition for S4S_{4}

We have a direct proof that the a priori necessary conditions on the pip_{i} are sufficient to ensure the existence of a convex Kraus-like decomposition in the context of S4S_{4} as well. We use methods that are similar to those presented above but more computationally involved due to the parameter space being of a larger dimension.

The character table for S4S_{4} is

χ=(1111131011201203101111111)\chi=\begin{pmatrix}1&1&1&1&1\\ 3&1&0&-1&-1\\ 2&0&-1&2&0\\ 3&-1&0&-1&1\\ 1&-1&1&1&-1\end{pmatrix}

and the sizes of the conjugacy classes are (Ci)i=1,,5=(1,6,8,3,6)(C_{i})_{i=1,\ldots,5}=(1,6,8,3,6).

Accordingly,

p1=\displaystyle p_{1}= 124(1+6etl2+8etl3+3etl4+6etl5)\displaystyle\frac{1}{24}\left(1+6e^{-tl_{2}}+8e^{-tl_{3}}+3e^{-tl_{4}}+6e^{-tl_{5}}\right)
p2=\displaystyle p_{2}= 18(1+2etl2etl42etl5)\displaystyle\frac{1}{8}\left(1+2e^{-tl_{2}}-e^{-tl_{4}}-2e^{-tl_{5}}\right)
p3=\displaystyle p_{3}= 112(14etl3+3etl4)\displaystyle\frac{1}{12}\left(1-4e^{-tl_{3}}+3e^{-tl_{4}}\right)
p4=\displaystyle p_{4}= 18(12etl2etl4+2etl5)\displaystyle\frac{1}{8}\left(1-2e^{-tl_{2}}-e^{-tl_{4}}+2e^{-tl_{5}}\right)
p5=\displaystyle p_{5}= 124(16etl2+8etl3+3etl46etl5).\displaystyle\frac{1}{24}\left(1-6e^{-tl_{2}}+8e^{-tl_{3}}+3e^{-tl_{4}}-6e^{-tl_{5}}\right).

Thus, at t=0t=0, p1=1p_{1}=1 and all the others are equal to 0.

The constraint that the probabilities be nonnegative at time ϵ>0\epsilon>0 for ϵ\epsilon arbitrarily small tells us that pi(t=0)0p_{i}^{\prime}(t=0)\geq 0 for 2i52\leq i\leq 5, In particular,

24p2(0)\displaystyle 24p_{2}^{\prime}(0) =6l2+3l4+6l50\displaystyle=-6l_{2}+3l_{4}+6l_{5}\geq 0
24p3(0)\displaystyle 24p_{3}^{\prime}(0) =8l36l40\displaystyle=8l_{3}-6l_{4}\geq 0
24p4(0)\displaystyle 24p_{4}^{\prime}(0) =6l2+3l46l50\displaystyle=6l_{2}+3l_{4}-6l_{5}\geq 0
24p5(0)\displaystyle 24p_{5}^{\prime}(0) =6l28l33l4+6l50.\displaystyle=6l_{2}-8l_{3}-3l_{4}+6l_{5}\geq 0.

We can clean this up a little bit by setting a=etl2a=e^{-tl_{2}}, b=etl3b=e^{-tl_{3}}, c=etl4c=e^{-tl_{4}}, d=etl5d=e^{-tl_{5}}.

Theorem A.1.

In the set up for S4S_{4} described above, all the pip_{i} are non-negative for all t0t\geq 0 if and only if the lengths are all non-negative and the above necessary conditions hold. Expressed in terms of a,b,ca,b,c and dd, this is equivalent to the following system of inequalities:

0a,b,c,d1.0\leq a,b,c,d\leq 1.
a2c1d21\displaystyle a^{2}c^{-1}d^{-2}\geq 1
b4c31\displaystyle b^{-4}c^{3}\geq 1
a2c1d21\displaystyle a^{-2}c^{-1}d^{2}\geq 1
a6b8c3d61.\displaystyle a^{-6}b^{8}c^{3}d^{-6}\geq 1.

.

Proof.

We first express our goal, the non-negativity of the pip_{i}, in terms of a,b,ca,b,c and dd as follows:

1+2ac2d0\displaystyle 1+2a-c-2d\geq 0
14b+3c0\displaystyle 1-4b+3c\geq 0
12ac+2d0\displaystyle 1-2a-c+2d\geq 0
16a+8b+3c6d0\displaystyle 1-6a+8b+3c-6d\geq 0

To prove these inequalities, we must consider a number of cases.

