arxiv:2106.05305v1 [quant-ph] 9 jun 2021

18
Linear growth of quantum circuit complexity Jonas Haferkamp, 1, 2 Philippe Faist, 1 Naga B. T. Kothakonda, 1, 3 Jens Eisert, 1, 2 and Nicole Yunger Halpern 4, 5, 6, 7, 8 1 Dahlem Center for Complex Quantum Systems, Freie Universität Berlin, 14195 Berlin, Germany 2 Helmholtz-Zentrum Berlin für Materialien und Energie, 14109 Berlin, Germany 3 Institute for Theoretical Physics, University of Cologne, D-50937, Berlin, Germany 4 ITAMP, Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138, USA 5 Department of Physics, Harvard University, Cambridge, MA 02138, USA 6 Research Laboratory of Electronics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA 7 Center for Theoretical Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA 8 Institute for Physical Science and Technology, University of Maryland, College Park, MD 20742, USA Quantifying quantum states’ complexity is a key problem in various subfields of science, from quantum computing to black-hole physics. We prove a prominent conjecture by Brown and Susskind about how random quantum circuits’ complexity increases. Consider constructing a unitary from Haar-random two-qubit quantum gates. Implementing the unitary exactly requires a circuit of some minimal number of gates—the unitary’s exact circuit complexity. We prove that this complexity grows linearly in the number of random gates, with unit probability, until saturating after exponentially many random gates. Our proof is surprisingly short, given the established difficulty of lower-bounding the exact circuit complexity. Our strategy combines differential topology and elementary algebraic geometry with an inductive construction of Clifford circuits. Quantum complexity is a pervasive concept at the inter- section of quantum computing, quantum many-body systems, and black-hole physics. In general, complexity quantifies the resources required to implement a computation. For example, a Boolean function’s complexity can be defined as the mini- mal number of gates, chosen from a given gate set, necessary to evaluate the function. In quantum computing, the circuit model provides a natural measure of complexity: A unitary transformation’s complexity is the size of the smallest circuit that effects the unitary. Similarly, a pure state has a complex- ity definable as the size of the smallest circuit that produces the state from a product state. Quantum complexity has applications in numerous sub- fields of quantum physics, including the foundations of quan- tum computing. Quantum computational complexity is rooted in circuit complexity: A problem is deemed “easy” if it is soluble with a circuit whose size grows polynomially with the input’s size. The problem is deemed “hard” if the cir- cuit’s size scales exponentially. Quantum complexity also features in the definition of phases of matter. For instance, a topological phase is defined in terms of a high-complexity quantum state. Recently, close connections have been discov- ered between gate complexity and holography in high-energy physics, in the context of the anti-de-Sitter-space/conformal- field-theory (AdS/CFT) correspondence [7, 10, 12, 52, 55]. The “complexity equals volume” conjecture [52] suggests that the correspondence’s boundary state has a complexity propor- tional to the volume behind the event horizon of a black hole in the bulk geometry. Similarly, the “complexity equals ac- tion” conjecture posits that a holographic state’s complexity is dual to a certain space-time region’s action [11]. These deep physical connections have motivated studies of quan- tum complexity as a means of illuminating quantum many- body systems’ complex behaviors. Lower-bounding the quan- tum complexity is a long-standing open problem in quan- tum information theory. Underlying this difficulty is that gates performed early in a circuit may cancel with gates per- formed later. Consequently, quantum-gate–synthesis algo- rithms, which decompose a given unitary into gates, run for times exponential in the system size [1]. Several attempts have been made to lower-bound unitaries’ circuit complexi- ties as defined geometrically by Nielsen [18, 19, 38, 40, 41]. A few claims have been proved about circuit complexity; for instance, polynomially deep circuits implement only a small fraction of the unitaries. Most unitaries have complexities exponentially large in the system size, according to a count- ing argument [30]. A key question in the study of quantum complexity is the following. Consider constructing deeper and deeper circuits for an n-qubit system, by applying ran- dom two-qubit gates. At what rate does the circuit complex- ity increase? Brown and Susskind conjectured that quantum circuits’ complexity generically grows linearly for an expo- nentially long time [12, 56]. Brown and Susskind supported the conjecture using the complexity geometry introduced by Nielsen [38, 41]. Further evidence for the conjecture arises from counting arguments [49]. Growth-of-complexity studies advanced further in [9], which connected a strong notion of quantum complexity to unitary t-designs (finite collections of unitaries that approximate completely random unitaries [42]). Our results govern a straightforward notion of circuit com- plexity, a notion based on exact implementations of a uni- tary [18, 38, 41]. The unitaries under consideration are con- structed from Haar-random two-qubit gates. We prove that the circuit complexity grows linearly with time (with the num- ber of gates applied). The proof is surprisingly short, given the difficulty of lower-bounding complexity. Our bound on the complexity holds for all random circuits described above, with probability 1. Instead of invoking unitary designs [9] or Nielsen’s geometric approach [18, 38, 40, 41], we employ el- ementary aspects of differential topology and algebraic geom- etry, combined with an inductive construction of Clifford cir- cuits. Clifford circuits play a pivotal role in quantum comput- ing, as circuits that can easily be implemented fault-tolerantly. This work is organized as follows. First, we introduce the setup and definitions. Second, we present the main result, the complexity’s exponentially long linear growth. We present a arXiv:2106.05305v1 [quant-ph] 9 Jun 2021

