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

Lower bound for the TT count via unitary stabilizer nullity

Jiaqing Jiang [email protected] Institute for Quantum Computing, Baidu Research, Beijing 100193, China Computing and Mathematical Sciences, California Institute of Technology, Pasadena, California, USA    Xin Wang [email protected] Institute for Quantum Computing, Baidu Research, Beijing 100193, China
Abstract

We introduce magic measures to quantify the nonstabilizerness of multiqubit quantum gates and establish lower bounds on the TT count for fault-tolerant quantum computation. First, we introduce the stabilizer nullity of multi-qubit unitary, which is based on the subgroup of the quotient Pauli group associated with the unitary. This unitary stabilizer nullity extends the state-stabilizer nullity by Beverland et al. to a dynamic version. In particular, we show this nonstabilizerness measure has desirable properties such as subadditivity under composition and additivity under tensor product. Second, we prove that a given unitary’s stabilizer nullity is a lower bound for the TT count, utilizing the above properties in gate synthesis. Third, we compare the state- and the unitary-stabilizer nullity, proving that the lower bounds for the TT count obtained by the unitary-stabilizer nullity are never less than the state-stabilizer nullity. Moreover, we show an explicit nn-qubit unitary family of unitary-stabilizer nullity 2n2n, which implies that its TT count is at least 2n2n. This gives an example where the bounds derived by the unitary-stabilizer nullity strictly outperform the state-stabilizer nullity by a factor of 22. We finally showcase the advantages of unitary-stabilizer nullity in estimating the TT count of quantum gates with interests.

I Introduction

A quantum circuit comprised of only Clifford gates confers no computational advantage since it can be efficiently simulated on a classical computer [1, 2]. However, the addition of TT gates into Clifford circuits [3, 4] can achieve universal quantum computation. The minimum number of TT gates used, or the TT count, plays a role in quantum computation. The study of TT count has applications in quantum simulation, quantum error correction, and quantum circuit synthesis. Specifically, in the theoretical study of classically simulating quantum computation, TT count serves as a quantifier of difficulty for such simulation in many algorithms [5, 6, 7]. It is shown that quantum circuits composed of Clifford gates and tt TT gates can be simulated on a classical computer in time, which is exponential in tt [5, 7, 8, 9]. Better estimation of the TT-count gives better analysis of the performance for those algorithms. On the other hand, in the framework of fault-tolerant quantum computation [10, 11], TT count dominates the cost of computation. This is because in many promising fault-tolerant architectures, the Clifford gates can be implemented transversally [12] and the required TT gate for universality can be performed via state injection [13, 10]. The key of this resolution is to perform magic state distillation [11, 14], which is a resource-intensive subroutine with a costly overhead [14, 15] (see Refs.[16, 17, 18, 19, 20, 21, 22, 23] for recent progress).

Quantum resource theories [24] offer a powerful framework for studying quantum resources for information processing. The resource theory of TT count has found applications in magic state distillation and non-Clifford gate synthesis [25]. To better realize quantum computation, there are many circuit synthesis algorithms that aim to reduce the number of TT gates in the implementation of a given unitary operator  [26, 27, 28, 29, 30], providing a better upper bound for its TT count.

This work aims to study lower bounds for the TT count. A lower bound for the TT count estimates the minimum magic resources [25] required for quantum computation, and gives a lower bound for the runtime of the classical simulation algorithms [8, 5, 6] based on Clifford+TT gate sets. Specifically, for a given nn-qubit unitary UU, write t(U)t(U) as the minimum number of TT gates among all implementations of UU over the Clifford+TT gate set, i.e., the TT count equals t(U)t(U). The task is to calculate a quantity v(U)v(U) such that v(U)t(U)v(U)\leq t(U). A good lower bound is a v(U)v(U) which is as large as possible, and the time complexity for computing v(U)v(U) is as small as possible. From the theoretical perspective, the explicit expressions of v(U)v(U) for some nn-qubit unitary families are useful and desirable. For the decision version of the lower-bound task, Gosset et al. [29] gives an algorithm which for any integer mm and nn-qubit unitary UU, it decides whether the TT count for UU is less than mm, with both time and space complexity of O(2nmpoly(m,2n))O(2^{nm}poly(m,2^{n})). Although this algorithm can be used to calculate the exact value of the TT count, the time complexity is quite high. Even for small mm such as m=nm=n, the complexity will scale as O(2cn2)O(2^{cn^{2}}) for some constant cc. Other methods for deriving lower bounds for the TT count are mostly based on magic monotones. Those magic monotones are primarily designed for quantifying the non-Clifford resources for constructing quantum states, such as quantifying the number of the magic state |T|T\rangle needed for constructing a given state |ψ|\psi\rangle. Those magic monotones could also naturally induce a lower bound for the TT count for a given unitary. Note that due to the adaptive operations allowed in the model of computation with magic states, which are not allowed in the model of computation with unitary, the lower bound of the TT count for UU induced by those magic monotones can be strictly smaller than the TT count for UU [31, 32, 29]. This subtle distinction is clarified in Appendix B.

The relevant magic monotones can be summarized as follows. For a pure state |ψ|\psi\rangle, the stabilizer rank [7] is defined as the minimum number of nonzero coefficients when decomposing |ψ|\psi\rangle onto a linear combination of stabilizer states. The stabilizer extent [7, 33] is defined as the square of the minimum l1l_{1} norm of the above coefficients. It relates to the approximate stabilizer rank, that is the stabilizer rank of states, which approximate |ψ|\psi\rangle within given precision. The robustness of nonstabilizerness [31] is also related to the minimum l1l_{1} norm but defined for mixed states. The stabilizer rank can also be generalized to mixed states in multiple ways [8]. The channel robustness and magic capacity [34] are closely related to the robustness of magic. They can be used for quantifying the non-Clifford resources for multiqubit operations. All of the above magic monotones are more or less involving optimization over the set of stabilizer states, and thus are costly to compute since the cardinality of stabilizer states increases as 2Ω(n2)2^{\Omega(n^{2})} [35]. Moreover, since they involve optimization, it is hard to derive explicit expressions for those magic monotones for nn-qubit unitaries or quantum states. Beyond the magic monotones based on optimization, Beverland et al. [25] proposed the state-stabilizer nullity based on quantities related to the Pauli group. More specifically, let s(|ψ)s(|\psi\rangle) be the size of the stabilizer of |ψ|\psi\rangle, where the stabilizer is the subgroup of the Pauli group for which |ψ|\psi\rangle is a +1+1 eigenvector, the state-stabilizer nullity of |ψ|\psi\rangle is defined as vs(|ψ)=nlog2s(|ψ)v_{s}(|\psi\rangle)=n-\log_{2}s(|\psi\rangle). In general, the state-stabilizer nullity can be computed in time O(24n)O(2^{4n}). Moreover, for some useful unitary families {Un}n\{U_{n}\}_{n}, the explicit expressions of the state-stabilizer nullity of Un|+nU_{n}|+\rangle^{\otimes n} can be easily computed. And this quantity can be used to lower bound the TT count for UnU_{n}. An example is showing that the nn-qubit Toffoli gate has TT count nn, implying that implementing such a gate needs at least nn |T|T\rangle states or TT gates. The state-stabilizer nullity is designed for quantifying non-Clifford resources for constructing states, although it can be used to derive lower bounds for the TT count for unitaries, a “unitary-stabilizer nullity” is desirable for better quantifying the nonstabilizerness and estimate the TT count of multiqubit gates. It is unclear and nontrivial to generalize the definition from states to unitaries. Beyond all the above, there are also magic monotones defined for qudit systems of odd prime dimension, which cannot be directly used for qubit systems [36, 23].

In this work, we define the unitary-stabilizer nullity. Instead of counting the operators in the Pauli group, which stabilize |ψ|\psi\rangle, for a target UU we count the ±1\pm 1s in the Pauli transfer matrix. In other words, we calculate the size of the intersection between U𝒫nUU{\cal P}_{n}U^{\dagger} and 𝒫n{\cal P}_{n}, where 𝒫n{\cal P}_{n} is the quotient Pauli group, that is the Pauli-group modulo phases. Denoting the size as s(U)s(U), the unitary-stabilizer nullity is defined as v(U)=2nlog2s(U)v(U)=2n-\log_{2}s(U). Intuitively, s(U)s(U) quantifies the similarity between UU and Clifford unitaries, where for Clifford unitary UU, s(U)=4ns(U)=4^{n}. To use v()v(\cdot) to derive a lower bound for the TT count, we prove two key properties: v()v(\cdot) satisfies faithfullness, i.e. v(U)0v(U)\geq 0 for any UU, v(U)=0v(U)=0 if and only if UU is Clifford, and subadditivity under composition, i.e. v(UV)v(U)+v(V)v(U\cdot V)\leq v(U)+v(V) or equivalently, s(U)s(V)4ns(UV)s(U)s(V)\leq 4^{n}s(UV). We further apply these properties to establish the lower bound for the TT count.

We then compare the lower bounds of the TT count obtained by the state-stabilizer nullity and the unitary-stabilizer nullity. We show that the bound given by the unitary-stabilizer nullity will never be smaller than the bound given by the state-stabilizer nullity, that is v(U)vs(U|ϕ),v(U)\geq v_{s}(U|\phi\rangle), for any stabilizer state |ϕ|\phi\rangle. We also prove that for any diagonal UU, v(U)=vs(U|+n)v(U)=v_{s}(U|+\rangle^{\otimes n}), which in turn shows that for any diagonal UU, the maximum lower bound obtained by the state-stabilizer nullity, i.e., vs(U|ϕ)v_{s}(U|\phi\rangle) for some stabilizer state |ϕ|\phi\rangle, is achieved by |ϕ=|+n|\phi\rangle=|+\rangle^{\otimes n}. Moreover, we give an explicit nn-qubit unitary family, which has unitary-stabilizer nullity 2n2n. Since the state-stabilizer nullity is always less than nn, such a unitary family illustrates that the lower bound of the TT count given by the unitary-stabilizer nullity can be strictly better than the bound given by the state-stabilizer nullity by a factor of 22.

In general, the unitary-stabilizer nullity can be computed in time O(27n)O(2^{7n}), which is less than O(2(1/2+o(1))n2)O(2^{(1/2+o(1))n^{2}}) but still costly. However, as mentioned above, for certain nn-qubit unitary families the explicit expressions of the unitary-stabilizer nullity can be easily computed. We believe this property will make it useful in the theoretical study. At the technical level, most of the proofs are based on connecting s(U)s(U) to certain subgroups and applying the Lagrangian theorem [37] and its variants, deriving relationships between the sizes of subgroups. Based on the group theory, an interesting fact is that unlike other magic monotones, both the state- and the unitary-stabilizer nullity are always integers. Note that since the unitary-stabilizer nullity is bounded by 2n2n, the lower bound for TT count it gives is up to 2n2n.

Finally, we connect the unitary-stabilizer nullity and the state-stabilizer nullity with auxiliary systems. Specifically, we prove that v(U)v(U) equals the best bound one can obtain by the state-stabilizer nullity with auxiliary systems. That is for any dd-qubit auxiliary system, for any (d+n)(d+n)-qubit stabilizer state |ψ|\psi\rangle, we have v(U)vs((I2dU|ψ)v(U)\geq v_{s}((I_{2^{d}}\otimes U|\psi\rangle). In particular, we show that v(U)=vs((I2nU)|ϕ)v(U)=v_{s}((I_{2^{n}}\otimes U)|\phi\rangle) where |ϕ|\phi\rangle is the 2n2n-qubit maximally entangled state. We also give numerical results that compare the TT count lower bound obtained by the state-stabilizer nullity, the unitary-stabilizer nullity, and the stabilizer extent. The results show that for several gates of practical interest, the unitary-stabilizer unitary can derive a better TT count lower bound.

Outline and main contributions

The outline and main contribution of this paper can be summarized as follows:

  • In section II, we set the notations and introduce the unitary-stabilizer nullity.

  • In section III, we prove basic properties of the unitary-stabilizer nullity, that is faithfulness, invariance under Clifford operations, and additivity under tensor product. Moreover, we show that the unitary-stabilizer nullity satisfies the subadditivity under composition.

  • In section IV, we show that the unitary-stabilizer nullity can derive a lower bound of the TT count for a given unitary UU. The method can be generalized to give lower bounds for arbitrary non-Clifford gates beyond the TT gate. We also compare the lower bound obtained by the unitary- and the state-stabilizer nullity. We prove that the bound by the unitary-stabilizer nullity will never perform worse than the state-stabilizer nullity. Moreover, we show that there is an explicit nn-qubit unitary family of unitary-stabilizer nullity 2n2n, illustrating our bounds can strictly outperform the state-stabilizer nullity. Besides, for diagonal unitaries, we prove our bounds coincide with the state-stabilizer nullity. Then connect the unitary-stabilizer nullity with the state-stabilizer nullity with auxiliary systems.

  • In section IV.5, we investigate the TT counts of a plethora of quantum gates of practical interest and the improvements are evident for many cases.

  • Finally in Section V, we conclude and discuss open problems.

II Preliminaries

II.1 Notations

In this part, we set the notations. We use ii as the imaginary number. We use \mathbb{C} as the complex field, +\mathbb{N}^{+} as the positive integers. For a complex number λ\lambda, we use λ\lambda^{\dagger} as its conjugate, |λ||\lambda| as its norm. Let A,BA,B be two subgroups of a group GG, we use A×BA\times B as the set A×B:={ab|aA,bB}A\times B\mathrel{\mathop{\mathchar 58\relax}}=\{ab|a\in A,b\in B\}. Note that when GG is Abelian, A×BA\times B is a subgroup of GG. For any nn-bit string x{0,1}n{{\textbf{x}}}\in\{0,1\}^{n}, we use (x)\odot({{\textbf{x}}}) as the product x1x2xnx_{1}x_{2}...x_{n}. We use 1n1^{n} as the nn-bit string of all ones. Given two strings x,y{0,1}n,{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}^{n}, we use xy{{\textbf{x}}}\cdot{{\textbf{y}}} as their inner product, xy{{\textbf{x}}}\oplus{{\textbf{y}}} as their bitwise XOR operation. We abbreviate x{1,+1}x\in\{-1,+1\} as x=±1x=\pm 1. We use IdI_{d} as the dd-dimensional identity matrix. We abbreviate IdI_{d} as II when dd is clear in the context. We write the state 12(|0+|1)\frac{1}{\sqrt{2}}(|0\rangle+|1\rangle) as |+|+\rangle. For a matrix UU, we use UTU^{T} as its transpose, UU^{\dagger} as its transpose conjugate. For integers dd and kk, we use d|kd|k to denote dd is a factor of kk. We use the δuv\delta_{{{\textbf{u}}}{{\textbf{v}}}} to denote the function that equals 11 if and only if u=v{{\textbf{u}}}={{\textbf{v}}} and 0 otherwise.

We use [n][n] as the set {0,1,,n1}.\{0,1,...,n-1\}. We use 4[n]4^{[n]} as the set of length-nn vectors with alphabet [4][4], i.e. 4[n]:={0,1,2,3}n4^{[n]}\mathrel{\mathop{\mathchar 58\relax}}=\{0,1,2,3\}^{n}. We use the lowercase character to denote a number, and the bold lowercase character to denote a vector, that is u[4]u\in[4] and u[4]n{{\textbf{u}}}\in[4]^{n} respectively. We represent Pauli operators as

σ0=I2=[1001],σ1=X=[0110],\displaystyle\sigma_{0}=I_{2}=\begin{bmatrix}1&0\\ 0&1\end{bmatrix},\sigma_{1}=X=\begin{bmatrix}0&1\\ 1&0\end{bmatrix}, (1)
σ2=Y=[0ii0],σ3=Z=[1001].\displaystyle\sigma_{2}=Y=\begin{bmatrix}0&-i\\ i&0\end{bmatrix},\sigma_{3}=Z=\begin{bmatrix}1&0\\ 0&-1\end{bmatrix}. (2)

We write phase gate (S), TT gate (T) and Hadamard gate (H) as

S=[100i],T=[100eπi4],H=12[1111].\displaystyle S=\begin{bmatrix}1&0\\ 0&i\end{bmatrix},T=\begin{bmatrix}1&0\\ 0&e^{\frac{\pi i}{4}}\end{bmatrix},H=\frac{1}{\sqrt{2}}\begin{bmatrix}1&1\\ 1&-1\end{bmatrix}. (3)

The CNOT gate is the two-qubit control NOT gate, i.e. CNOT|x1|x2=|x1|x1x2CNOT|x_{1}\rangle|x_{2}\rangle=|x_{1}\rangle|x_{1}\oplus x_{2}\rangle. The Cn1ZC^{n-1}Z is the multicontrolled Z gate, i.e. Cn1Z|x1xn=(1)x1x2xn|x1xnC^{n-1}Z|x_{1}...x_{n}\rangle=(-1)^{x_{1}x_{2}\cdots x_{n}}|x_{1}...x_{n}\rangle.