The a=da=d case

We want to show that

b4c31\displaystyle b^{-4}c^{3}\geq 1
a12b8c31\displaystyle a^{-12}b^{8}c^{3}\geq 1

implies

14b+3c0\displaystyle 1-4b+3c\geq 0
112a+8b+3c0,\displaystyle 1-12a+8b+3c\geq 0,

since the other inequalities become trivial in this case. The inequality 14b+3c01-4b+3c\geq 0 holds true since 14b+3c14b+3b4/31-4b+3c\geq 1-4b+3b^{4/3}, which is nonnegative on [0,1][0,1] since its derivative is nonpositive and it evaluates to 0 at b=1b=1.

Note that ab2/3c1/4a\leq b^{2/3}c^{1/4}, so 112a+8b+3c112b2/3c1/4+8b+3c1-12a+8b+3c\geq 1-12b^{2/3}c^{1/4}+8b+3c. We want to minimize this subject to b4c3b^{4}\leq c^{3}, and show that the minimum is at least 0. If b4=c3b^{4}=c^{3}, we obtain that the lower bound is given by 112b2/3b1/3+8b+3b4/3=14b+3b4/31-12b^{2/3}b^{1/3}+8b+3b^{4/3}=1-4b+3b^{4/3}. This is the same situation as in the previous case and as such, is non-negative for b[0,1]b\in[0,1]. Taking a partial derivative with respect to bb, we get 8b1/3c1/4+88b1/3b1/3+8=0-8b^{-1/3}c^{1/4}+8\leq-8b^{-1/3}b^{1/3}+8=0, so 112b2/3c1/4+8b+3c1-12b^{2/3}c^{1/4}+8b+3c is non-negative at points (b,c)(b,c) where b[0,c3/4]b\in[0,c^{3/4}]. This is precisely the range of values of bb allowed by the first inequality, so we are done.

Relaxing the condition a=da=d: We now treat the full set of inequalities. We rewrite our assumptions in terms of cc:

cmin(a2d2,a2d2)c\leq\min(a^{2}d^{-2},a^{-2}d^{2})

and

cmax(b4/3,a2d2b8/3).c\geq\max(b^{4/3},a^{2}d^{2}b^{-8/3}).

In particular, we have that

max(b4/3,a2d2b8/3)min(a2d2,a2d2),\max(b^{4/3},a^{2}d^{2}b^{-8/3})\leq\min(a^{2}d^{-2},a^{-2}d^{2}),

and so

b2/3a\displaystyle b^{2/3}a db2/3\displaystyle\leq d\leq b^{2/3}
b2/3d\displaystyle b^{2/3}d ab2/3\displaystyle\leq a\leq b^{2/3}

are necessary conditions.

We first prove that 1+2ac2d01+2a-c-2d\geq 0. Note that everything is symmetric in aa and in dd, so without loss of genereality, suppose ada\geq d, then

1+2ac2d1+2ad2/a22d=2(ad)+1d2/a20.1+2a-c-2d\geq 1+2a-d^{2}/a^{2}-2d=2(a-d)+1-d^{2}/a^{2}\geq 0.

Next, we show 1+2dc2a01+2d-c-2a\geq 0. Continuing to assume that ada\geq d, we note that

1+2dc2a1+2dd2/a22a.1+2d-c-2a\geq 1+2d-d^{2}/a^{2}-2a.

We want to minimize 1+2dd2/a22a1+2d-d^{2}/a^{2}-2a over the region of permitted values of aa and dd, so we start by fixing bb and aa, and requiring dd to lie between b2/3ab^{2/3}a and b2/3b^{2/3}. (Note that we will be a bit lackadaisical with the constraint on aa. This is OK since we just need a lower bound on the constraint region, which is guaranteed if we work with a region containing the constraint region.)

Taking a partial derivative with respect to dd yields 22d/a22-2d/a^{2} which changes sign at d=a2d=a^{2}, from positive to negative, so this is a maximum. We must evaluate this expression at the boundary of the allowed dd values as well. These are d=b2/3ad=b^{2/3}a and d=ad=a (since, by assumption, dad\leq a). Evaluating at d=b2/3ad=b^{2/3}a, we obtain 12(1b2/3)ab4/312(1b2/3)b2/3b4/3=12b2/3+b4/3=(1b2/3)201-2(1-b^{2/3})a-b^{4/3}\geq 1-2(1-b^{2/3})b^{2/3}-b^{4/3}=1-2b^{2/3}+b^{4/3}=(1-b^{2/3})^{2}\geq 0, where in the first inequality, we used ab2/3a\leq b^{2/3}. For d=ad=a, we get 1+2a12a=01+2a-1-2a=0.