Upload: others

Post on 18-Dec-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Linear growth of quantum circuit complexity

Jonas Haferkamp,1, 2 Philippe Faist,1 Naga B. T. Kothakonda,1, 3 Jens Eisert,1, 2 and Nicole Yunger Halpern4, 5, 6, 7, 8

1Dahlem Center for Complex Quantum Systems, Freie Universität Berlin, 14195 Berlin, Germany2Helmholtz-Zentrum Berlin für Materialien und Energie, 14109 Berlin, Germany

3Institute for Theoretical Physics, University of Cologne, D-50937, Berlin, Germany4ITAMP, Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138, USA

5Department of Physics, Harvard University, Cambridge, MA 02138, USA6Research Laboratory of Electronics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA

7Center for Theoretical Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA8Institute for Physical Science and Technology, University of Maryland, College Park, MD 20742, USA

Quantifying quantum states’ complexity is a key problem in various subfields of science, from quantumcomputing to black-hole physics. We prove a prominent conjecture by Brown and Susskind about how randomquantum circuits’ complexity increases. Consider constructing a unitary from Haar-random two-qubit quantumgates. Implementing the unitary exactly requires a circuit of some minimal number of gates—the unitary’sexact circuit complexity. We prove that this complexity grows linearly in the number of random gates, withunit probability, until saturating after exponentially many random gates. Our proof is surprisingly short, giventhe established difficulty of lower-bounding the exact circuit complexity. Our strategy combines differentialtopology and elementary algebraic geometry with an inductive construction of Clifford circuits.

Quantum complexity is a pervasive concept at the inter-section of quantum computing, quantum many-body systems,and black-hole physics. In general, complexity quantifies theresources required to implement a computation. For example,a Boolean function’s complexity can be defined as the mini-mal number of gates, chosen from a given gate set, necessaryto evaluate the function. In quantum computing, the circuitmodel provides a natural measure of complexity: A unitarytransformation’s complexity is the size of the smallest circuitthat effects the unitary. Similarly, a pure state has a complex-ity definable as the size of the smallest circuit that producesthe state from a product state.

Quantum complexity has applications in numerous sub-fields of quantum physics, including the foundations of quan-tum computing. Quantum computational complexity is rootedin circuit complexity: A problem is deemed “easy” if it issoluble with a circuit whose size grows polynomially withthe input’s size. The problem is deemed “hard” if the cir-cuit’s size scales exponentially. Quantum complexity alsofeatures in the definition of phases of matter. For instance,a topological phase is defined in terms of a high-complexityquantum state. Recently, close connections have been discov-ered between gate complexity and holography in high-energyphysics, in the context of the anti-de-Sitter-space/conformal-field-theory (AdS/CFT) correspondence [7, 10, 12, 52, 55].The “complexity equals volume” conjecture [52] suggests thatthe correspondence’s boundary state has a complexity propor-tional to the volume behind the event horizon of a black holein the bulk geometry. Similarly, the “complexity equals ac-tion” conjecture posits that a holographic state’s complexityis dual to a certain space-time region’s action [11]. Thesedeep physical connections have motivated studies of quan-tum complexity as a means of illuminating quantum many-body systems’ complex behaviors. Lower-bounding the quan-tum complexity is a long-standing open problem in quan-tum information theory. Underlying this difficulty is thatgates performed early in a circuit may cancel with gates per-formed later. Consequently, quantum-gate–synthesis algo-