The one-qubit Pauli group is the group generated by the Pauli operators, i.e. 𝒫^1={ikσh|k,h[4]}\hat{{\cal P}}_{1}=\{i^{k}\sigma_{h}|k,h\in[4]\}. The nn-qubit Pauli group is defined to be the nn-fold tensor product of the one-qubit Pauli-group, i.e. 𝒫^n:=𝒫^1n\hat{{\cal P}}_{n}\mathrel{\mathop{\mathchar 58\relax}}=\hat{{\cal P}}_{1}^{\otimes n}. In the following, we are uninterested in the overall phases. We define the one-qubit Pauli-group modulo phases to be 𝒫1=𝒫^1/{ikI2|k[4]}={σh|h[4]}{\cal P}_{1}=\hat{{\cal P}}_{1}/\{i^{k}I_{2}|k\in[4]\}=\{\sigma_{h}|h\in[4]\}, where the last equality means we use σh\sigma_{h} as a representative of elements in {ikσh|k[4]}\{i^{k}\sigma_{h}|k\in[4]\}. Similarly, we define the nn-qubit Pauli group modulo phases, or the nn-qubit quotient Pauli group, to be 𝒫n:=𝒫1n={σh|h[4]}n{\cal P}_{n}\mathrel{\mathop{\mathchar 58\relax}}={\cal P}_{1}^{\otimes n}=\{\sigma_{h}|h\in[4]\}^{\otimes n}. Notice that 𝒫n{\cal P}_{n} is an Abelian group with respect to matrix-product modulo phases. We have |𝒫1|=4|{\cal P}_{1}|=4 and |𝒫n|=4n|{\cal P}_{n}|=4^{n} where |||\cdot| is the size of the group. Elements in 𝒫n{\cal P}_{n} can be represented by a string u4[n]{{\textbf{u}}}\in 4^{[n]}, that is σu=k=1nσuk\sigma_{{\textbf{u}}}=\otimes_{k=1}^{n}\sigma_{{{\textbf{u}}}_{k}}.

The nn-qubit Clifford group 𝒞n{\cal C}_{n} is the set of nn-qubit unitary UU which maps Pauli group to Pauli group, i.e. U𝒫^nU=𝒫^nU\hat{{\cal P}}_{n}U^{\dagger}=\hat{{\cal P}}_{n}. An nn-qubit state |ψ|\psi\rangle is called a stabilizer state if |ψ=U|0n|\psi\rangle=U|0\rangle^{\otimes n} for some U𝒞nU\in{\cal C}_{n}. One can verify that U𝒫nU=𝒫nU{\cal P}_{n}U^{\dagger}={\cal P}_{n}, with respect to matrix-product modulo phases. In the following context, when involving the quotient group 𝒫n{\cal P}_{n}, we always assume the operation is matrix product modulo phases.

For a given unitary UU, we define its TT count t(U)t(U) as the minimum number of TT gates required when decomposing UU into a sequence of gates over Clifford plus TT gate set, without auxiliary qubits and measurement. Here we view Clifford gates, or Clifford unitaries, as free resources. More clarifications can be seen in the quantum computation with unitary part in Appendix B. If UU cannot be exactly synthesized by Clifford plus TT gate set, we denote its TT count as ++\infty.

II.2 Pauli function

In this part, we introduce the state Pauli function and the unitary Pauli function, which are the basic elements for defining the state- and the unitary-stabilizer nullity. Those functions are related to the values of the matrix representation of a linear map in the Pauli basis. We adopt the terminology from Ref. [25].

We name them the Pauli functions in order to unify the framework of the state- and the unitary-stabilizer nullity, and to better analyze their properties.

Definition 1 (State Pauli function)

For u4[n]{{\textbf{u}}}\in 4^{[n]}, the Pauli function of an nn-qubit pure state |ψ|\psi\rangle is defined as P|ψ(u):=Tr|ψψ|σuP_{|\psi\rangle}({{\textbf{u}}})\mathrel{\mathop{\mathchar 58\relax}}=\operatorname{Tr}|\psi\rangle\!\langle\psi|\sigma_{{{\textbf{u}}}}.

Definition 2 (Unitary Pauli function)

For u,v4[n]{{\textbf{u}}},{{\textbf{v}}}\in 4^{[n]}, the Pauli function of an nn-qubit unitary UU is defined as PU(u|v)=Tr(σuUσvU)/2nP_{U}({{\textbf{u}}}|{{\textbf{v}}})=\operatorname{Tr}(\sigma_{{{\textbf{u}}}}U\sigma_{{{\textbf{v}}}}U^{\dagger})/2^{n}.

The matrix whose (u,v)({{\textbf{u}}},{{\textbf{v}}}) th element is PU(u|v)P_{U}({{\textbf{u}}}|{{\textbf{v}}}) is known to be the Pauli transfer matrix of UU. The Pauli transfer matrix is the matrix representation of the linear map U()UU(\cdot)U^{\dagger} in the Pauli basis. Equivalently, one can show that UσvU=uPU(u|v)σu.U\sigma_{{\textbf{v}}}U^{\dagger}=\sum_{{\textbf{u}}}P_{U}({{\textbf{u}}}|{{\textbf{v}}})\sigma_{{\textbf{u}}}. A useful fact is that the Pauli transfer matrix of unitary is orthogonal. That is for unitary UU,

uPU(u|v)PU(u|v)=Tr(UσvUUσvU)/2n=δvv.\displaystyle\!\!\!\sum_{{{\textbf{u}}}}P_{U}({{\textbf{u}}}|{{\textbf{v}}})P_{U}({{\textbf{u}}}|{{\textbf{v}}}^{\prime})=\operatorname{Tr}(U\sigma_{{{\textbf{v}}}}U^{\dagger}U\sigma_{{{\textbf{v}}}^{\prime}}U^{\dagger})/2^{n}=\delta_{{{\textbf{v}}}{{\textbf{v}}}^{\prime}}. (4)

Both the state and the unitary Pauli functions take values in [1,1][-1,1]. To analyze the properties of the unitary Pauli function, we connect UU to a subgroup of the quotient Pauli group.

Definition 3 (Pauli subgroup associated with UU)

Given an nn-qubit unitary UU, we define the Pauli subgroup associated with UU to be 𝒫U:=U𝒫nU𝒫n{\cal P}_{U}\mathrel{\mathop{\mathchar 58\relax}}=U{\cal P}_{n}U^{\dagger}\cap{\cal P}_{n}.

One can verify that 𝒫U{\cal P}_{U} is a subgroup of 𝒫n{\cal P}_{n}. It is worth noting that U𝒫UUU{\cal P}_{U}U^{\dagger} may not be equal to 𝒫U{\cal P}_{U}, and 𝒫U{\cal P}_{U} is not U𝒫nUU{\cal P}_{n}U^{\dagger}.

II.3 Stabilizer nullity

In this part, we give the definitions of the state- and the unitary-stabilizer nullity.

Definition 4 (state-stabilizer nullity [25])

For an nn-qubit state |ψ|\psi\rangle, the stabilizer nullity of |ψ|\psi\rangle is defined as vs(|ψ):=nlog2s(|ψ)v_{s}(|\psi\rangle)\mathrel{\mathop{\mathchar 58\relax}}=n-\log_{2}s(|\psi\rangle), where s(|ψ)s(|\psi\rangle) is equal to the number of ±\pm1s in the Pauli function of |ψ|\psi\rangle when varying u4[n]{{\textbf{u}}}\in 4^{[n]}.

To better explore the power of stabilizer nullity in quantifying the TT count (or nonstabilizerness) of quantum dynamics, we introduce the unitary-stabilizer nullity as follows.

Definition 5 (unitary-stabilizer nullity)

For an nn-qubit unitary UU, the stabilizer nullity of UU is defined as

v(U):=2nlog2s(U).\displaystyle v(U)\mathrel{\mathop{\mathchar 58\relax}}=2n-\log_{2}s(U). (5)

where s(U)s(U) is equal to the number of ±1\pm 1s in the Pauli function of UU when varying u,v4[n]{{\textbf{u}}},{{\textbf{v}}}\in 4^{[n]}.

Note that Beverland et al [25] originally define the state-stabilizer nullity as ν(|ψ)=nlog2|Stab|ψ|\nu(|\psi\rangle)=n-\log_{2}{|\mathrm{Stab}\,|\psi\rangle}|, where Stab|ψ={P𝒫^n:P|ψ=|ψ}\mathrm{Stab}\,|\psi\rangle=\{P\in\hat{{\cal P}}_{n}\mathrel{\mathop{\mathchar 58\relax}}P|\psi\rangle=|\psi\rangle\}, that is Stab|ψ\mathrm{Stab}\,|\psi\rangle is the subgroup of the Pauli group 𝒫^n\hat{{\cal P}}_{n} for which |ψ|\psi\rangle is a +1+1 eigenvector. Definition 4 is equivalent to Beverland’s definition, that is vs(|ψ)=ν(|ψ)v_{s}(|\psi\rangle)=\nu(|\psi\rangle), as illustrated in Appendix A.

For any stabilizer state |ψ|\psi\rangle, Beverland et al [25] show that vs(|ψ)=0v_{s}(|\psi\rangle)=0. For any Clifford unitary UU, one can verify that v(U)=0v(U)=0 by Lemma 1.

III Basic properties of unitary-stabilizer nullity

III.1 Basic properties of the unitary Pauli function

We firstly analyze the basic properties of the unitary Pauli function. Roughly speaking, the unitary Pauli function measures the similarity between UU and the Clifford operators, where the similarity is quantified by how well the map U()UU(\cdot)U^{\dagger} keeps the Pauli group invariant. To see this intuition, we prove some basic properties of the unitary Pauli function.

We give short proofs to Lemma 1, Lemma 2 and Lemma 3 assuming the familiarity of quantum information. The more detailed proofs are put into Appendix C.4.

Lemma 1

If UU is an nn-qubit Clifford unitary, for u,v4[n]{{\textbf{u}}},{{\textbf{v}}}\in 4^{[n]},

  • for every u, there is only one v such that PU(u|v)=±1P_{U}({{\textbf{u}}}|{{\textbf{v}}})=\pm 1, and for all vv{{\textbf{v}}}^{\prime}\neq{{\textbf{v}}}, PU(u|v)=0P_{U}({{\textbf{u}}}|{{\textbf{v}}}^{\prime})=0.

  • for every v, there is only one u such that PU(u|v)=±1P_{U}({{\textbf{u}}}|{{\textbf{v}}})=\pm 1, and for all uu{{\textbf{u}}}^{\prime}\neq{{\textbf{u}}}, PU(u|v)=0P_{U}({{\textbf{u}}}^{\prime}|{{\textbf{v}}})=0.

Proof.  We give only proof to the first statement, the second is similar. Recall that

PU(u|v)\displaystyle P_{U}({{\textbf{u}}}|{{\textbf{v}}}) =Tr(σuUσvU)/2n,\displaystyle=\operatorname{Tr}(\sigma_{{\textbf{u}}}U\sigma_{{\textbf{v}}}U^{\dagger})/2^{n}, (6)
Tr(σuσv)\displaystyle\operatorname{Tr}(\sigma_{{\textbf{u}}}\sigma_{{\textbf{v}}}) ={2n, if u=v0, else.\displaystyle=\left\{\begin{aligned} &2^{n},&\text{ if ${{\textbf{u}}}={{\textbf{v}}}$}\\ &0,&\text{ else.}\end{aligned}\right. (7)

Since UU is Clifford, U()UU(\cdot)U^{\dagger} is an isomorphism of 𝒫^n\hat{{\cal P}}_{n}, i.e. U𝒫^nU=𝒫^nU\hat{{\cal P}}_{n}U^{\dagger}=\hat{{\cal P}}_{n}. In other words, U()UU(\cdot)U^{\dagger} is a permutation over the Pauli operators {σu}u4[n]\{\sigma_{{\textbf{u}}}\}_{{{\textbf{u}}}\in 4^{[n]}}, up to a ±1\pm 1 sign. Thus by Eq. (6) and Eq. (LABEL:eq:c01), we know PU(u|v)=±1P_{U}({{\textbf{u}}}|{{\textbf{v}}})=\pm 1 if and only if UσvU=±σuU\sigma_{{{\textbf{v}}}}U^{\dagger}=\pm\sigma_{{\textbf{u}}}. Thus the lemma holds. \blacksquare

Similar results also hold for general unitary UU beyond Clifford operators, which is proved in the following lemma.

Lemma 2

For any nn-qubit unitary UU, if there exist u,v4[n]{{\textbf{u}}},{{\textbf{v}}}\in 4^{[n]} such that PU(u|v)=±1P_{U}({{\textbf{u}}}|{{\textbf{v}}})=\pm 1. Then

UσvU=±σu,\displaystyle U\sigma_{{\textbf{v}}}U^{\dagger}=\pm\sigma_{{\textbf{u}}}, (8)
uu,PU(u|v)=0.\displaystyle\forall{{\textbf{u}}}^{\prime}\neq{{\textbf{u}}},P_{U}({{\textbf{u}}}^{\prime}|{{\textbf{v}}})=0. (9)

Proof.  Recall that from equation (4) we know the Pauli transfer matrix of UU, whose (u,v)({{\textbf{u}}},{{\textbf{v}}})-th element is PU(u,v)P_{U}({{\textbf{u}}},{{\textbf{v}}}), is orthogonal. Specially, choosing v=v{{\textbf{v}}}^{\prime}={{\textbf{v}}} in Eq. (4) we know

u|PU(u|v)|2=1.\displaystyle\sum_{{{\textbf{u}}}}|P_{U}({{\textbf{u}}}|{{\textbf{v}}})|^{2}=1. (10)

Thus if PU(u|v)=±1P_{U}({{\textbf{u}}}|{{\textbf{v}}})=\pm 1, we have uu,PU(u|v)=0.\forall{{\textbf{u}}}^{\prime}\neq{{\textbf{u}}},P_{U}({{\textbf{u}}}^{\prime}|{{\textbf{v}}})=0. Since UσvU=uPU(u|v)σu,U\sigma_{{\textbf{v}}}U^{\dagger}=\sum_{{\textbf{u}}}P_{U}({{\textbf{u}}}|{{\textbf{v}}})\sigma_{{\textbf{u}}}, thus we complete the proof.

\blacksquare

In the following, we give a simple but useful connection between the unitary-stabilizer nullity and the subgroup associated with UU. Recall that s(U)s(U) is the number of ±1\pm 1s in the unitary Pauli function, as in Definition 5. In the following lemma, we prove that s(U)s(U) equals to the size of 𝒫U{\cal P}_{U}, that is the intersection of U𝒫nUU{\cal P}_{n}U^{\dagger} and 𝒫n{\cal P}_{n}. Note that 𝒫n{\cal P}_{n} is the quotient Pauli group rather than the Pauli group 𝒫n^\hat{{\cal P}_{n}}.

Lemma 3

For any nn-qubit unitary UU, 𝒫U{\cal P}_{U} is a subgroup of 𝒫n{\cal P}_{n}. What is more,

|𝒫U|=|𝒫U|=s(U).\displaystyle|{\cal P}_{U}|=|{\cal P}_{U^{\dagger}}|=s(U). (11)

Proof.  By Lemma 2, we know every ±1\pm 1 in the Pauli transfer matrix will imply a distinct element in 𝒫U{\cal P}_{U}. That is if PU(uv|v)=±1P_{U}({{\textbf{u}}}_{{{\textbf{v}}}}|{{\textbf{v}}})=\pm 1, then σuv𝒫U\sigma_{{{\textbf{u}}}_{{{\textbf{v}}}}}\in{\cal P}_{U}. Besides, for different v, such uv{{\textbf{u}}}_{{{\textbf{v}}}} are different, this comes from the fact that the Pauli transfer matrix is orthogonal, thus any row cannot have two ±1\pm 1s. Thus |𝒫U||s(U)||{\cal P}_{U}|\geq|s(U)|. On the other hand, for every element in 𝒫U{\cal P}_{U}, that is a σu\sigma_{{\textbf{u}}} such that UσvU=±σuU\sigma_{{\textbf{v}}}U^{\dagger}=\pm\sigma_{{\textbf{u}}} for some v, we would have |PU(u|v)|=1|P_{U}({{\textbf{u}}}|{{\textbf{v}}})|=1, thus |𝒫U||s(U)||{\cal P}_{U}|\leq|s(U)|. Thus |PU|=s(U)|P_{U}|=s(U). Besides, we have Tr(σuUσvU)=Tr(σvUσuU)\operatorname{Tr}(\sigma_{{\textbf{u}}}U\sigma_{{\textbf{v}}}U^{\dagger})=\operatorname{Tr}(\sigma_{{\textbf{v}}}U^{\dagger}\sigma_{{\textbf{u}}}U), thus the Pauli transfer matrix of UU^{\dagger} is the transpose of the Pauli transfer matrix of UU. Thus |𝒫U|=|s(U)|=|𝒫U||{\cal P}_{U^{\dagger}}|=|s(U)|=|{\cal P}_{U}|. \blacksquare

An interesting corollary is that the unitary-stabilizer nullity is always an integer, although we do not utilize this property in the following sections.

Corollary 4

For any nn-qubit unitary UU, v(U)v(U) must be an integer, or equivalently, s(U)=2ks(U)=2^{k} for some non-negative integer kk.

By Lemma 3, we know s(U)=|𝒫U|s(U)=|{\cal P}_{U}|. Since 𝒫U{\cal P}_{U} is a subgroup of 𝒫n{\cal P}_{n}, where |𝒫n|=4n|{\cal P}_{n}|=4^{n}, by group theory, we know that s(U)|4ns(U)|4^{n}. Thus v(U)=2nlog2s(U)v(U)=2n-\log_{2}s(U) is always an integer.

III.2 Basic properties of the unitary-stabilizer nullity

In this section, we prove the basic properties of the unitary-stabilizer nullity, that is the faithfulness, invariance under Clifford operations, additivity under tensor product and subadditivity under composition.