Next, we want to show that

14b+3c0.1-4b+3c\geq 0.

Since cb4/3,c\geq b^{4/3}, we obtain 14b+3c14b+3b4/301-4b+3c\geq 1-4b+3b^{4/3}\geq 0, as we showed earlier.

Finally, we want to show

16(a+d)+8b+3c0.1-6(a+d)+8b+3c\geq 0.

It is natural to split into two cases.

Case 1: b2adb^{2}\leq ad: In this case, we have c(ad)2b8/3c\geq(ad)^{2}b^{-8/3} as the stronger lower bound. Using this bound yields

16(a+d)+8b+3c16(a+d)+8b+3(ad)2b8/3.1-6(a+d)+8b+3c\geq 1-6(a+d)+8b+3(ad)^{2}b^{-8/3}.

We will minimize the sum 16(a+d)+8b+3(ad)2b8/31-6(a+d)+8b+3(ad)^{2}b^{-8/3} for fixed bb. We can still assume without loss of generality that ada\geq d. Recall that b2/3adb^{2/3}a\leq d. So we want to minimize 16(a+d)+8b+3(ad)2b8/31-6(a+d)+8b+3(ad)^{2}b^{-8/3} as a function of aa and dd subject to the two constraints

b2/3adab2/3b^{2/3}a\leq d\leq a\leq b^{2/3}

and

adb2ad\geq b^{2}

The shape of this domain is a three-sided region dependent on the value of bb. In the plane with aa as its y-axis and dd as its x-axis, we have a region upper bounded by the line a=b2/3a=b^{2/3}, right-bounded by the line d=ad=a, and southwest-bounded by the curve ad=b2ad=b^{2}. The corners of the region are (d=b4/3,a=b2/3)(d=b^{4/3},a=b^{2/3}), (d=b2/3,a=b2/3)(d=b^{2/3},a=b^{2/3}), and (d=b,a=b)(d=b,a=b).

For fixed aa, we can differentiate with respect to dd, obtaining 6+6a2db8/3-6+6a^{2}db^{-8/3}. Note that aba\geq b and db4/3d\geq b^{4/3} on this region, so we have an upper bound on this derivative of 6+6b2b4/3b8/3=6+6b2/30.-6+6b^{2}b^{4/3}b^{-8/3}=-6+6b^{2/3}\leq 0. Thus, the derivative is non-positive, and the function is non-increasing in dd. As such, it will suffice to take dd lying on the boundary to the right, where d=ad=a. We are thus left with considering d=ad=a in the expression 112a+8b+3a4b8/31-12a+8b+3a^{4}b^{-8/3}444Note that we have not considered this case yet. When we had d=ad=a earlier, we were directly considering the inequalities we wanted to prove, not the strengthened one we are working with here.. Taking a derivative with respect to aa, we obtain 12+12a3b8/3-12+12a^{3}b^{-8/3}, which equals 0 when a3b8/3=1a^{3}b^{-8/3}=1, which is to say, when a=b8/9a=b^{8/9}. Note that 12+12a3b8/3<0-12+12a^{3}b^{-8/3}<0 for a<b8/9a<b^{8/9} and 12+12a3b8/3>0-12+12a^{3}b^{-8/3}>0 for a>b8/9a>b^{8/9}, so we obtain a minimum at a=b8/9a=b^{8/9} and it suffices to use this value going forward. Thus, we want to minimize 112b8/9+8b+3b32/9b8/3=19b8/9+8b.1-12b^{8/9}+8b+3b^{32/9}b^{-8/3}=1-9b^{8/9}+8b. This has derivative 88b1/908-8b^{-1/9}\leq 0, so it suffices to verify 19b8/9+8b01-9b^{8/9}+8b\geq 0 at b=1b=1, which certainly holds.

Case 2: b2adb^{2}\geq ad: Then cb4/3c\geq b^{4/3} is the meaningful lower bound on cc. So we wish to minimize

16(a+d)+8b+3b4/31-6(a+d)+8b+3b^{4/3}

on the domain

b2/3adab3/2b^{2/3}a\leq d\leq a\leq b^{3/2}

and

adb2.ad\leq b^{2}.

This domain is also a three-sided region dependent on the value of bb. It is bounded on the left by the line d=b2/3ad=b^{2/3}a, on the right by the line a=da=d, on the northeast by the curve ad=b2ad=b^{2}. The corners of the region are (0,0)(0,0), (d=b4/3,a=b2/3)(d=b^{4/3},a=b^{2/3}), and (b,b)(b,b).