rithms, which decompose a given unitary into gates, run fortimes exponential in the system size [1]. Several attemptshave been made to lower-bound unitaries’ circuit complexi-ties as defined geometrically by Nielsen [18, 19, 38, 40, 41].A few claims have been proved about circuit complexity; forinstance, polynomially deep circuits implement only a smallfraction of the unitaries. Most unitaries have complexitiesexponentially large in the system size, according to a count-ing argument [30]. A key question in the study of quantumcomplexity is the following. Consider constructing deeperand deeper circuits for an n-qubit system, by applying ran-dom two-qubit gates. At what rate does the circuit complex-ity increase? Brown and Susskind conjectured that quantumcircuits’ complexity generically grows linearly for an expo-nentially long time [12, 56]. Brown and Susskind supportedthe conjecture using the complexity geometry introduced byNielsen [38, 41]. Further evidence for the conjecture arisesfrom counting arguments [49]. Growth-of-complexity studiesadvanced further in [9], which connected a strong notion ofquantum complexity to unitary t-designs (finite collections ofunitaries that approximate completely random unitaries [42]).

Our results govern a straightforward notion of circuit com-plexity, a notion based on exact implementations of a uni-tary [18, 38, 41]. The unitaries under consideration are con-structed from Haar-random two-qubit gates. We prove that thecircuit complexity grows linearly with time (with the num-ber of gates applied). The proof is surprisingly short, giventhe difficulty of lower-bounding complexity. Our bound onthe complexity holds for all random circuits described above,with probability 1. Instead of invoking unitary designs [9] orNielsen’s geometric approach [18, 38, 40, 41], we employ el-ementary aspects of differential topology and algebraic geom-etry, combined with an inductive construction of Clifford cir-cuits. Clifford circuits play a pivotal role in quantum comput-ing, as circuits that can easily be implemented fault-tolerantly.

This work is organized as follows. First, we introduce thesetup and definitions. Second, we present the main result, thecomplexity’s exponentially long linear growth. We present a

arX

iv:2

106.

0530

5v1

[qu

ant-

ph]

9 J

un 2

021

2

Figure 1. The complexity ramp: Evolution of quantum complexity,as conjectured in Ref. [12].

high-level overview of the proof third. The key mathematicalsteps follow, in the methods section. Two corollaries follow:an extension to random arrangements of gates and an exten-sion to slightly imperfect gates. In the discussion, we compareour results with known results and explain our work’s impli-cations for various subfields of quantum physics. Finally, wediscuss the opportunities engendered by this work.

In Appendix A, we review elementary algebraic geometryrequired for the proof. Proof details appear in Appendix B. Weelaborate on states’ complexities in Appendix C. We prove thetwo corollaries in Appendices D and E. Finally, we comparenotions of circuit complexity in Appendix F.

Preliminaries. This work concerns a system of n qubits.For convenience, we assume that n is even. We simplifytensor-product notation as |0k〉 := |0〉⊗k, for k = 1, 2, . . . , n;and 1k denotes the k-qubit identity operator. Let Uj,k denotea unitary gate that operates on qubits j and k. Such gates neednot couple the qubits together and need not be geometricallylocal.

An architecture is an arrangement of some fixed number Rof gates. Formally:

Definition 1. An architecture is a directed acyclic graph thatcontains R ∈ Z>0 vertices (gates). Two edges (qubits) entereach vertex, and two edges exit.

Figures 2(b) and 2(c) illustrate example architectures gov-erned by our results.

• A brickwork is the architecture of any circuit formedas follows: Apply a string of two-qubit gates: U1,2 ⊗U3,4 ⊗ . . .⊗ Un−1,n. Then, apply a staggered string ofgates, as shown in Fig. 2(b). Perform this pair of stepsT times total, using possibly different gates each time.