Theorem 5 (Faithfulness)

Let UU be an nn-qubit unitary. Then v(U)0v(U)\geq 0. Furthermore, v(U)=0v(U)=0 if and only if UU is a Clifford unitary. Or equivalently, s(U)4ns(U)\leq 4^{n} and s(U)=4ns(U)=4^{n} if and only if UU is a Clifford unitary.

Proof.  For any nn-qubit unitary UU, we firstly prove v(U)0v(U)\geq 0, or equivalently, s(U)4ns(U)\leq 4^{n}. By Lemma 2, if there exist u,v{{\textbf{u}}},{{\textbf{v}}} such that PU(u|v)=1P_{U}({{\textbf{u}}}|{{\textbf{v}}})=1 then for any uu,𝒫U(u|v)=0{{\textbf{u}}}^{\prime}\neq{{\textbf{u}}},{\cal P}_{U}({{\textbf{u}}}^{\prime}|{{\textbf{v}}})=0. Thus every ±1\pm 1 in the Pauli function must come together with 4n14^{n}-1 number of 0s0s. Thus s(U)4ns(U)\leq 4^{n}.

If s(U)=4ns(U)=4^{n}, by Lemma 2, for every Pauli operator σv\sigma_{{\textbf{v}}}, UσvUU\sigma_{{\textbf{v}}}U^{\dagger} is a Pauli operator, thus UU is a Clifford unitary. On the other hand, if UU is a Clifford operator, by Lemma 1, it is easy to verify s(U)=4n.s(U)=4^{n}. \blacksquare

Theorem 6 (Invariant under Clifford operation)

For an nn-qubit unitary UU and an nn-qubit Clifford unitary VV,

v(VU)=v(U),s(VU)=s(U).\displaystyle v(VU)=v(U),\quad s(VU)=s(U). (12)

Similarly, it also holds that v(UV)=v(U),s(UV)=s(U)v(UV)=v(U),s(UV)=s(U).

Proof.  Recall that the unitary Pauli function

PVU(u|v)\displaystyle P_{VU}({{\textbf{u}}}|{{\textbf{v}}}) =TrσuVUσvUV/2n\displaystyle=\operatorname{Tr}\sigma_{{\textbf{u}}}VU\sigma_{{\textbf{v}}}U^{\dagger}V^{\dagger}/2^{n}
=TrVσuVUσvU/2n.\displaystyle=\operatorname{Tr}V^{\dagger}\sigma_{{\textbf{u}}}VU\sigma_{{\textbf{v}}}U^{\dagger}/2^{n}. (13)

Since VV is Clifford, then V()VV^{\dagger}(\cdot)V is an isomorphism over the Pauli group, i.e. V𝒫^nV=𝒫^nV^{\dagger}\hat{{\cal P}}_{n}V=\hat{{\cal P}}_{n}. Further, since VσuVV^{\dagger}\sigma_{{\textbf{u}}}V does not change the eigenvalue of σu\sigma_{{\textbf{u}}}, where is ±1\pm 1. Thus

V{σu|u4[n]}V={puσu|u4[n],pu=±1}.\displaystyle V^{\dagger}\{\sigma_{{\textbf{u}}}|{{\textbf{u}}}\in 4^{[n]}\}V=\{p_{{\textbf{u}}}\sigma_{{\textbf{u}}}|{{\textbf{u}}}\in 4^{[n]},p_{{\textbf{u}}}=\pm 1\}. (14)

In other words, the set of unitary Pauli functions satisfies

{PVU(u|v)}={puPU(u|v)|pu=±1}.\displaystyle\{P_{VU}({{\textbf{u}}}|{{\textbf{v}}})\}=\{p_{{\textbf{u}}}P_{U}({{\textbf{u}}}|{{\textbf{v}}})\,|\,p_{{\textbf{u}}}=\pm 1\}. (15)

Thus s(VU)=s(U)s(VU)=s(U), thus v(VU)=s(U)v(VU)=s(U). Similarly, it holds that v(VU)=v(U),s(VU)=s(U)v(VU)=v(U),s(VU)=s(U). \blacksquare

Theorem 7 (Additivity under tensor product)

Given nn-qubit unitary UU and mm-qubit unitary VV, the following holds

v(UV)=v(U)+v(V).\displaystyle v(U\otimes V)=v(U)+v(V). (16)

Proof.  For simplicity, given u4[n+m]{{\textbf{u}}}\in 4^{[n+m]}, we write u(1){{\textbf{u}}}^{(1)} as the first nn bits of u, i.e. u(1)=u1u2un{{\textbf{u}}}^{(1)}={{\textbf{u}}}_{1}{{\textbf{u}}}_{2}...{{\textbf{u}}}_{n}. We write u(2){{\textbf{u}}}^{(2)} as the last mm bits of u, i.e. u(2)=un+1un+2un+m{{\textbf{u}}}^{(2)}={{\textbf{u}}}_{n+1}{{\textbf{u}}}_{n+2}...{{\textbf{u}}}_{n+m}. It suffices to notice that

PUV(u|v)\displaystyle P_{U\otimes V}({{\textbf{u}}}|{{\textbf{v}}})
=Trσu(1)σu(2)(UV)σv(1)σv(2)(UV)/2(n+m)\displaystyle=\operatorname{Tr}\sigma_{{{\textbf{u}}}^{(1)}}\otimes\sigma_{{{\textbf{u}}}^{(2)}}\left(U\otimes V\right)\sigma_{{{\textbf{v}}}^{(1)}}\otimes\sigma_{{{\textbf{v}}}^{(2)}}\left(U^{\dagger}\otimes V^{\dagger}\right)/2^{(n+m)}
=Trσu(1)Uσv(1)U/2nTrσu(2)Vσv(2)V/2m\displaystyle=\operatorname{Tr}\sigma_{{{\textbf{u}}}^{(1)}}U\sigma_{{{\textbf{v}}}^{(1)}}U^{\dagger}/2^{n}\cdot\operatorname{Tr}\sigma_{{{\textbf{u}}}^{(2)}}V\sigma_{{{\textbf{v}}}^{(2)}}V^{\dagger}/2^{m}
=PU(u(1)|v(1))PV(u(2)|v(2)).\displaystyle=P_{U}({{\textbf{u}}}^{(1)}|{{\textbf{v}}}^{(1)})\cdot P_{V}({{\textbf{u}}}^{(2)}|{{\textbf{v}}}^{(2)}). (17)

Thus we have s(UV)=s(U)s(V)s(U\otimes V)=s(U)s(V), 2(n+m)log2s(UV)=(2nlog2s(U))+(2mlog2s(V))2(n+m)-\log_{2}s(U\otimes V)=(2n-\log_{2}s(U))+(2m-\log_{2}s(V)), and then v(UV)=v(U)+v(V)v(U\otimes V)=v(U)+v(V). \blacksquare

Lemma 8

For any nn-qubit unitary UU and VV,

|𝒫U𝒫V|s(UV).\displaystyle|{\cal P}_{U^{\dagger}}\cap{\cal P}_{V}|\leq s(UV). (18)

Recall that

𝒫UV\displaystyle{\cal P}_{UV} =UV𝒫nVU𝒫n\displaystyle=UV{\cal P}_{n}V^{\dagger}U^{\dagger}\cap{\cal P}_{n}
=U(V𝒫nVU𝒫nU)U.\displaystyle=U\left(V{\cal P}_{n}V^{\dagger}\cap U^{\dagger}{\cal P}_{n}U\right)U^{\dagger}. (19)

Noting that U()UU(\cdot)U^{\dagger} is an injection, we have

|𝒫UV|\displaystyle|{\cal P}_{UV}| =|V𝒫nVU𝒫nU|\displaystyle=|V{\cal P}_{n}V^{\dagger}\cap U^{\dagger}{\cal P}_{n}U|
|V𝒫nVU𝒫nU𝒫n|\displaystyle\geq|V{\cal P}_{n}V^{\dagger}\cap U^{\dagger}{\cal P}_{n}U\cap{\cal P}_{n}|
|𝒫V𝒫U|.\displaystyle\geq|{\cal P}_{V}\cap{\cal P}_{U^{\dagger}}|. (20)

By Lemma 3 we know |𝒫UV|=s(UV)|{\cal P}_{UV}|=s(UV), we conclude |𝒫U𝒫V|s(UV)|{\cal P}_{U^{\dagger}}\cap{\cal P}_{V}|\leq s(UV).

The following lemma is a variation of Lagrange’s theorem in group theory. It connects the sizes between subgroups and their intersections. A detailed proof is provided in Appendix. C.4.

Lemma 9

Suppose A,BA,B are subgroups of HH, then |A×B|=|A||B||AB||A\times B|=\frac{|A|\cdot|B|}{|A\cap B|}. Since A×BHA\times B\subseteq H we have |A||B||H||AB||A|\cdot|B|\leq|H|\cdot|A\cap B|.

Utilizing Lemma 9, we finally arrive at the statement that the unitary-stabilizer nullity satisfies the subadditivity under composition.

Theorem 10 (Subadditivity under composition)

Given nn-qubit unitary UU and VV, the following holds

v(UV)v(U)+v(V).\displaystyle v(UV)\leq v(U)+v(V). (21)

Or equivalently s(U)s(V)4ns(UV)s(U)s(V)\leq 4^{n}s(UV).

Proof.  Let H=𝒫n,A=𝒫U,B=𝒫VH={\cal P}_{n},A={\cal P}_{U^{\dagger}},B={\cal P}_{V}, we know A,BA,B are subgroups of HH. By Lemma 9 we have

|𝒫U||𝒫V||𝒫U𝒫V|4n.\displaystyle|{\cal P}_{U^{\dagger}}|\cdot|{\cal P}_{V}|\leq|{\cal P}_{U^{\dagger}}\cap{\cal P}_{V}|\cdot 4^{n}. (22)

Combine with Lemma 8, we have

|𝒫U||𝒫V||s(UV)|4n.\displaystyle|{\cal P}_{U^{\dagger}}|\cdot|{\cal P}_{V}|\leq|s(UV)|\cdot 4^{n}. (23)

Further by Lemma 3, we know

|𝒫V|=s(V),|𝒫U|=|𝒫U|=s(U).\displaystyle|{\cal P}_{V}|=s(V),|{\cal P}_{U^{\dagger}}|=|{\cal P}_{U}|=s(U). (24)

thus finally we get

s(U)s(V)4ns(UV).\displaystyle s(U)s(V)\leq 4^{n}s(UV). (25)

\blacksquare

IV Applications

Interestingly, by the properties in section III, especially the subadditivity under composition, we can use v()v(\cdot) to give a lower bound for the TT count. In section IV.1, we prove this connection. In section IV.2 we compare the lower bound for TT count obtained by the state- and the unitary-stabilizer nullity, and prove that the unitary-stabilizer nullity never performs worse than the state-stabilizer nullity. We also prove that for diagonal unitary, the two lower bounds coincide. Finally, in section IV.3, we give an nn-qubit unitary family which has unitary-stabilizer nullity 2n2n. This circuit family gives an example where the lower bound obtained by the unitary-stabilizer nullity strictly outperforms the state-stabilizer nullity by a factor of 22. Furthermore, in section IV.4, we connect the unitary-stabilizer nullity and the state-stabilizer nullity with auxiliary systems, showing that adding auxiliary systems and choosing proper stabilizer states can strictly improve the lower bound obtained by the state-stabilizer nullity.

IV.1 Lower bounds for the TT count

Recall that for a given unitary UU, its TT count t(U)t(U) is the minimum number of TT gates used among any implementation of UU that decomposes UU into a sequence of gates from Clifford plus TT-gate set, without ancillary qubits and measurement. In this section, we would use the unitary-stabilizer nullity to lower bound the TT count. The main idea is to use the fact that v()v(\cdot) satisfies the subadditivity under composition.

Theorem 11 (Lower bound for the TT count)

For any nn-qubit unitary UU, its TT count is lower bounded by its unitary-stabilizer nullity, that is

t(U)v(U).\displaystyle t(U)\geq v(U). (26)

Proof.  Notice that v(T)=1v(T)=1 by verifying the unitary Pauli function of TT

{PT(u|v)}=[10000222200222200001].\displaystyle\{P_{T}({{\textbf{u}}}|{{\textbf{v}}})\}=\begin{bmatrix}1&0&0&0\\ 0&\frac{\sqrt{2}}{2}&-\frac{\sqrt{2}}{2}&0\\ 0&\frac{\sqrt{2}}{2}&\frac{\sqrt{2}}{2}&0\\ 0&0&0&1\end{bmatrix}. (27)

Thus, we have v(T)=2log22=1v(T)=2-\log_{2}2=1.

Besides, since I2I_{2} is a Clifford operator, by the faithfulness, i.e. Theorem 5, we have

v(I2)=0.\displaystyle v(I_{2})=0. (28)

By the additivity under tensor product, i.e. Theorem 7, we have

v(TI2I2I2)=v(T)+0=1.\displaystyle v(T\otimes I_{2}\otimes I_{2}...\otimes I_{2})=v(T)+0=1. (29)

From Theorem 5 we know that all the Clifford unitaries have unitary-stabilizer nullity 0. Suppose UU is decomposed into sequence of Clifford gates and t(U)t(U) TT gates, using the subadditivity of composition sequentially, that is Theorem 10, we have

v(U)\displaystyle v(U) 0+t(U)v(T)\displaystyle\leq 0+t(U)\cdot v(T) (30)
=t(U).\displaystyle=t(U).

This concludes the theorem. \blacksquare

To compare the lower bound got by the unitary-stabilizer nullity and the lower bound by the state-stabilizer nullity, we would use the following theorem from [25].

Theorem 12

[25] For any nn-qubit unitary UU, for any (n+t)(n+t)-qubit stabilizer state |ϕ|\phi\rangle, the TT count of UU is lower bounded by

t(U)vs((I2tU)|ϕ).\displaystyle t(U)\geq v_{s}\left((I_{2^{t}}\otimes U)|\phi\rangle\right). (31)

To generalize our result a little bit, consider the task that decomposing UU into sequence of Clifford gates and WW gates, where WW gate is a gate beyond Clifford. Using similar proofs as in the Theorem 11, we can also use the unitary-stabilizer nullity to lower bound the number of WW gates.

Theorem 13 (Lower bound for gate synthesis)

For any nn-qubit unitary UU, the number of gates WW required to realize UU under Clifford unitaries, denoted as w(U)w(U), satisfies

w(U)v(U)v(W).\displaystyle{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}w(U)}\geq\frac{v(U)}{v(W)}. (32)

Proof.  Note that the subadditivity under composition (Theorem 10) holds for arbitrary two unitaries. Suppose UU is decomposed into Clifford and WW gates, using Theorem 10 we get similar results as Eq. (29) and Eq. (30), i.e. v(WI2I2I2)=v(W)v(W\otimes I_{2}\otimes I_{2}...\otimes I_{2})=v(W) and v(U)w(U)v(W)v(U)\leq w(U)v(W). Thus we prove the theorem. \blacksquare

IV.2 Comparison between the unitary and state-stabilizer nullity

Beverland et al. [25] showed that for a unitary UU, for any stabilizer state |ψ|\psi\rangle, the state-stabilizer nullity can derive a lower bound for TT count for UU as vs(U|ψ)v_{s}(U|\psi\rangle). We compare this lower bound with the lower bound obtained by the unitary-stabilizer nullity from Theorem 11. We prove in Theorem 14 that our method will never perform worse and provide the detailed proof in Appendix D. We further show that our bound can perform strictly better for a certain unitary family by a factor of 2, as in Corollary 21. We also prove that for diagonal unitary, the lower bounds for TT count obtained by the state- and the unitary-stabilizer nullity are the same, as in Theorem 18.

Theorem 14

For any nn-qubit unitary UU, for any nn-qubit stabilizer state |ψ|\psi\rangle,

v(U)vs(U|ψ).\displaystyle v(U)\geq v_{s}(U|\psi\rangle). (33)

That is

v(U)maxC𝒞nvs(UC|0n).\displaystyle v(U)\geq\max_{C\in{\cal C}_{n}}v_{s}(UC|0\rangle^{\otimes n}). (34)

The equality holds if and only if there exists Clifford C𝒞nC\in{\cal C}_{n} such that

{u4[n]|0n|CUσuUC|0n=±1}\displaystyle\left\{{{\textbf{u}}}\in 4^{[n]}\,|\,\langle 0^{\otimes n}|C^{\dagger}U^{\dagger}\sigma_{{\textbf{u}}}UC|0\rangle^{\otimes n}=\pm 1\right\}
={u4[n]|CUσuUC±{I,Z}n}.\displaystyle=\left\{{{\textbf{u}}}\in 4^{[n]}\,|\,C^{\dagger}U^{\dagger}\sigma_{{\textbf{u}}}UC\in\pm\{I,Z\}^{\otimes n}\right\}. (35)

and

C{I,Z}nC×(U𝒫nU𝒫n)=𝒫n.\displaystyle C\{I,Z\}^{\otimes n}C^{\dagger}\times\left(U^{\dagger}{\cal P}_{n}U\cap{\cal P}_{n}\right)={\cal P}_{n}. (36)

The equality in equation (36) is with respect to matrix-product modulo phases.

In the following, we prove several lemmas which will be used to show the unitary-stabilizer nullity and the state-stabilizer nullity coincide for the diagonal unitary.