The partial derivative of 16(a+d)+8b+3b4/31-6(a+d)+8b+3b^{4/3} with respect to dd is 6-6, so due to the geometry of the domain boundary, this expression is minimized on the d=ad=a boundary if aba\leq b and on the ad=b2ad=b^{2} boundary if aba\geq b.

For aba\leq b and d=ad=a, we have 16(a+d)+8b+3b4/3=112a+8b+3b4/314b+3b4/31-6(a+d)+8b+3b^{4/3}=1-12a+8b+3b^{4/3}\geq 1-4b+3b^{4/3}. This is non-negative, as we showed earlier.

For aba\geq b and ad=b2ad=b^{2}, we obtain 16(a+d)+8b+3b4/3=16(a+b2/a)+8b+3b4/31-6(a+d)+8b+3b^{4/3}=1-6(a+b^{2}/a)+8b+3b^{4/3}. Taking a derivative with respect to aa, one 6(1b2/a2)-6(1-b^{2}/a^{2}). Thus, 16(a+b2/a)+8b+3b4/31-6(a+b^{2}/a)+8b+3b^{4/3} has a maximum at a=ba=b, and achieves its minimums at a boundary value of aa. The boundary values of aa in the region we are considering are a=b2/3a=b^{2/3} and a=ba=b. If a=b2/3a=b^{2/3}, we get 1+8b3b4/36b2/31+8b-3b^{4/3}-6b^{2/3}, which has derivative 4(1+b1/3)2b1/30-\frac{4(-1+b^{1/3})^{2}}{b^{1/3}}\leq 0. Thus, 1+8b3b4/36b2/31+8b-3b^{4/3}-6b^{2/3} is minimized when b=1b=1, yielding 0. If a=ba=b, we get 14b+3b4/31-4b+3b^{4/3}, which is non-negative, as shown earlier. ∎

References

  • [CDH+20] Shawn X. Cui, Dawei Ding, Xizhi Han, Geoffrey Penington, Daniel Ranard, Brandon C. Rayhaun, and Zhou Shangnan. Kitaev’s quantum double model as an error correcting code. Quantum, 4:331, September 2020.
  • [Cho75] Man Duen Choi. Completely positive linear maps on complex matrices. Linear Algebra Appl., 10:285–290, 1975.
  • [Jan66] G. J. Janusz. Primitive idempotents in group algebras. Proceedings of the American Mathematical Society, 17:520–523, 1966.
  • [JPPP17] Marius Junge, Carlos Palazuelos, Javier Parcet, and Mathilde Perrin. Hypercontractivity in group von Neumann algebras. Mem. Amer. Math. Soc., 249(1183):xii+83, 2017.
  • [Kit03] A. Yu Kitaev. Fault-tolerant quantum computation by anyons. Annals of Physics, 303:2–30, 2003.
  • [Lin21] Robert Lin. Certain linear combinations of exponential functions are positive under semidefinite linear constraints, November 2021. arXiv:2112.00134.
  • [Mor98] K. E. Morrison. A generalization of circulant matrices for non-abelian groups, 1998. Research report.
  • [NC10] Michael A. Nielsen and Isaac L. Chuang. Quantum Computation and Quantum Information. Cambridge University Press, 2010. Tenth anniversary edition.
  • [NS06] Sergey Neshveyev and Erling Størmer. Dynamical entropy in operator algebras, volume 50 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. [Results in Mathematics and Related Areas. 3rd Series.] A Series of Modern Surveys in Mathematics. Springer-Verlag, Berlin, 2006.
  • [Sch38] Isaac J Schoenberg. Metric spaces and positive definite functions. Transactions of the American Mathematical Society, 44(3):522–536, 1938.
  • [Sti55] W. Forrest Stinespring. Positive functions on CC^{*}-algebras. Proc. Amer. Math. Soc., 6:211–216, 1955.
  • [Uhl77] A. Uhlmann. Relative entropy and the Wigner-Yanase-Dyson-Lieb concavity in an interpolation theory. Comm. Math. Phys., 54(1):21–32, 1977.
  • [vdBF75] Christian van den Berg and Gunnar Forst. Potential Theory on Locally Compact Abelian Groups. Springer, 1975.
  • [Wir22] Melchior Wirth. Christensen-Evans theorem and extensions of GNS-symmetric quantum markov semigroups, March 2022. arXiv:2203.00341.