Lemma 15

For any u4[n]{{\textbf{u}}}\in 4^{[n]},

x,y{0,1}n|xσuy||xy|{I,X}n.\displaystyle\sum_{{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}^{n}}|\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}\rangle|\,|{{\textbf{x}}}\rangle\langle{{\textbf{y}}}|\in\{I,X\}^{\otimes n}. (37)

Here |xσuy||\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}\rangle| is the norm of the complex number x|σu|y\langle{{\textbf{x}}}|\sigma_{{\textbf{u}}}|{{\textbf{y}}}\rangle.

Proof.  Firstly we prove this lemma for n=1n=1. We can directly verify that

x,y{0,1}|xσy||xy|={I,If σ{I,Z}X,If σ{X,Y}\displaystyle\sum_{{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}}|\langle{{\textbf{x}}}\sigma{{\textbf{y}}}\rangle|\,|{{\textbf{x}}}\rangle\langle{{\textbf{y}}}|=\left\{\begin{aligned} &I,\quad\text{If }\sigma\in\{I,Z\}\\ &X,\quad\text{If }\sigma\in\{X,Y\}\\ \end{aligned}\right. (38)

It is worth noting that x,y{0,1}|xσy||xy|\sum_{{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}}|\langle{{\textbf{x}}}\sigma{{\textbf{y}}}\rangle|\,|{{\textbf{x}}}\rangle\langle{{\textbf{y}}}| records the positions of the nonzero values in σ\sigma. From this point of view, we can generalize this proof to all nn. \blacksquare

By Theorem 14 we prove that the lower bound for the TT-count obtained by the unitary-stabilizer nullity will never perform worse than the bound obtained by the state-stabilizer nullity. In particular, we prove that for the diagonal unitary, the two bounds coincide.

We would use Eq. (14) and set C=HnC=H^{\otimes n}. Before showing the main result of this section, we would like to present the following lemmas with detailed proof in Appendix D.

Lemma 16

For any diagonal nn-qubit unitary UU,

{u4[n]|+|nUσuU|+n=±1}\displaystyle\{{{\textbf{u}}}\in 4^{[n]}\,|\,\langle+|^{\otimes n}U^{\dagger}\sigma_{{\textbf{u}}}U|+\rangle^{\otimes n}=\pm 1\}
={u4[n]|UσuU±{I,X}n}.\displaystyle=\{{{\textbf{u}}}\in 4^{[n]}\,|\,U^{\dagger}\sigma_{{\textbf{u}}}U\in\pm\{I,X\}^{\otimes n}\}. (39)
Lemma 17

For any diagonal unitary UU,

{I,X}n×(U𝒫nU𝒫n)=𝒫n.\displaystyle\{I,X\}^{\otimes n}\times\left(U^{\dagger}{\cal P}_{n}U\cap{\cal P}_{n}\right)={\cal P}_{n}. (40)

w.r.t product modulo phases.

Now we are ready to establish the result for diagonal unitary as follows.

Theorem 18

For any diagonal nn-qubit unitary UU,

v(U)=maxC𝒞nvs(UC|0n).\displaystyle v(U)=\max_{C\in{\cal C}_{n}}v_{s}(UC|0\rangle^{\otimes n}). (41)

Proof.  By setting C=HnC=H^{\otimes n} in Theorem 14 and then using Lemma 16, Lemma 17 to show that equations (14)(36)(\ref{eq:diageqc})(\ref{eq:noless}) are satisfied. \blacksquare

Corollary 19

For any diagonal nn-qubit unitary UU, the maximum of its corresponding state-stabilizer nullity, i.e. vs(U|ψ)v_{s}(U|\psi\rangle) over all nn-qubit stabilizer state |ψ|\psi\rangle, is achieved by |ψ=|+n.|\psi\rangle=|+\rangle^{\otimes n}.

IV.3 Explicit expressions of the unitary-stabilizer nullity for unitary families

As proved in the previous sections, for the task of lower bounding the TT count for a given unitary, the bound obtained by the unitary-stabilizer nullity is always no less than the bound obtained by the state-stabilizer nullity. In this section, we give an example where the unitary-stabilizer nullity performs strictly better.

In Proposition 4.1 of Ref. [25] it showed that for n3n\geq 3, the state-stabilizer nullity of the states related to the nn-bit controlled ZZ gate is nn, i.e. vs(Cn1Z|+n)=nv_{s}(C^{n-1}Z|+\rangle^{\otimes n})=n . This implies the TT count for Cn1ZC^{n-1}Z is at least nn. One may wonder whether the unitary-stabilizer nullity can improve these bounds. Unfortunately, since Cn1ZC^{n-1}Z is diagonal, by Theorem 18 we know its unitary-stabilizer nullity is also nn. To find examples to differentiate the state- and unitary-stabilizer nullity, we should focus on nondiagonal unitaries.

In the following, we give a family of nn-qubit unitary, which has unitary-stabilizer nullity 2n2n. This implies constructing such a unitary needs at least 2n2n TT gates by Theorem 11. This unitary, or its corresponding quantum circuit, consists of an nn-qubit controlled ZZ gate, a layer of Hadamard gate, and an nn-qubit controlled ZZ gate.

Theorem 20

For n3n\geq 3,

v(Cn1ZHnCn1Z)=2n.\displaystyle v(C^{n-1}Z\,H^{\otimes n}\,C^{n-1}Z)=2n. (42)

Theorem 20 implies that the lower bound given by the unitary-stabilizer nullity can outperform the bound given by the stabilizer nullity. The detailed proof is provided in Appendix C. We further have the following corollary.

Corollary 21

There exists a family of nn-qubit unitary UnU_{n} such that

v(Un)>maxC𝒞nvs(UnC|0n)).\displaystyle v(U_{n})>\max_{C\in{\cal C}_{n}}v_{s}(U_{n}C|0\rangle^{\otimes n})). (43)

IV.4 Adding auxiliary systems improves the state-stabilizer nullity

In this section, we reconnect the state-stabilizer nullity and the unitary-stabilizer nullity. For any nn-qubit unitary UU, instead of comparing vs(U|ψ)v_{s}(U|\psi\rangle) and v(U)v(U) for some nn-qubit stabilizer state |ψ|\psi\rangle as in section IV.2, we add auxiliary systems and carefully choose the stabilizer states in the computation of vs()v_{s}(\cdot), i.e. vs(I2dU|ϕ)v_{s}(I_{2^{d}}\otimes U|\phi\rangle) for some dd-qubit auxiliary system and (d+n)(d+n)-qubit stabilizer state |ϕ|\phi\rangle. By Theorem 22 and Corollary 23, we prove that adding auxiliary systems can strictly increase the value of the state-stabilizer nullity, thus strictly improving the lower bound for the TT count through Theorem 12. Furthermore, we prove that the unitary-stabilizer nullity v(U)v(U) equals the best lower bound for the TT count which one can get through this method, i.e. computing the state-stabilizer nullity with auxiliary systems, as in Theorem 24.

It is worth noting that although the state-stabilizer nullity vs()v_{s}(\cdot) can derive a lower bound for the TT count, it is not the TT count itself, and it holds different properties to the TT count. Thus although the TT count for I2dUI_{2^{d}}\otimes U is no more than the TT count for UU, the state-stabilizer nullity for I2dUI_{2^{d}}\otimes U (with respect to some stabilizer state) can be strictly larger than the state-stabilizer nullity for UU, as in Corollary 23.

Theorem 22

Let |Φ|\Phi\rangle be the maximally entangled states on 2n2n-qubit systems, i.e. |Φ=x{0,1}n12n|x|x|\Phi\rangle=\sum_{{{\textbf{x}}}\in\{0,1\}^{n}}\frac{1}{\sqrt{2^{n}}}|{{\textbf{x}}}\rangle|{{\textbf{x}}}\rangle. For any nn-qubit unitary UU,

v(U)=vs(I2nU|Φ).\displaystyle v(U)=v_{s}\left(I_{2^{n}}\otimes U|\Phi\rangle\right). (44)
Corollary 23

For the 2n2n-qubit maximally entangled states |Φ=x{0,1}n12n|x|x|\Phi\rangle=\sum_{{{\textbf{x}}}\in\{0,1\}^{n}}\frac{1}{\sqrt{2^{n}}}|{{\textbf{x}}}\rangle|{{\textbf{x}}}\rangle, for any nn-qubit unitary UU,

vs(I2nU|Φ)maxC𝒞nvs(UC|0n).\displaystyle v_{s}(I_{2^{n}}\otimes U|\Phi\rangle)\geq\max_{C\in{\cal C}_{n}}v_{s}(UC|0\rangle^{\otimes n}). (45)

The above inequality can be strict, that is

U such that vs(I2nU|Φ)>maxC𝒞nvs(UC|0n).\displaystyle\exists U\text{ such that }v_{s}(I_{2^{n}}\otimes U|\Phi\rangle)>\max_{C\in{\cal C}_{n}}v_{s}(UC|0\rangle^{\otimes n}). (46)

Before giving proofs to the theorem and the corollary, we want to make several clarifications:

  • One can easily check the following argument holds as in Appendix C.3

    maxC𝒞2nvs(I2nU)C|02n)maxC𝒞nvs(UC|0n).\displaystyle\max_{C^{\prime}\in{\cal C}_{2n}}v_{s}\left(I_{2^{n}}\otimes U\right)C^{\prime}|0^{\otimes 2n}\rangle)\geq\max_{C\in{\cal C}_{n}}v_{s}(UC|0\rangle^{\otimes n}). (47)

    That is when adding the auxiliary systems, the best lower bounds for the TT count we can get by the state-stabilizer nullity, is never less than the case when there is no auxiliary system. Compared to this easy argument, Theorem 22 has two key differences. The first one is finding the best bound in the left side of Eq. (47), or equivalent enumerating all possible 2n2n-qubit stabilizer state C|02nC^{\prime}|0^{2n}\rangle, need at least O(2Ω(n2))O(2^{\Omega(n^{2})}) time and thus is hard. In contrast, in Corollary 23 and Theorem 24 we show that to find the best bound, it suffices to set the 2n2n-qubit stabilizer state to be a fixed state, that is the maximally entangled state. The second difference is Eq. (47) shows only \geq, Corollary 23 shows that this inequality can be strict >> in some cases. Combined with Theorem 12, this implies analyzing the state-stabilizer nullity with auxiliary systems can strictly improve the lower bound for the TT count.

  • One may expect whether using higher-dimensional auxiliary systems or choosing other stabilizer states other than the maximally entangled states can further improve the bound by the state-stabilizer nullity. The answer is no. See as in Theorem 24.

  • Since in general the state-stabilizer nullity does not satisfy the subadditivity under composition (counterexample can be seen as in Appendix C.2), thus using properties of the state-stabilizer nullity and Theorem 22 is not sufficient to imply the subadditivity under composition of the unitary-stabilizer nullity, that is Theorem 10.

  • Corollary 23 and equation (46) do not violate the property that the state-stabilizer nullity is invariant under Clifford unitaries, i.e. for any unitary VV, one has vs(CV|ψ)=vs(V|ψ)v_{s}(CV|\psi\rangle)=v_{s}(V|\psi\rangle) for any Clifford unitary CC and stabilizer state |ψ|\psi\rangle. More clarifications can be seen at Appendix C.

Proof. [of Theorem 22] By the symmetry of the maximally entangled states, for any nn-qubit operator M2n2nM\in\mathbb{C}^{2^{n}\otimes 2^{n}}, we have (MI2n)|Φ=(I2nMT)|Φ(M\otimes I_{2^{n}})|\Phi\rangle=(I_{2^{n}}\otimes M^{T})|\Phi\rangle, which is known as the transpose trick.

For simplicity, in the following we abbreviate I2nI_{2^{n}} as II. Suppose the 2n2n-qubit maximally entangled state |Φ|\Phi\rangle is on systems ABA\otimes B, where both A,BA,B are nn-qubit systems. Then for any 2n2n-qubit Pauli operator σuσv\sigma_{{\textbf{u}}}\otimes\sigma_{{\textbf{v}}}, where u,v4[n]{{\textbf{u}}},{{\textbf{v}}}\in 4^{[n]}, the corresponding state Pauli function satisfies

P(IU)|Φ(uv)\displaystyle P_{(I\otimes U)|\Phi\rangle}({{\textbf{u}}}{{\textbf{v}}})
=TrAB(σuσv(IU)|ΦΦ|(IU))\displaystyle=\operatorname{Tr}_{AB}\left(\sigma_{{\textbf{u}}}\otimes\sigma_{{\textbf{v}}}\,(I\otimes U)\,|\Phi\rangle\langle\Phi|\,(I\otimes U^{\dagger})\right)
=TrAB((IU)σuσv(IU)|ΦΦ|)\displaystyle=\operatorname{Tr}_{AB}\left((I\otimes U^{\dagger})\,\sigma_{{\textbf{u}}}\otimes\sigma_{{\textbf{v}}}\,(I\otimes U)\,|\Phi\rangle\langle\Phi|\right)
=TrAB((IUσvU)(σuI)|ΦΦ|)\displaystyle=\operatorname{Tr}_{AB}\left(\,(I\otimes U^{\dagger}\sigma_{{\textbf{v}}}U)\,(\sigma_{{\textbf{u}}}\otimes I)\,|\Phi\rangle\langle\Phi|\right)
=TrAB((IUσvUσuT)|ΦΦ|)\displaystyle=\operatorname{Tr}_{AB}\left(\,(I\otimes U^{\dagger}\sigma_{{\textbf{v}}}U\sigma_{{{\textbf{u}}}}^{T})|\Phi\rangle\langle\Phi|\right)
=γTrAB((IUσvUσu)|ΦΦ|).\displaystyle=\gamma\operatorname{Tr}_{AB}\left(\,(I\otimes U^{\dagger}\sigma_{{\textbf{v}}}U\sigma_{{{\textbf{u}}}})|\Phi\rangle\langle\Phi|\right). (48)

The last two are obtained by the transpose trick and the fact that γ{1,+1})\gamma\in\{-1,+1\}). Further, notice that

|ΦΦ|\displaystyle|\Phi\rangle\langle\Phi| =12nx,y{0,1}n|x|xy|y|.\displaystyle=\frac{1}{2^{n}}\sum_{{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}^{n}}|{{\textbf{x}}}\rangle|{{\textbf{x}}}\rangle\langle{{\textbf{y}}}|\langle{{\textbf{y}}}|. (49)

The above equations can be further simplified by

P(IU)|Φ(uv)\displaystyle P_{(I\otimes U)|\Phi\rangle}({{\textbf{u}}}{{\textbf{v}}})
=γTrAB((IUσvUσu)12nx,y{0,1}n|x|xy|y|)\displaystyle=\gamma\operatorname{Tr}_{AB}\left(\,(I\otimes U^{\dagger}\sigma_{{\textbf{v}}}U\sigma_{{{\textbf{u}}}})\frac{1}{2^{n}}\sum_{{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}^{n}}|{{\textbf{x}}}\rangle|{{\textbf{x}}}\rangle\langle{{\textbf{y}}}|\langle{{\textbf{y}}}|\right)
=γTrB(UσvUσu12nx{0,1}n|xx|)\displaystyle=\gamma\operatorname{Tr}_{B}\left(\,U^{\dagger}\sigma_{{\textbf{v}}}U\sigma_{{{\textbf{u}}}}\frac{1}{2^{n}}\sum_{{{\textbf{x}}}\in\{0,1\}^{n}}|{{\textbf{x}}}\rangle\langle{{\textbf{x}}}|\right)
=γTrB(UσvUσu12nI)\displaystyle=\gamma\operatorname{Tr}_{B}\left(\,U^{\dagger}\sigma_{{\textbf{v}}}U\sigma_{{{\textbf{u}}}}\frac{1}{2^{n}}I\right)
=γ2nTrB(σvUσuU)\displaystyle=\frac{\gamma}{2^{n}}\operatorname{Tr}_{B}\left(\sigma_{{\textbf{v}}}U\sigma_{{{\textbf{u}}}}U^{\dagger}\right)
=γPU(v|u).\displaystyle={\gamma}\,P_{U}({{\textbf{v}}}|{{\textbf{u}}}). (50)

In other words, if we ignore the γ=±1\gamma=\pm 1 factor, then there is a one-to-one correspondence between the values of state Pauli functions and the unitary Pauli function. Thus the two Pauli functions have the same number of ±1\pm 1s, that is s(U)=s((IU)|Φ)s(U)=s((I\otimes U)|\Phi\rangle). Since |Φ|\Phi\rangle is a 2n2n-qubit state, we obtain 2ns(U)=2ns((IU)|Φ)2n-s(U)=2n-s((I\otimes U)|\Phi\rangle). Then we have v(U)=vs((IU)|Φ)v(U)=v_{s}\left((I\otimes U)|\Phi\rangle\right). \blacksquare

Moreover, the proof of Corollary 23 can be proved by combining Theorem 22, Theorem 14 and Corollary 21.

In the following, we find that v(U)v(U) equals the best lower bound for the TT count that can be obtained from the state-stabilizer nullity with auxiliary systems. In the following theorem the maximization is over dd-qubit auxiliary systems for any positive integer dd. The detailed proof is provided in Appendix D.

Theorem 24

For any nn-qubit unitary UU, we have

v(U)=maxd+maxC𝒞d+nvs((I2dU)C|0(d+n)).\displaystyle v(U)=\max_{d\in\mathbb{N}^{+}}\max_{C\in{\cal C}_{d+n}}v_{s}\left((I_{2^{d}}\otimes U)\,C|0\rangle^{\otimes(d+n)}\right). (51)

IV.5 Estimating the TT count for quantum gates of interests

In this section, we showcase the specific applications of unitary-stabilizer nullity in gate synthesis. We present numerical results of using the state-stabilizer nullity, the unitary-stabilizer nullity, and the stabilizer extent [7] to derive lower bounds for the TT count. We compare their performance and summarize the results in Table 1. The definition and properties of stabilizer extent are put in Appendix B.4. The quantum Fourier transform QFTn is the unitary where QFT|nj:=12nk=02n1e2πijk/2n|k{}_{n}|j\rangle\mathrel{\mathop{\mathchar 58\relax}}=\frac{1}{\sqrt{2^{n}}}\sum_{k=0}^{2^{n}-1}e^{2\pi ijk/2^{n}}|k\rangle.

As shown in table 1, when compared with state-stabilizer nullity, the lower bounds obtained by unitary-stabilizer nullity is the same as for diagonal gates, and better in the general case. When compared with stabilizer extent, for small nn, the unitary-stabilizer nullity outperforms for some gates. For larger nn, the stabilizer extent might perform better, but it is harder to compute since the cardinality of stabilizer states increases as 2Ω(n2)2^{\Omega(n^{2})} [35].

Unitary Corresponding state Lower bounds obtained by
state-stabilizer nullity unitary-stabilizer nullity stabilizer extent
TT T|+T|+\rangle 1 1 1
T\sqrt{T} T|+\sqrt{T}|+\rangle 1 1 0.7549
CCZCCZ CCZ|+3CCZ|+\rangle^{\otimes 3} 3 3 3.6349
C3ZC^{3}Z C3Z|+4C^{3}Z|+\rangle^{\otimes 4} 4 4 5.1226
THT\sqrt{T}H\sqrt{T} THT|+\sqrt{T}H\sqrt{T}|+\rangle 1 2 0.9109
THTTHT THT|+THT|+\rangle 1 2 1.4542
QFT3 QFT|3x{}_{3}|x\rangle
x0mod2x\equiv 0\mod 2 0 4 0.0000
x1mod2x\equiv 1\mod 2 1 4 1.0000
QFT4 QFT|4x{}_{4}|x\rangle
x1mod2x\equiv 1\mod 2 2 6 1.7555
x0mod4x\equiv 0\mod 4 0 6 0.0000
x2mod4x\equiv 2\mod 4 1 6 1.0000
Table 1: Lower bounds for TT count for the given unitary. The first column is the given unitary UU. The second column is the corresponding state |ψ|\psi\rangle. The next three columns are the lower bounds obtained by state-stabilizer nullity vsv_{s}, the unitary-stabilizer nullity vv, and the stabilizer extent ξ\xi. The corresponding value are vs(|ψ)v_{s}(|\psi\rangle), v(U)v(U) and log(ξ(|ψ)logξ(|T)\frac{\log(\xi(|\psi\rangle)}{\log\xi(|T\rangle)}, as proved in Theorem 11, Theorem 12. Since QFT|n+n=|0{}_{n}|+\rangle^{\otimes n}=|0\rangle, similar to Ref. [25] instead we use QFT|nx{}_{n}|x\rangle as its corresponding states.

V Conclusion

Resource theory for magic monotones has wide applications in quantum information. In this work, we introduce a magic monotone for unitaries, and use it to lower bound the TT count in the circuit synthesis task. This lower bound quantifies the minimum magic resources required for the computation, and lower bounds the runtime of classical simulation algorithms based on the Clifford+TT gate set. We also give numerical results which show that our magic monotone can give better TT count lower bounds for some gates when compared with other magic monotones.

Specifically, in this paper, for any nn-qubit unitary UU, we introduce its unitary-stabilizer nullity as v(U)=2ns(U)v(U)=2n-s(U), where s(U)s(U) is the number of ±1s\pm 1s in the Pauli transfer matrix. We prove that the unitary-stabilizer nullity satisfies several desirable properties, and in particular establish that it is a lower bound for the TT count of UU. We show that the lower bound obtained by the unitary-stabilizer nullity never performs worse than the state-stabilizer nullity. We also give an nn-qubit unitary family of unitary-stabilizer nullity 2n2n, providing examples where the unitary-stabilizer nullity strictly outperforms. For the diagonal unitaries, we prove the lower bounds obtained by the state- and the unitary-stabilizer nullity are the same. Besides, we connect the unitary-stabilizer nullity and the state-stabilizer nullity with auxiliary systems. Moreover, the results in this work may shed lights for the study of resource theories of quantum dynamics (e.g., Refs. [38, 39, 40, 41, 42, 43, 44]).

Meaningful magic monotones with desirable properties are very useful to exploit the power of nonstabilizer resources in quantum computation. Our work has, in particular, shown the applications of unitary-stabilizer nullity in estimating the TT count, which is central for resource estimation for fault-tolerant computation [11], performance analysis for classical-simulation algorithms [8, 5, 6], and even for the complexity of the parameterized Quantum Merlin-Arthur problem (QMA problem) [45]. There are many other applications beyond. For one thing, we could apply the unitary-stabilizer nullity to benchmark the gate synthesis task under stabilizer operations [46, 28, 27]. Using the result in Theorem 13, the benchmark method can be generalized for any non-Clifford gates. For another, one might use the unitary-stabilizer nullity to quantify the nonstabilizerness of quantum circuits [47, 48] on near-term quantum devices, as we could extract the unitary Pauli function on the quantum hardware, via firstly measuring the states U(σv+I)/2nUU(\sigma_{v}+I)/{2^{n}}U^{\dagger} and U(Iσv)/2nUU(I-\sigma_{{\textbf{v}}})/{2^{n}}U^{\dagger} with measurement operator σu\sigma_{{\textbf{u}}}, and then obtain the Pauli function. Protocols for evaluating the nonstabilizerness experimentally [47, 48] can be used to analyze the performance of quantum devices.

Beyond this work, there are still various interesting open questions on the quantification of non-Clifford resources:

  • For the task of lower bounding TT gates for a given unitary, the unitary-stabilizer nullity gives a bound up to 2n2n. However, in general the TT count can be much larger than 2n2n. For example, it is known that only unitaries with elements in certain rings can be exactly synthesized by Clifford and TT gates [49]. Is it possible to get a superlinear bound by devising alternative magic monotones?

  • Here we define the TT count to be the minimum number of TT gates used for exactly synthesizing a given unitary UU. Is it possible to lower bound the number of TT gates for unitary VV which approximates UU, i.e. |VU|ϵ|V-U|\leq\epsilon? Beverland et al. [25] have considered this setting for single-qubit unitaries.

Acknowledgements. We thank the anonymous reviewers for their valuable suggestions, which help us better present the results and proofs. This work was done when J. J. was a research intern at Baidu Research.

Appendix A Equivalence between the definitions of the state-stabilizer nullity

In this section, we show Definition 4 is equivalent to the definition of the state-stabilizer nullity in the original paper [25]. Recall that for an nn-qubit state |ψ|\psi\rangle, Beverland et al [25] define the state-stabilizer nullity as

ν(|ψ)=nlog2|Stab|ψ|,\displaystyle\nu(|\psi\rangle)=n-\log_{2}{|\mathrm{Stab}\,|\psi\rangle}|, (52)

where Stab|ψ\mathrm{Stab}\,|\psi\rangle is the stabilizer of |ψ|\psi\rangle, that is

Stab|ψ={P𝒫^n:P|ψ=|ψ}.\displaystyle\mathrm{Stab}\,|\psi\rangle=\{P\in\hat{{\cal P}}_{n}\mathrel{\mathop{\mathchar 58\relax}}P|\psi\rangle=|\psi\rangle\}. (53)

Here we prove our definition vs(|ψ)v_{s}(|\psi\rangle) is equivalent to ν(|ψ)\nu(|\psi\rangle), that is vs(|ψ)=ν(|ψ)v_{s}(|\psi\rangle)=\nu(|\psi\rangle), by building a one-to-one correspondence between u4[n]{{\textbf{u}}}\in 4^{[n]} where P|ψ(u)=±1P_{|\psi\rangle}({{\textbf{u}}})=\pm 1, and P𝒫^nP\in\hat{{\cal P}}_{n} where P|ψ=|ψP|\psi\rangle=|\psi\rangle. The key is noticing that the state Pauli functions are ranging over the quotient group 𝒫n{\cal P}_{n}, while the Stab\mathrm{Stab} is ranging over the Pauli group 𝒫^n\hat{{\cal P}}_{n} where

𝒫n={σu|u4[n]},\displaystyle{\cal P}_{n}=\{\sigma_{{\textbf{u}}}|{{\textbf{u}}}\in 4^{[n]}\}, (54)
𝒫^n={±σu,±iσu|u4[n]}.\displaystyle\hat{{\cal P}}_{n}=\{\pm\sigma_{{\textbf{u}}},\pm i\sigma_{{\textbf{u}}}|{{\textbf{u}}}\in 4^{[n]}\}. (55)

More precisely,

\rightarrow For one direction, if any u satisfies that P|ψ(u)=γP_{|\psi\rangle}({{\textbf{u}}})=\gamma where γ=±1\gamma=\pm 1, since σu\sigma_{{\textbf{u}}} only has ±1\pm 1 eigenvalues and |ψ|\psi\rangle is unit, we must have

σu|ψ=γ|ψ.\displaystyle\sigma_{{\textbf{u}}}|\psi\rangle=\gamma|\psi\rangle. (56)

Thus we find P(u):=γσu𝒫^nP^{({{\textbf{u}}})}\mathrel{\mathop{\mathchar 58\relax}}=\gamma\sigma_{{\textbf{u}}}\in\hat{{\cal P}}_{n} such that

P(u)Stab|ψ.\displaystyle P^{({{\textbf{u}}})}\in\mathrm{Stab}\,|\psi\rangle. (57)

Further notice that for uv{{\textbf{u}}}\neq{{\textbf{v}}} where P|ψ(u)=P|ψ(v)=±1P_{|\psi\rangle}({{\textbf{u}}})=P_{|\psi\rangle}({{\textbf{v}}})=\pm 1, we have P(u)P(v)P^{({{\textbf{u}}})}\neq P^{({{\textbf{v}}})}. Thus the map from u to P(u)P^{({{\textbf{u}}})} is a bijection.

\leftarrow For the other direction, notice that since iσui\sigma_{{\textbf{u}}} has eigenvalues ±i\pm i, we have

iσu|ψ|ψ,|ψ.\displaystyle i\sigma_{{\textbf{u}}}|\psi\rangle\neq|\psi\rangle,\forall|\psi\rangle. (58)

Then for any PStab|ψP\in\mathrm{Stab}\,|\psi\rangle, there exists uP4[n]{{\textbf{u}}}_{P}\in 4^{[n]} such that P=γσuPP=\gamma\sigma_{{{\textbf{u}}}_{P}} where γ=±1\gamma=\pm 1. We can verify that

P|ψ(up)=γ.\displaystyle P_{|\psi\rangle}({{\textbf{u}}}_{p})=\gamma. (59)

Also notice that by the linearity, if PStab|ψP\in\mathrm{Stab}\,|\psi\rangle, then PStab|ψ-P\not\in\mathrm{Stab}\,|\psi\rangle. Thus for any PPStab|ψP\neq P^{\prime}\in\mathrm{Stab}\,|\psi\rangle, uPuP{{\textbf{u}}}_{P}\neq{{\textbf{u}}}_{P^{\prime}}. Thus the map from PStab|ψP\in\mathrm{Stab}\,|\psi\rangle to uP{{\textbf{u}}}_{P} is also a bijection.

Combing the above argument that there is a bijection from u where P|ψ(u)=±1P_{|\psi\rangle}({{\textbf{u}}})=\pm 1 to the elements of Stab|ψ\mathrm{Stab}\,|\psi\rangle and a bijection vice versa, we conclude that

vs(|ψ)=ν(|ψ).\displaystyle v_{s}(|\psi\rangle)=\nu(|\psi\rangle). (60)

Appendix B Non-Clifford resources in two computational models

In this section, we specify two computation models which we mentioned in this manuscript, that is the quantum computation model with magic states, and the quantum computation model with unitary. We would use examples from [31, 32, 29] to clarify the relationship between the number of non-Clifford resources used in the two models, that is Proposition 25 and Proposition 26.

To begin with, we firstly clarify the two computation models. We write TT gate as T=[100eiπ/4]T=\begin{bmatrix}1&0\\ 0&e^{i\pi/4}\end{bmatrix}, T state as |T=T|+.|T\rangle=T|+\rangle.

B.1 Two computation models

Quantum computation with magic state [11]. In this model, the computational task is to construct a target nn-qubit quantum state ρ\rho, either a pure state or a mixed state. The valid operations are as follows.

  • Initiate qubits in state |0m|Ts|0\rangle^{\otimes m}|T\rangle^{\otimes s} for some integer m,sm,s.

  • Perform stabilizer operations, including using Clifford unitaries and qubit measurement in the computational basis. Here the Clifford unitary is performed adaptively, that is the Clifford unitary used can depend on previous measurement outcomes.

  • After the computation, use the first nn qubits as the output, which should be in state ρ\rho.

There are variants of this model which allow using catalyst states in the initiation, we do not consider those variants here. For readers who are interested, more can be seen in [25].

Quantum computation with unitary. In this model, the computational task is to construct a target nn-qubit unitary UU using basic gates. More specifically, the task is to decompose UU into a sequence of gates from the Clifford+TT gate set, without using ancillary qubits and measurement. Mathematically, we aim to find unitary U1,,UlU_{1},...,U_{l} such that

U=U1U2Ul,\displaystyle U=U_{1}U_{2}...U_{l}, (61)

where each UiU_{i} corresponds to either a Clifford gate, or a TT gate. Here ll is an integer corresponding to the size of the circuit.

Differences: Note that in the quantum computation with magic state, one can use auxiliary qubits, do measurements and perform Clifford unitary adaptively, while in the quantum computation with unitary, one cannot use auxiliary qubits, cannot do measurements and there are no adaptive operations.

B.2 Non-Clifford resources used in the two models

In the quantum computation with magic state model, the T state |T|T\rangle is the non-Clifford resource, while the stabilizer operations are viewed as free operations. For an nn-qubit state ρ\rho, we use ws(ρ)w_{s}(\rho) to denote the minimum ss, that is the minimum number of |T|T\rangle needed to construct ρ\rho. We denote ws(ρ)=+w_{s}(\rho)=+\infty if ρ\rho cannot be exactly constructed through the above process.

In the quantum computation with unitary model, the TT gate is the non-Clifford resources, while the Clifford gates, or Clifford unitaries, are viewed as free resources. We use t(U)t(U) to denote the minimum number of TT gates used in Eq. (61). We denote t(U)=+t(U)=+\infty if UU cannot be exactly synthesized with Clifford+TT gate set. We also call t(U)t(U) the TT count of UU.

B.3 Relation between number of non-Clifford resources in two models

As we inexplicitly used in the manuscript, firstly the lower bound of |T|T\rangle can derive a lower bound of TT. This fact is used in [25].

Proposition 25

[25] For an nn-qubit unitary UU, for any nn-qubit stabilizer state |ϕ|\phi\rangle,

t(U)ws(U|ϕ).\displaystyle t(U)\geq w_{s}(U|\phi\rangle). (62)

Proof.  This proof is direct. Suppose we decompose UU into sequences of gates in Clifford+TT gate set with in total t(U)t(U) TT gates. Then use gate injection [10, 50] we can construct state U|ϕU|\phi\rangle by using t(U)t(U) copies of |T|T\rangle. \blacksquare

Furthermore, in the other direction, one may wonder if t(U)t(U) can be completely characterized by ws()w_{s}(\cdot), via enumerating all stabilizer states |ϕ|\phi\rangle. We emphasize here this is not true.

Proposition 26

[31, 32, 29] There exists 33-qubit unitary UU such that

t(U)>max|ϕws(U|ϕ),\displaystyle t(U)>\max_{|\phi\rangle}w_{s}(U|\phi\rangle), (63)

where the maximum is over all 33-qubit stabilizer states |ϕ|\phi\rangle.

Proof.  This fact is implied by [31, 32, 29]. Here UU is set to be the Toffoli gate. Using algorithm which outputs exactly w()w(\cdot), Gosset shows t(U)=7t(U)=7 [29]. On the other hand, by utilizing the adaptive computation, ws(U|ϕ)4w_{s}(U|\phi\rangle)\leq 4 [32]. \blacksquare

B.4 Background of stabilizer extent

For completeness, here we write down the definition and properties of stabilizer extent, which is from [25].

Definition 6 ([25].)

For an arbitrary pure state |ψ|\psi\rangle, the stabilizer extent ξ(|ψ)\xi(|\psi\rangle) is defined as

ξ(|ψ)=min(c1,c2,,ck)12 s.t |ψ=α=1kcα|ϕα,\displaystyle\xi(|\psi\rangle)=\min\|(c_{1},c_{2},...,c_{k})\|_{1}^{2}\text{ s.t }|\psi\rangle=\sum_{\alpha=1}^{k}c_{\alpha}|\phi_{\alpha}\rangle, (64)

where the minimization is over all complex linear combination of stabilizer states {|ϕα}\{|\phi_{\alpha}\rangle\}.

Theorem 27 ([25])

For any nn-qubit unitary UU, for any single-qubit stabilizer states |ψ1,,|ψn|\psi_{1}\rangle,...,|\psi_{n}\rangle, we have

t(U)logξ(U|ψ1|ψn)logξ(|T).\displaystyle t(U)\geq\frac{\log\xi(U|\psi_{1}\rangle...|\psi_{n}\rangle)}{\log\xi(|T\rangle)}. (65)

Appendix C Clarification on properties of the state-stabilizer nullity

C.1 Corollary 23 does not violate properties of the state-stabilizer nullity

Corollary 23 and equation (46) does not violate the property that the state-stabilizer nullity is invariant under Clifford unitaries, i.e., for any nn-qubit unitary VV, for any nn-qubit Clifford unitary C𝒞nC\in{\cal C}_{n} and any nn-qubit stabilizer state |ψ|\psi\rangle, one have

vs(CV|ψ)=vs(V|ψ).\displaystyle v_{s}(CV|\psi\rangle)=v_{s}(V|\psi\rangle). (66)

What the corollary implies is slightly different. Specifically, what it implies is V,C\exists V,C such that

vs(VC|ψ)vs(V|ψ),\displaystyle v_{s}(VC|\psi\rangle)\neq v_{s}(V|\psi\rangle), (67)

Or equivalently there may exist different nn-qubit stabilizer states |ψ1|\psi_{1}\rangle and |ψ2|\psi_{2}\rangle such that

vs(V|ψ1)vs(V|ψ2).\displaystyle v_{s}(V|\psi_{1}\rangle)\neq v_{s}(V|\psi_{2}\rangle). (68)

Write |Φ|\Phi\rangle to be the 2n2n-qubit maximally entangled states, Corollary 23 implies \exists nn-qubit unitary UU such that setting V=(I2nU)V=(I_{2^{n}}\otimes U)

vs(V|Φ)>vs(V(I2nC′′)|02n),\displaystyle v_{s}(V|\Phi\rangle)>v_{s}(V(I_{2^{n}}\otimes C^{\prime\prime})|0^{\otimes 2n}\rangle), (69)

for any nn-qubit Clifford C′′𝒞nC^{\prime\prime}\in{\cal C}_{n}.

C.2 The state-stabilizer nullity does not satisfy the subadditivity under composition

In this section, we show that for any 11-qubit stabilizer state |ψ|\psi\rangle, there exist single-qubit unitaries U,VU,V such that

vs(UV|ψ)>vs(U|ψ)+vs(V|ψ).\displaystyle v_{s}(UV|\psi\rangle)>v_{s}(U|\psi\rangle)+v_{s}(V|\psi\rangle). (70)

Thus the state-stabilizer nullity does not satisfy the subadditivity under composition. In particular, suppose

|ψ=C|0,C𝒞1.\displaystyle|\psi\rangle=C|0\rangle,C\in{\cal C}_{1}. (71)

Set

U=eiXHC1,V=CHC1.\displaystyle U=e^{iX}HC^{-1},V=CHC^{-1}. (72)

Since |+|+\rangle is an eigenstate of eiXe^{iX}, one can verify that

vs(U|ψ)=vs(eiX|+)=0.\displaystyle v_{s}(U|\psi\rangle)=v_{s}(e^{iX}|+\rangle)=0. (73)

Since CHCH is Clifford, then

vs(V|ψ)=vs(CH|0)=0.\displaystyle v_{s}(V|\psi\rangle)=v_{s}(CH|0\rangle)=0. (74)

Finally, one can verify by calculating the state Pauli function that

vs(UV|ψ)=vs(eiX|0)=1.\displaystyle v_{s}(UV|\psi\rangle)=v_{s}(e^{iX}|0\rangle)=1. (75)

Thus Eq. (70) holds. This counterexample can be generalized to nn-qubit case.

C.3 Non-decreasing of state-stabilizer nullity with higher-dimensional systems

Lemma 28

For any nn-qubit unitary UU, for any ddd\geq d^{\prime}, we have

maxC𝒞d+nvs((I2dU)C|0d+n)maxC𝒞d+nvs((I2dU)C|0d+n).\displaystyle\max_{C\in{\cal C}_{d+n}}v_{s}\left((I_{2^{d}}\otimes U)C|0\rangle^{\otimes d+n}\right)\geq\max_{C^{\prime}\in{\cal C}_{d^{\prime}+n}}v_{s}\left((I_{2^{d^{\prime}}}\otimes U)C^{\prime}|0\rangle^{\otimes d^{\prime}+n}\right). (76)

Proof.  It suffices to notice that, for any C𝒞d+nC^{\prime}\in{\cal C}_{d^{\prime}+n}, let

C:=I2ddC.\displaystyle C\mathrel{\mathop{\mathchar 58\relax}}=I_{2^{d-d^{\prime}}}\otimes C^{\prime}. (77)

we have

vs((I2dU)C|0d+n)\displaystyle v_{s}\left((I_{2^{d}}\otimes U)C|0\rangle^{\otimes d+n}\right) =vs((I2ddI2dU)C|0d+n)\displaystyle=v_{s}\left((I_{2^{d-d^{\prime}}}\otimes I_{2^{d^{\prime}}}\otimes U)C|0\rangle^{\otimes d+n}\right)
=vs(|0dd(I2dU)C|0d+n)\displaystyle=v_{s}\left(|0\rangle^{d-d^{\prime}}\otimes(I_{2^{d^{\prime}}}\otimes U)C^{\prime}|0\rangle^{\otimes d^{\prime}+n}\right)
=vs(|0dd)+vs((I2dU)C|0d+n)\displaystyle=v_{s}\left(|0\rangle^{d-d^{\prime}}\right)+v_{s}\left((I_{2^{d^{\prime}}}\otimes U)C^{\prime}|0\rangle^{\otimes d^{\prime}+n}\right) (78)
=vs((I2dU)C|0d+n).\displaystyle=v_{s}\left((I_{2^{d^{\prime}}}\otimes U)C^{\prime}|0\rangle^{\otimes d^{\prime}+n}\right). (79)

The Eq. (78) is obtained by the fact that the state-stabilizer nullity is additive under tensor product. The Eq. (79) is by the fact that vs(|ϕ)=0v_{s}(|\phi\rangle)=0 for stabilizer state |ϕ|\phi\rangle. The proof of the two facts can be seen in [25]. \blacksquare

C.4 Detailed proofs of Lemma 1, Lemma 2, and Lemma 9

Proof.  [of Lemma 1] We only give proof to the first statement, the second is similar. Recall that

PU(u|v)\displaystyle P_{U}({{\textbf{u}}}|{{\textbf{v}}}) =Tr(σuUσvU)/2n,\displaystyle=\operatorname{Tr}(\sigma_{{\textbf{u}}}U\sigma_{{\textbf{v}}}U^{\dagger})/2^{n}, (80)
Tr(σuσv)\displaystyle\operatorname{Tr}(\sigma_{{\textbf{u}}}\sigma_{{\textbf{v}}}) ={2n, if u=v0, else.\displaystyle=\left\{\begin{aligned} &2^{n},&\text{ if ${{\textbf{u}}}={{\textbf{v}}}$}\\ &0,&\text{ else.}\end{aligned}\right. (81)

Since UU is Clifford, U()UU(\cdot)U^{\dagger} is an isomorphism of 𝒫^n\hat{{\cal P}}_{n}, i.e. U𝒫^nU=𝒫^nU\hat{{\cal P}}_{n}U^{\dagger}=\hat{{\cal P}}_{n}. Thus u4[n]\forall{{\textbf{u}}}\in 4^{[n]}, there exists v4[n],k[4]{{\textbf{v}}}\in 4^{[n]},k\in[4] such that UikσvU=σu,Ui^{k}\sigma_{{\textbf{v}}}U^{\dagger}=\sigma_{{\textbf{u}}}, or equivalently UσvU=(i)kσuU\sigma_{{\textbf{v}}}U^{\dagger}=(-i)^{k}\sigma_{{\textbf{u}}}. Furthermore, since the conjugate operation would not change the eigenvalues of σv\sigma_{{\textbf{v}}}, which are ±1\pm 1. Thus the global phase (i)k(-i)^{k} must be ±1\pm 1. Thus there exists at least one v such that PU(u|v)=±1P_{U}({{\textbf{u}}}|{{\textbf{v}}})=\pm 1.

On the other hand, by the linearity of U()UU(\cdot)U^{\dagger}, if there exists vv{{\textbf{v}}}\neq{{\textbf{v}}}^{\prime} such that

UσvU\displaystyle U\sigma_{{\textbf{v}}}U^{\dagger} =(i)kvσu,\displaystyle=(-i)^{k_{{\textbf{v}}}}\sigma_{{\textbf{u}}}, (82)
UσvU\displaystyle U\sigma_{{{\textbf{v}}}^{\prime}}U^{\dagger} =(i)kvσu.\displaystyle=(-i)^{k_{{{\textbf{v}}}^{\prime}}}\sigma_{{\textbf{u}}}. (83)

Then we have U((i)kvσv(i)kvσv)U=0U\left((-i)^{k_{{{\textbf{v}}}^{\prime}}}\sigma_{{\textbf{v}}}-(-i)^{k_{{{\textbf{v}}}}}\sigma_{{{\textbf{v}}}^{\prime}}\right)U^{\dagger}=0, which leads to a contradiction by the fact that σv\sigma_{{\textbf{v}}} and σv\sigma_{{{\textbf{v}}}^{\prime}} are linear independent. Thus we prove that there is only one such v which satisfies PU(u|v)=±1P_{U}({{\textbf{u}}}|{{\textbf{v}}})=\pm 1. Since U()UU(\cdot)U^{\dagger} is an isomorphism of 𝒫^n\hat{{\cal P}}_{n} by Eq. (LABEL:eq:c01) we know for any vv{{\textbf{v}}}^{\prime}\neq{{\textbf{v}}}, we have PU(u|v)=0P_{U}({{\textbf{u}}}|{{\textbf{v}}}^{\prime})=0. \blacksquare

Proof. [of Lemma 2] Since σu\sigma_{{\textbf{u}}} is a Pauli operator, 2n12^{n-1} number of its eigenvalues are +1+1, and 2n12^{n-1} eigenvalues are 1-1. We write

σu=σu+σu,\displaystyle\sigma_{{\textbf{u}}}=\sigma_{{\textbf{u}}}^{+}-\sigma_{{\textbf{u}}}^{-}, (84)

where σu±\sigma_{{\textbf{u}}}^{\pm} are the projections onto the ±1\pm 1 eigenspace of σu\sigma_{{\textbf{u}}}. Similarly, we write

σv=σv+σv,Qv+=Uσv+U,Qv=UσvU.\displaystyle\sigma_{{\textbf{v}}}=\sigma_{{\textbf{v}}}^{+}-\sigma_{{\textbf{v}}}^{-},\quad Q_{{\textbf{v}}}^{+}=U\sigma_{{\textbf{v}}}^{+}U^{\dagger},\quad Q_{{\textbf{v}}}^{-}=U\sigma_{{\textbf{v}}}^{-}U^{\dagger}. (85)

Then

PU(u|v)\displaystyle P_{U}({{\textbf{u}}}|{{\textbf{v}}}) =TrσuUσvU/2n\displaystyle=\operatorname{Tr}\sigma_{{\textbf{u}}}U\sigma_{{\textbf{v}}}U^{\dagger}/2^{n} (86)
=Tr(2σu+I)(2Qv+I)/2n (Since σu++σu=I)\displaystyle=\operatorname{Tr}(2\sigma_{{\textbf{u}}}^{+}-I)(2Q_{{\textbf{v}}}^{+}-I)/2^{n}\text{ (Since $\sigma_{{\textbf{u}}}^{+}+\sigma_{{\textbf{u}}}^{-}=I$)}
=(4Trσu+Qv+2n)/2n.\displaystyle=(4\operatorname{Tr}\sigma_{{\textbf{u}}}^{+}Q_{{\textbf{v}}}^{+}-2^{n})/2^{n}. (87)

Since PU(u|v)=±1P_{U}({{\textbf{u}}}|{{\textbf{v}}})=\pm 1, we have either

Trσu+Qv+=0.\displaystyle\operatorname{Tr}\sigma_{{\textbf{u}}}^{+}Q_{{\textbf{v}}}^{+}=0. (88)

or

Trσu+Qv+=2n1.\displaystyle\operatorname{Tr}\sigma_{{\textbf{u}}}^{+}Q_{{\textbf{v}}}^{+}=2^{n-1}. (89)

Note that the eigenvalues of σu+\sigma_{{\textbf{u}}}^{+} (or Qv+Q_{{\textbf{v}}}^{+}) are 2n12^{n-1} 11-eigenvalue and 2n12^{n-1} 0-eignenvalue.

If equation (88) holds, we must have Qv+=σuQ_{{\textbf{v}}}^{+}=\sigma_{{\textbf{u}}}^{-}. To prove this, let x1,,x2n1{{\textbf{x}}}_{1},...,{{\textbf{x}}}_{2^{n-1}} be an orthogonal basis of the +1+1-eigenspace of QvQ_{{\textbf{v}}}. Since Qv+Q_{{\textbf{v}}}^{+} and σu\sigma_{{\textbf{u}}}^{-} are all projections, it suffices to prove vector spaces {x|Qv+x=x}={x|σux=x}\{{{\textbf{x}}}|Q_{{\textbf{v}}}^{+}{{\textbf{x}}}={{\textbf{x}}}\}=\{{{\textbf{x}}}|\sigma_{{\textbf{u}}}^{-}{{\textbf{x}}}={{\textbf{x}}}\}. Since the two spaces have the same dimension of 2n12^{n-1}, it suffices to prove

σuxi=xi,i.\displaystyle\sigma_{{\textbf{u}}}^{-}{{\textbf{x}}}_{i}={{\textbf{x}}}_{i},\forall i. (90)

or by the fact σu++σu=I\sigma_{{\textbf{u}}}^{+}+\sigma_{{\textbf{u}}}^{-}=I, equivalently to prove

σu+xi=0,i.\displaystyle\sigma_{{\textbf{u}}}^{+}{{\textbf{x}}}_{i}=0,\forall i. (91)

If the above is not true, i0,σu+xi00\exists i_{0},\sigma_{{\textbf{u}}}^{+}{{\textbf{x}}}_{i_{0}}\neq 0. Extend {xi}i=12n1\{{{\textbf{x}}}_{i}\}_{i=1}^{2^{n-1}} to the basis of the whole (2n2^{n}-dimensional) space, we have

Trσu+Qv+\displaystyle\operatorname{Tr}\sigma_{{\textbf{u}}}^{+}Q_{{\textbf{v}}}^{+} =i=12nxi|σu+Qv+|xi=i=12n1xi|σu+|xixi0|σu+|xi0>0,\displaystyle=\sum_{i=1}^{2^{n}}\langle{{\textbf{x}}}_{i}|\sigma_{{\textbf{u}}}^{+}Q_{{\textbf{v}}}^{+}|{{\textbf{x}}}_{i}\rangle=\sum_{i=1}^{2^{n-1}}\langle{{\textbf{x}}}_{i}|\sigma_{{\textbf{u}}}^{+}|{{\textbf{x}}}_{i}\rangle\geq\langle{{\textbf{x}}}_{i_{0}}|\sigma_{{\textbf{u}}}^{+}|{{\textbf{x}}}_{i_{0}}\rangle>0, (92)

which leads to a contradiction. Thus we conclude Qv+=σuQ_{{\textbf{v}}}^{+}=\sigma_{{\textbf{u}}}^{-}, that is

UσvU\displaystyle U\sigma_{{\textbf{v}}}U^{\dagger} =2Qv+I=2σuI=σu.\displaystyle=2Q_{{\textbf{v}}}^{+}-I=2\sigma_{{\textbf{u}}}^{-}-I=-\sigma_{{\textbf{u}}}. (93)

Then uu\forall{{\textbf{u}}}^{\prime}\neq{{\textbf{u}}}, PU(u|v)=0.P_{U}({{\textbf{u}}}^{\prime}|{{\textbf{v}}})=0.

If equation (89) holds, notice that

TrσuQv+\displaystyle\operatorname{Tr}\sigma_{{\textbf{u}}}^{-}Q_{{\textbf{v}}}^{+} =Tr(Iσu+)Qv+=2n12n1=0.\displaystyle=\operatorname{Tr}(I-\sigma_{{\textbf{u}}}^{+})Q_{{\textbf{v}}}^{+}=2^{n-1}-2^{n-1}=0. (94)

Similarly, we conclude Qv+=σu+Q_{{\textbf{v}}}^{+}=\sigma_{{\textbf{u}}}^{+}. \blacksquare

Proof. [of Lemma 3] Notice that

|𝒫U|\displaystyle|{\cal P}_{U}| =|U𝒫nU𝒫n|=|U(𝒫nU𝒫nU)U|\displaystyle=|U{\cal P}_{n}U^{\dagger}\cap{\cal P}_{n}|=|U({\cal P}_{n}\cap U^{\dagger}{\cal P}_{n}U)U^{\dagger}|
=|𝒫nU𝒫nU|=|𝒫U|.\displaystyle=|{\cal P}_{n}\cap U^{\dagger}{\cal P}_{n}U|=|{\cal P}_{U^{\dagger}}|. (95)

In the following we show that |𝒫U|=s(U)|{\cal P}_{U}|=s(U). Since 𝒫U{\cal P}_{U} is an intersection of two groups, thus 𝒫U{\cal P}_{U} is a group. Since 𝒫U𝒫n{\cal P}_{U}\subseteq{\cal P}_{n}, thus further 𝒫U{\cal P}_{U} is a subgroup of 𝒫n{\cal P}_{n}. To prove |𝒫U|=s(U)|{\cal P}_{U}|=s(U), it suffices to prove that there is an one-to-one correspondence between σu𝒫U\sigma_{{\textbf{u}}}\in{\cal P}_{U} and a pair (u,v)({{\textbf{u}}},{{\textbf{v}}}) such that PU(u|v)=±1P_{U}({{\textbf{u}}}|{{\textbf{v}}})=\pm 1.

The proof is as follows. On one hand if σu𝒫U=U𝒫nU𝒫n\sigma_{{\textbf{u}}}\in{\cal P}_{U}=U{\cal P}_{n}U^{\dagger}\cap{\cal P}_{n}, then by definition there exists vu4[n]{{\textbf{v}}}_{{\textbf{u}}}\in 4^{[n]} such that

UσvuU=±σu.\displaystyle U\sigma_{{{\textbf{v}}}_{{\textbf{u}}}}U^{\dagger}=\pm\sigma_{{\textbf{u}}}. (96)

Thus, we have PU(u|vu)=±1P_{U}({{\textbf{u}}}|{{\textbf{v}}}_{{\textbf{u}}})=\pm 1. Further we know such vu{{\textbf{v}}}_{{{\textbf{u}}}} is unique since Eq. (96) implies UσuU=±σvuU^{\dagger}\sigma_{{{\textbf{u}}}}U=\pm\sigma_{{{\textbf{v}}}_{{{\textbf{u}}}}}, and for any vvu{{\textbf{v}}}\neq{{{\textbf{v}}}_{{{\textbf{u}}}}} we have Tr(σvuσv)=0\operatorname{Tr}(\sigma_{{{\textbf{v}}}_{{{\textbf{u}}}}}\sigma_{{\textbf{v}}})=0.

One the other hand, for any (u,v)({{\textbf{u}}},{{\textbf{v}}}) such that PU(u|v)=±1P_{U}({{\textbf{u}}}|{{\textbf{v}}})=\pm 1, by Lemma 2 we have UσvU=±σuU\sigma_{{\textbf{v}}}U^{\dagger}=\pm\sigma_{{\textbf{u}}} thus

σu𝒫U=U𝒫nU𝒫n.\displaystyle\sigma_{{\textbf{u}}}\in{\cal P}_{U}=U{\cal P}_{n}U^{\dagger}\cap{\cal P}_{n}. (97)

\blacksquare

Proof. [of Lemma 9] For any a1,a2Aa_{1},a_{2}\in A, notice that

a21a1Ba1B=a2B,\displaystyle a_{2}^{-1}a_{1}\in B\rightarrow a_{1}B=a_{2}B, (98)
a21a1Ba1Ba2B=.\displaystyle a_{2}^{-1}a_{1}\not\in B\rightarrow a_{1}B\cap a_{2}B=\emptyset. (99)

Thus A×BA\times B can be written as the union of disjoint cosets: If we write A^\hat{A} as the set of coset representatives, that is a maximum set which satisfies

{A^A,a1,a2A^,a21a1B.\displaystyle\left\{\begin{aligned} &\hat{A}\subseteq A,\\ &\forall a_{1},a_{2}\in\hat{A},\,a_{2}^{-1}a_{1}\not\in B.\end{aligned}\right. (100)

Then we have

A×B=aA^aB,\displaystyle A\times B=\cup_{a\in\hat{A}}aB, (101)
a1,a2A^,a1Ba2B=.\displaystyle\forall a_{1},a_{2}\in\hat{A},a_{1}B\cap a_{2}B=\emptyset. (102)

Thus

|A×B|=|A^||B|\displaystyle|A\times B|=|\hat{A}|\cdot|B| (103)

Since A,BA,B are groups, thus ABA\cap B is a group. Further notice that since A^A\hat{A}\subseteq A, Eq. (100) is equivalent to

{A^A,a1,a2A^,a21a1BA.\displaystyle\left\{\begin{aligned} &\hat{A}\subseteq A,\\ &\forall a_{1},a_{2}\in\hat{A},\,a_{2}^{-1}a_{1}\not\in B\cap A.\end{aligned}\right. (104)

This implies A^\hat{A} is also the set of coset representative for ABA\cap B in AA. Thus

|A^|=|A||AB|.\displaystyle|\hat{A}|=\frac{|A|}{|A\cap B|}. (105)

Combine Eq. (103) and Eq. (105) we finally conclude that

|A×B|=|A||B||AB|.\displaystyle|A\times B|=\frac{|A|\cdot|B|}{|A\cap B|}. (106)

\blacksquare

Appendix D Detailed proofs of main results

D.1 Detailed proof for Theorem 14

Proof.  [Proof of Theorem 14] Fix any nn-qubit stabilizer state, that is |ψ=C|0n|\psi\rangle=C|0\rangle^{\otimes n} for some Clifford operator C𝒞n.C\in{\cal C}_{n}. For any u4[n]{{\textbf{u}}}\in 4^{[n]}, notice that the condition that |0n|0\rangle^{\otimes n} is an eigenvector of CUσuUCC^{\dagger}U^{\dagger}\sigma_{{\textbf{u}}}UC, would imply PU|ψ(u)=±1,P_{U|\psi\rangle}({{\textbf{u}}})=\pm 1, then

CUσuUC±{I,Z}nPU|ψ(u)=±1.\displaystyle C^{\dagger}U^{\dagger}\sigma_{{\textbf{u}}}UC\in\pm\{I,Z\}^{\otimes n}\rightarrow P_{U|\psi\rangle}({{\textbf{u}}})=\pm 1. (107)

Note that in the following all the matrix product is the matrix product modulo phases. For Clifford operator C𝒞nC\in{\cal C}_{n}, define

H\displaystyle H :=𝒫n,\displaystyle\mathrel{\mathop{\mathchar 58\relax}}={\cal P}_{n},
A\displaystyle A :=C{I,Z}nC,\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=C\{I,Z\}^{\otimes n}C^{\dagger},
B\displaystyle B :=𝒫U=U𝒫nU𝒫n.\displaystyle\mathrel{\mathop{\mathchar 58\relax}}={\cal P}_{U^{\dagger}}=U^{\dagger}{\cal P}_{n}U\cap{\cal P}_{n}. (108)

Notice that AA, BB and ABA\cap B are subgroups of HH, with respect to the matrix-product modulo phases. Besides, we have

|H|\displaystyle|H| =4n,\displaystyle=4^{n},
|A|\displaystyle|A| =2n,\displaystyle=2^{n},
|B|\displaystyle|B| =|𝒫U|=|𝒫U|=s(U).\displaystyle=|{\cal P}_{U^{\dagger}}|=|{\cal P}_{U}|=s(U). (by Lemma 3) (109)

Further, we have

|AB|\displaystyle|A\cap B| =|C{I,Z}nCU𝒫nU𝒫n|\displaystyle=|C\{I,Z\}^{\otimes n}C^{\dagger}\cap U^{\dagger}{\cal P}_{n}U\cap{\cal P}_{n}|
=|{I,Z}nCU𝒫nUC𝒫n|\displaystyle=|\{I,Z\}^{\otimes n}\cap C^{\dagger}U^{\dagger}{\cal P}_{n}UC\cap{\cal P}_{n}| (110)
=|{I,Z}nCU𝒫nUC|\displaystyle=|\{I,Z\}^{\otimes n}\cap C^{\dagger}U^{\dagger}{\cal P}_{n}UC| (111)
s(U|ψ).\displaystyle\leq s(U|\psi\rangle). (112)

The Eqs. (110) (111) (112) use the fact that C𝒫nC=𝒫nC^{\dagger}{\cal P}_{n}C={\cal P}_{n}, {I,Z}n𝒫n\{I,Z\}^{\otimes n}\in{\cal P}_{n} and Eq. (107) respectively. By Lemma 9, we conclude

2ns(U)\displaystyle 2^{n}\cdot s(U) |AB|4n,\displaystyle\leq|A\cap B|\cdot 4^{n}, (113)
s(U|ψ)4n,\displaystyle\leq s(U|\psi\rangle)\cdot 4^{n},
nlog2s(U|ψ)\displaystyle n-\log_{2}s(U|\psi\rangle) 2nlog2s(U).\displaystyle\leq 2n-\log_{2}s(U).

Thus

vs(U|ψ)v(U).\displaystyle v_{s}(U|\psi\rangle)\leq v(U). (114)

The equality in (114) holds if and only if equality holds for Eq. (112) and Eq. (113), that is

{u4[n]|0nCUσuUC|0n=±1}\displaystyle\{{{\textbf{u}}}\in 4^{[n]}\,|\,\langle 0^{\otimes n}C^{\dagger}U^{\dagger}\sigma_{{\textbf{u}}}UC|0\rangle^{\otimes n}=\pm 1\}
={u4[n]|CUσuUC±{I,Z}n},\displaystyle=\{{{\textbf{u}}}\in 4^{[n]}|C^{\dagger}U^{\dagger}\sigma_{{\textbf{u}}}UC\in\pm\{I,Z\}^{\otimes n}\}, (115)

and A×B=HA\times B=H. \blacksquare

D.2 Detailed proof of Lemma  16

Proof. [Proof of Lemma 16] We prove the inclusion relationships for both sides.

\leftarrow Suppose u4[n]{{\textbf{u}}}\in 4^{[n]} satisfies that UσuU±{I,X}nU^{\dagger}\sigma_{{\textbf{u}}}U\in\pm\{I,X\}^{\otimes n}. Since

Hn{I,X}nHn={I,Z}n.\displaystyle H^{\otimes n}\{I,X\}^{\otimes n}H^{\otimes n}=\{I,Z\}^{\otimes n}. (116)

We have

+|nUσuU|+n\displaystyle\langle+|^{\otimes n}U^{\dagger}\sigma_{{\textbf{u}}}U|+\rangle^{\otimes n} =±1.\displaystyle=\pm 1. (117)

\rightarrow Since UU is a diagonal unitary, we can write

U=diag(,λx,),x{0,1}n,|λx|=1.\displaystyle U=diag(\cdots,\lambda_{{\textbf{x}}},\cdots),{{\textbf{x}}}\in\{0,1\}^{n},|\lambda_{{\textbf{x}}}|=1. (118)

Since

U|+n=12nx{0,1}nλx|x,\displaystyle U|+\rangle^{\otimes n}=\frac{1}{\sqrt{2^{n}}}\sum_{{{\textbf{x}}}\in\{0,1\}^{n}}\lambda_{{{\textbf{x}}}}|{{\textbf{x}}}\rangle, (119)

then

+|nUσuU|+n\displaystyle\langle+|^{\otimes n}U^{\dagger}\sigma_{{\textbf{u}}}U|+\rangle^{\otimes n} =12nx,y{0,1}nλxλyxσuy..\displaystyle=\frac{1}{2^{n}}\sum_{{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}^{n}}\lambda_{{\textbf{x}}}^{\dagger}\lambda_{{\textbf{y}}}\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}\rangle.. (120)

Note that |xσuy|{0,1}|\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}\rangle|\in\{0,1\} and Eq. (120) has in total 2n×2n2^{n}\times 2^{n} terms. However, for fixed u, for any x{0,1}n{{\textbf{x}}}\in\{0,1\}^{n}, there exists exactly one yx{0,1}n{{\textbf{y}}}_{{{\textbf{x}}}}\in\{0,1\}^{n} such that |xσuyx|=1|\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}_{{{\textbf{x}}}}\rangle|=1, and other terms are all 0, that is

|xσuyx|=1,\displaystyle|\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}_{{{\textbf{x}}}}\rangle|=1,
|xσuy|=0,yyx.\displaystyle|\langle{{\textbf{x}}}\sigma_{{{\textbf{u}}}}{{\textbf{y}}}\rangle|=0,\forall{{\textbf{y}}}\neq{{\textbf{y}}}_{{{\textbf{x}}}}. (121)

Thus if +|nUσuU|+n=1\langle+|^{\otimes n}U^{\dagger}\sigma_{{\textbf{u}}}U|+\rangle^{\otimes n}=1 (The case for 1-1 is similar), we must have all the nonzero term λxλyxxσuyx=1\lambda_{{\textbf{x}}}^{\dagger}\lambda_{{{\textbf{y}}}_{{{\textbf{x}}}}}\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}_{{{\textbf{x}}}}\rangle=1. Together with the fact that xσuy=0,yyx\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}\rangle=0,\forall{{\textbf{y}}}\neq{{\textbf{y}}}_{{\textbf{x}}}. Thus we have

λxλyxσuy=|λxλyxσuy|,x,y{0,1}n.\displaystyle\lambda_{{\textbf{x}}}^{\dagger}\lambda_{{\textbf{y}}}\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}\rangle=|\lambda_{{\textbf{x}}}^{\dagger}\lambda_{{\textbf{y}}}\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}\rangle|,\quad\forall{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}^{\otimes n}. (122)

Note that for any operator ()(\cdot), the following holds

()=x,y{0,1}nx()y|xy|.\displaystyle(\cdot)=\sum_{{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}^{n}}\langle{{\textbf{x}}}(\cdot){{\textbf{y}}}\rangle|{{\textbf{x}}}\rangle\langle{{\textbf{y}}}|. (123)

Thus

UσuU\displaystyle U^{\dagger}\sigma_{{\textbf{u}}}U =x,y{0,1}nx(UσuU)y|xy|\displaystyle=\sum_{{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}^{n}}\langle{{\textbf{x}}}(U^{\dagger}\sigma_{{\textbf{u}}}U){{\textbf{y}}}\rangle|{{\textbf{x}}}\rangle\langle{{\textbf{y}}}|
=x,y{0,1}nλxλyxσuy|xy|\displaystyle=\sum_{{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}^{n}}\lambda_{{\textbf{x}}}^{\dagger}\lambda_{{\textbf{y}}}\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}\rangle|{{\textbf{x}}}\rangle\langle{{\textbf{y}}}|
=x,y{0,1}n|λxλyxσuy||xy|\displaystyle=\sum_{{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}^{n}}|\lambda_{{\textbf{x}}}^{\dagger}\lambda_{{\textbf{y}}}\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}\rangle|\cdot|{{\textbf{x}}}\rangle\langle{{\textbf{y}}}|
=x,y{0,1}n|xσuy||xy|\displaystyle=\sum_{{{\textbf{x}}},{{\textbf{y}}}\in\{0,1\}^{n}}|\langle{{\textbf{x}}}\sigma_{{\textbf{u}}}{{\textbf{y}}}\rangle|\cdot|{{\textbf{x}}}\rangle\langle{{\textbf{y}}}|
{I,X}n.\displaystyle\in\{I,X\}^{\otimes n}. (124)

The above four equalities and the final inclusion are obtained by Eq. (123), the fact that U|y=λy|yU|{{\textbf{y}}}\rangle=\lambda_{{\textbf{y}}}|{{\textbf{y}}}\rangle, Eq. (122), the fact that |λx|=|λy|=1|\lambda_{{\textbf{x}}}|=|\lambda_{{\textbf{y}}}|=1 and Lemma 15 respectively. \blacksquare

D.3 Detailed proof of Theorem 20

Proof. [of Theorem 20] We prove this theorem by directly calculating the unitary Pauli functions. Recall that for any x{0,1}n{{\textbf{x}}}\in\{0,1\}^{n}, (x):=x1x2xn\odot({{\textbf{x}}})\mathrel{\mathop{\mathchar 58\relax}}=x_{1}x_{2}...x_{n}, and 1n1^{n} is the nn-bit string of all ones. If we ignore a global phase ±1\pm 1, we may write any σu,σv𝒫n\sigma_{{\textbf{u}}},\sigma_{{\textbf{v}}}\in{\cal P}_{n} as

σu=XsZt,σv=XpZq.\displaystyle\sigma_{{\textbf{u}}}=X^{{{\textbf{s}}}}Z^{{{\textbf{t}}}},\sigma_{{\textbf{v}}}=X^{{{\textbf{p}}}}Z^{{{\textbf{q}}}}. (125)

where s,t,p,q{0,1}n{{\textbf{s}}},{{\textbf{t}}},{{\textbf{p}}},{{\textbf{q}}}\in\{0,1\}^{n}. Notice that ignoring the global phase ±1\pm 1 does not influence the number of ±1\pm 1s in the Pauli functions. Write

U=Cn1ZHnCn1Z.\displaystyle U=C^{n-1}Z\,H^{\otimes n}\,C^{n-1}Z. (126)


One can verify that for any x{0,1}n{{\textbf{x}}}\in\{0,1\}^{n},

x|σuUσvU|x\displaystyle\langle{{\textbf{x}}}|\sigma_{{\textbf{u}}}U\sigma_{{\textbf{v}}}U^{\dagger}|{{\textbf{x}}}\rangle
=12n(1)sp+st(1)x(p+t)(1)(x)+(s+x)\displaystyle=\frac{1}{2^{n}}(-1)^{{{\textbf{s}}}\cdot{{\textbf{p}}}+{{\textbf{s}}}\cdot{{\textbf{t}}}}(-1)^{{{\textbf{x}}}\cdot({{\textbf{p}}}+{{\textbf{t}}})}(-1)^{\odot({{\textbf{x}}})+\odot({{\textbf{s}}}+{{\textbf{x}}})}
×y{0,1}n(1)(q+s)y(1)(y)+(y+p).\displaystyle\,\quad\times\!\!\!\!\sum_{{{\textbf{y}}}\in\{0,1\}^{n}}(-1)^{({{\textbf{q}}}+{{\textbf{s}}})\cdot{{\textbf{y}}}}(-1)^{\odot({{\textbf{y}}})+\odot({{\textbf{y}}}+{{\textbf{p}}})}. (127)

Define

f(q,s,p)=y{0,1}n(1)(q+s)y(1)(y)+(y+p).\displaystyle f({{\textbf{q}}},{{\textbf{s}}},{{\textbf{p}}})=\sum_{{{\textbf{y}}}\in\{0,1\}^{n}}(-1)^{({{\textbf{q}}}+{{\textbf{s}}})\cdot{{\textbf{y}}}}(-1)^{\odot({{\textbf{y}}})+\odot({{\textbf{y}}}+{{\textbf{p}}})}. (128)

We can see that

Tr(σuUσvU)\displaystyle\operatorname{Tr}(\sigma_{{\textbf{u}}}U\sigma_{{\textbf{v}}}U^{\dagger}) =x{0,1}nx|σuUσvU|x\displaystyle=\sum_{{{\textbf{x}}}\in\{0,1\}^{n}}\langle{{\textbf{x}}}|\sigma_{{\textbf{u}}}U\sigma_{{\textbf{v}}}U^{\dagger}|{{\textbf{x}}}\rangle (129)
=12n(1)sp+stf(p,t,s)f(q,s,p).\displaystyle=\frac{1}{2^{n}}(-1)^{{{\textbf{s}}}\cdot{{\textbf{p}}}+{{\textbf{s}}}\cdot{{\textbf{t}}}}f({{\textbf{p}}},{{\textbf{t}}},{{\textbf{s}}})f({{\textbf{q}}},{{\textbf{s}}},{{\textbf{p}}}). (130)

To further simplify the equations, similar as in [25] we notice that

(1)(y)+(y+p)={1 y, if p=0;1 if p0y1n and yp1n;1 if p0y=1n or p+1n\displaystyle(-1)^{\odot({{\textbf{y}}})+\odot({{\textbf{y}}}+{{\textbf{p}}})}=\left\{\begin{aligned} &1\text{\quad$\forall{{\textbf{y}}}$, if ${{\textbf{p}}}=0$};\\ &1\text{\quad if ${{\textbf{p}}}\neq 0$;\, ${{\textbf{y}}}\neq 1^{n}$ and ${{\textbf{y}}}\neq{{\textbf{p}}}\oplus 1^{n}$};\\ &-1\text{\quad if ${{\textbf{p}}}\neq 0$;\, ${{\textbf{y}}}=1^{n}$ or ${{\textbf{p}}}+1^{n}$. }\end{aligned}\right. (131)

Combine the fact that

y{0,1}n(1)zy={2n, if z=0,0, else.\displaystyle\sum_{{{\textbf{y}}}\in\{0,1\}^{n}}(-1)^{{{\textbf{z}}}\cdot{{\textbf{y}}}}=\left\{\begin{aligned} &2^{n},\text{ if ${{\textbf{z}}}=0$,}\\ &0,\text{ else.}\end{aligned}\right. (132)

We can see that

f(q,s,p)={2n, if p=0,q=s;0, if p=0,qs;2n4, if p0,q=s;2(1)(q+s)1n2(1)(q+s)(1n+p), if p0,qs.\displaystyle f({{\textbf{q}}},{{\textbf{s}}},{{\textbf{p}}})=\left\{\begin{aligned} &\bullet\quad 2^{n},\text{ if ${{\textbf{p}}}=0,{{\textbf{q}}}={{\textbf{s}}}$};\\ &\bullet\quad 0,\text{ if ${{\textbf{p}}}=0,{{\textbf{q}}}\neq{{\textbf{s}}}$};\\ &\bullet\quad 2^{n}-4,\text{ if ${{\textbf{p}}}\neq 0,{{\textbf{q}}}={{\textbf{s}}}$};\\ &\bullet\quad-2(-1)^{({{\textbf{q}}}+{{\textbf{s}}})\cdot 1^{n}}-2(-1)^{({{\textbf{q}}}+{{\textbf{s}}})\cdot(1^{n}+{{\textbf{p}}})},\\ &\quad\quad\text{ if ${{\textbf{p}}}\neq 0,{{\textbf{q}}}\neq{{\textbf{s}}}$.}\\ \end{aligned}\right. (133)

Combine with Eq. (129) we know that the unitary Pauli function Tr(σuUσvU)/2n=±1\operatorname{Tr}(\sigma_{{\textbf{u}}}U\sigma_{{\textbf{v}}}U^{\dagger})/2^{n}=\pm 1 if and only if both f(p,t,s)=2n,f(q,s,p)=2nf({{\textbf{p}}},{{\textbf{t}}},{{\textbf{s}}})=2^{n},f({{\textbf{q}}},{{\textbf{s}}},{{\textbf{p}}})=2^{n}, thus by Eq. (133) we conclude this implies

p=t=q=s=0.\displaystyle{{\textbf{p}}}={{\textbf{t}}}={{\textbf{q}}}={{\textbf{s}}}=0. (134)

Thus we conclude that

s(U)=1,\displaystyle s(U)=1,
v(U)=2nlog2s(U)=2n.\displaystyle v(U)=2n-\log_{2}s(U)=2n. (135)

\blacksquare

D.4 Detailed proof of Theorem 24

Proof. [Proof of Theorem 24] Since for any ddd\geq d^{\prime}, we have

maxC𝒞d+nvs((I2dU)C|0d+n)\displaystyle\max_{C\in{\cal C}_{d+n}}v_{s}\left((I_{2^{d}}\otimes U)C|0\rangle^{\otimes d+n}\right)
maxC𝒞d+nvs((I2dU)C|0d+n).\displaystyle\geq\max_{C^{\prime}\in{\cal C}_{d^{\prime}+n}}v_{s}\left((I_{2^{d^{\prime}}}\otimes U)C^{\prime}|0\rangle^{\otimes d^{\prime}+n}\right). (136)
(Appendix C.3)

It suffices to give proofs for sufficiently large dd, namely dnd\geq n. For any dnd\geq n, by Theorem 14 we know that

v(IdU)maxC𝒞d+nvs((I2dU)C|0d+n).\displaystyle v(I_{d}\otimes U)\geq\max_{C\in{\cal C}_{d+n}}v_{s}\left((I_{2^{d}}\otimes U)C|0\rangle^{\otimes d+n}\right). (137)

Besides notice that since the unitary-stabilizer nullity satisfies the additivity under tensor product and faithfulness, we have

v(IdU)\displaystyle v(I_{d}\otimes U) =v(Id)+v(U),\displaystyle=v(I_{d})+v(U), (Theorem 7)
=0+v(U).\displaystyle=0+v(U). (Theorem 5) (138)

Thus, we have

v(U)maxC𝒞d+nvs((I2dU)C|0d+n).\displaystyle v(U)\geq\max_{C\in{\cal C}_{d+n}}v_{s}\left((I_{2^{d}}\otimes U)C|0\rangle^{\otimes d+n}\right). (139)

Moreover, by Theorem 22 we have

v(U)\displaystyle v(U) =vs(I2nU|ϕ)\displaystyle=v_{s}\left(I_{2^{n}}\otimes U|\phi\rangle\right)
maxC𝒞2nvs((I2nU)C|02n)\displaystyle\leq\max_{C\in{\cal C}_{2n}}v_{s}\left((I_{2^{n}}\otimes U)C|0\rangle^{\otimes 2n}\right)
maxC𝒞d+nvs((I2dU)C|0d+n).\displaystyle\leq\max_{C\in{\cal C}_{d+n}}v_{s}\left((I_{2^{d}}\otimes U)C|0\rangle^{\otimes d+n}\right). (140)

The last inequality is obtained by noticing dnd\geq n and using Eq. (137). Combine the above inequalities (139) and (140) we conclude for dnd\geq n,

v(U)=maxC𝒞d+nvs((I2dU)C|0d+n).\displaystyle v(U)=\max_{C\in{\cal C}_{d+n}}v_{s}\left((I_{2^{d}}\otimes U)C|0\rangle^{\otimes d+n}\right). (141)

Combine with Eq. (137), we conclude

v(U)=maxd+maxC𝒞d+nvs((I2dU)C|0(d+n)).\displaystyle v(U)=\max_{d\in\mathbb{N}^{+}}\max_{C\in{\cal C}_{d+n}}v_{s}\left((I_{2^{d}}\otimes U)\,C|0\rangle^{\otimes(d+n)}\right). (142)

\blacksquare

References

  • Gottesman [1998] D. Gottesman, The heisenberg representation of quantum computers, arXiv preprint quant-ph/9807006  (1998).
  • Aaronson and Gottesman [2004] S. Aaronson and D. Gottesman, Improved simulation of stabilizer circuits, Physical Review A 70, 052328 (2004).
  • Dawson and Nielsen [2006] C. Dawson and M. Nielsen, The solovay-kitaev algorithm, Quantum Information and Computation 6, 81 (2006).
  • Nielsen and Chuang [2002] M. A. Nielsen and I. Chuang, Quantum computation and quantum information, American Association of Physics Teachers  (2002).
  • Bravyi and Gosset [2016] S. Bravyi and D. Gosset, Improved classical simulation of quantum circuits dominated by clifford gates, Physical review letters 116, 250501 (2016).
  • Bravyi et al. [2016] S. Bravyi, G. Smith, and J. A. Smolin, Trading classical and quantum computational resources, Physical Review X 6, 021043 (2016).
  • Bravyi et al. [2019] S. Bravyi, D. Browne, P. Calpin, E. Campbell, D. Gosset, and M. Howard, Simulation of quantum circuits by low-rank stabilizer decompositions, Quantum 3, 181 (2019).
  • Seddon et al. [2021] J. R. Seddon, B. Regula, H. Pashayan, Y. Ouyang, and E. T. Campbell, Quantifying quantum speedups: Improved classical simulation from tighter magic monotones, PRX Quantum 2, 010345 (2021).
  • Kocia [2020] L. Kocia, Improved strong simulation of universal quantum circuits, arXiv preprint arXiv:2012.11739  (2020).
  • Zhou et al. [2000] X. Zhou, D. W. Leung, and I. L. Chuang, Methodology for quantum logic gate construction, Physical Review A 62, 052316 (2000).
  • Bravyi and Kitaev [2005] S. Bravyi and A. Kitaev, Universal quantum computation with ideal clifford gates and noisy ancillas, Physical Review A 71, 022316 (2005).
  • Gottesman [1997] D. Gottesman, Stabilizer Codes and Quantum Error Correction, Ph.D. thesis, California Institute of Technology (1997).
  • Gottesman and Chuang [1999] D. Gottesman and I. L. Chuang, Demonstrating the viability of universal quantum computation using teleportation and single-qubit operations, Nature 402, 390 (1999).
  • Bravyi and Haah [2012] S. Bravyi and J. Haah, Magic-state distillation with low overhead, Physical Review A 86, 052329 (2012).
  • Campbell et al. [2017] E. T. Campbell, B. M. Terhal, and C. Vuillot, Review Roads towards fault-tolerant universal quantum computation, Nature Publishing Group 549, 172 (2017).
  • Hastings and Haah [2018] M. B. Hastings and J. Haah, Distillation with Sublogarithmic Overhead, Physical Review Letters 120, 050504 (2018).
  • Chamberland and Cross [2019] C. Chamberland and A. W. Cross, Fault-tolerant magic state preparation with flag qubits, Quantum 3, 143 (2019).
  • Wang et al. [2020] X. Wang, M. M. Wilde, and Y. Su, Efficiently Computable Bounds for Magic State Distillation, Physical Review Letters 124, 090505 (2020).
  • O’Gorman and Campbell [2017] J. O’Gorman and E. T. Campbell, Quantum computation with realistic magic-state factories, Physical Review A 95, 032338 (2017).
  • Krishna and Tillich [2019] A. Krishna and J.-P. Tillich, Towards Low Overhead Magic State Distillation, Physical Review Letters 123, 070507 (2019).
  • Fang and Liu [2020] K. Fang and Z.-W. Liu, No-Go Theorems for Quantum Resource Purification, Physical Review Letters 125, 060405 (2020).
  • Regula et al. [2020] B. Regula, K. Bu, R. Takagi, and Z.-W. Liu, Benchmarking one-shot distillation in general quantum resource theories, Physical Review A 101, 062315 (2020).
  • Wang et al. [2019] X. Wang, M. M. Wilde, and Y. Su, Quantifying the magic of quantum channels, New Journal of Physics 21, 103002 (2019).
  • Chitambar and Gour [2019] E. Chitambar and G. Gour, Quantum resource theories, Reviews of Modern Physics 91, 025001 (2019).
  • Beverland et al. [2020] M. E. Beverland, E. Campbell, M. Howard, and V. Kliuchnikov, Lower bounds on the non-clifford resources for quantum computations, Quantum Science and Technology 10.1088/2058-9565/ab8963 (2020).
  • Selinger [2015] P. Selinger, Efficient clifford+ t approximation of single-qubit operators, Quantum Information & Computation 15, 159 (2015).
  • Ross and Selinger [2016] N. J. Ross and P. Selinger, Optimal ancilla-free clifford+ t approximation of z-rotations., Quantum Inf. Comput. 16, 901 (2016).
  • Heyfron and Campbell [2018] L. E. Heyfron and E. T. Campbell, An efficient quantum compiler that reduces t count, Quantum Science and Technology 4, 015004 (2018).
  • Gosset et al. [2014] D. Gosset, V. Kliuchnikov, M. Mosca, and V. Russo, An algorithm for the TT count, Quantum Information & Computation 14, 1261 (2014).
  • Mosca and Mukhopadhyay [2020] M. Mosca and P. Mukhopadhyay, A polynomial time and space heuristic algorithm for TT count, arXiv preprint arXiv:2006.12440  (2020).
  • Howard and Campbell [2017] M. Howard and E. Campbell, Application of a resource theory for magic states to fault-tolerant quantum computing, Physical review letters 118, 090501 (2017).
  • Jones [2013] C. Jones, Low-overhead constructions for the fault-tolerant toffoli gate, Physical Review A 87, 022328 (2013).
  • Regula [2018] B. Regula, Convex geometry of quantum resource quantification, Journal of Physics A: Mathematical and Theoretical 51, 045303 (2018).
  • Seddon and Campbell [2019] J. R. Seddon and E. T. Campbell, Quantifying magic for multi-qubit operations, Proceedings of the Royal Society A 475, 20190251 (2019).
  • Gross [2006] D. Gross, Hudson’s theorem for finite-dimensional quantum systems, Journal of mathematical physics 47, 122107 (2006).
  • Veitch et al. [2014] V. Veitch, S. A. H. Mousavian, D. Gottesman, and J. Emerson, The resource theory of stabilizer quantum computation, New Journal of Physics 16, 013009 (2014).
  • Dummit and Foote [2004] D. S. Dummit and R. M. Foote, Abstract algebra, Vol. 3 (Wiley Hoboken, 2004).
  • Díaz et al. [2018] M. G. Díaz, K. Fang, X. Wang, M. Rosati, M. Skotiniotis, J. Calsamiglia, and A. Winter, Using and reusing coherence to realize quantum processes, Quantum 2, 100 (2018).
  • Wang and Wilde [2019] X. Wang and M. M. Wilde, Resource theory of asymmetric distinguishability for quantum channels, Physical Review Research 1, 033169 (2019).
  • Rosset et al. [2018] D. Rosset, F. Buscemi, and Y.-C. Liang, Resource Theory of Quantum Memories and Their Faithful Verification with Minimal Assumptions, Physical Review X 8, 021033 (2018).
  • Li et al. [2020] L. Li, K. Bu, and Z.-w. Liu, Quantifying the resource content of quantum channels: An operational approach, Physical Review A 101, 022335 (2020).
  • Regula and Takagi [2021] B. Regula and R. Takagi, Fundamental limitations on distillation of quantum channel resources, Nature Communications 12, 1 (2021).
  • Liu and Winter [2019] Z.-W. Liu and A. Winter, Resource theories of quantum channels and the universal role of resource erasure, arXiv preprint arXiv:1904.04201  (2019).
  • Wang and Wilde [2023] X. Wang and M. M. Wilde, Exact entanglement cost of quantum states and channels under positive-partial-transpose-preserving operations, Physical Review A 107, 012429 (2023).
  • Arunachalam et al. [2022] S. Arunachalam, S. Bravyi, C. Nirkhe, and B. O’Gorman, The parameterized complexity of quantum verification, arXiv preprint arXiv:2202.08119  (2022).
  • Welch et al. [2014] J. Welch, A. Bocharov, and K. M. Svore, Efficient approximation of diagonal unitaries over the clifford+ t basis, arXiv preprint arXiv:1412.5608  (2014).
  • Oliviero et al. [2022] S. F. Oliviero, L. Leone, A. Hamma, and S. Lloyd, Measuring magic on a quantum processor, arXiv preprint arXiv:2204.00015  (2022).
  • Haug and Kim [2022] T. Haug and M. Kim, Scalable measures of magic for quantum computers, arXiv preprint arXiv:2204.10061  (2022).
  • Giles and Selinger [2013] B. Giles and P. Selinger, Exact synthesis of multiqubit clifford+ t circuits, Physical Review A 87, 032332 (2013).
  • Fowler et al. [2012] A. G. Fowler, M. Mariantoni, J. M. Martinis, and A. N. Cleland, Surface codes: Towards practical large-scale quantum computation, Physical Review A 86, 032324 (2012).