• A staircase is the architecture of any circuit formed asin Fig. 2(c): Apply a stepwise string of two-qubit gates:Un,n−1Un−2,n−1 . . . U2,1. Repeat this process T times,using possibly different gates each time.

The total number of gates in the brickwork architecture, as inthe staircase architecture, isR = (n−1)T . Our results extendto more-general architectures, e.g., the architecture depictedin Fig. 2(a) and architectures of non-nearest-neighbor gates.Circuits of a given architecture can be formed randomly.

Definition 2 (Random quantum circuit). Let A denote an ar-bitrary architecture. A probability distribution can be inducedover the architecture-A circuits as follows: For each vertex inA, draw a gate Haar-randomly from SU(4). Then, contractthe unitaries along the edges ofA. Each circuit so constructedis called a random quantum circuit.

Implementing a unitary with the optimal gates, in the optimalarchitecture, concretizes the notion of complexity.

Definition 3 (Exact circuit complexities). Let U ∈ SU(2n)denote an n-qubit unitary. The (exact) circuit complexityCu(U) is the least number of two-qubit gates in any circuitthat implements U . Similarly, let |ψ〉 denote a pure quantumstate vector. The (exact) state complexity Cstate(|ψ〉) is theleast number r of two-qubit gates U1, U2, . . . , Ur, arrangedin any architecture, such that U1U2 . . . Ur|0n〉 = |ψ〉.

One more concept merits defining: a causal slice. Considercreating two vertical cuts in a circuit (dashed lines in Fig. 2).The gates between the cuts form a slice. The slice is causalif, viewed as a graph, it contains a complete backwards lightcone: Some qubit t links to each other qubit t′ via a directedpath of gates (a path that may be unique to t′). A causal sliceis, in a sense, well-connected. A slice need not be a timeslice: The gates in a slice need not all be simultaneously im-plementable.

Main result: Linear growth of complexity in random quan-tum circuits. Our main result is a lower bound on the complex-ities of random unitaries and states. The bound holds with unitprobability.

Theorem 1 (Linear growth of complexity). Let U denote aunitary implemented by a random quantum circuit in an ar-chitecture formed from T = R/L causal slices of L gateseach. The unitary’s circuit complexity is lower-bounded as

Cu(U) ≥ R

9L− n

3, (1)

with unit probability, until the number of gates grows to T =R/L ≥ 4n−1. The same bound holds for Cstate(U |0n〉), untilT = R/L ≥ 2n+1 − 1.

The theorem governs all architectures formed from causalslices. The brickwork architecture forms a familiar specialcase. Let us choose for a brickwork’s causal slice to con-tain 2n time slices, or 2n of the columns in Fig. 2(b). Eachcausal slice contains L = n(n − 1) gates (in the absenceof periodic boundary conditions), yielding the lower boundCu(U) ≥ R

9n(n−1) −n3 . Another familiar example is the stair-

case architecture. A staircase’s causal slices can have the leastL possible, n− 1, which yields the strongest bound.

High-level overview of the proof of Theorem 1. Considerfixing anR-gate architectureA, then choosing the gates in thearchitecture. The resulting circuit implements some n-qubitunitary. All the unitaries implementable with A form a setU(A). Our proof relies on U(A)—namely, on the number ofdegrees of freedom in U(A). We define this number as the ar-chitecture’s accessible dimension, dA = dim(U(A)) (Fig. 3).The following section contains a formal definition; here, we

3

(a) (b) (c)

Figure 2. Our result relies on architectures and their causal slices. (a) An architecture specifies howR 2-qubit gates are arranged in an n-qubitcircuit. The gates need not be applied to neighboring qubits, though they are depicted this way for convenience. A causal slice is a section ofthe architecture (between the gray dotted lines) with the following property: There exists a qubit reachable from each other qubit via a path(red dashed line), possibly unique to the latter qubit, that passes only through gates in the slice. (b) The brickwork architecture interlaces layersof gates on a one-dimensional (1D) chain. In a 1D architecture with geometrically local gates, such as the brickwork architecture, each causalslice has an inverted light cone (light red region) that touches the qubit chain’s edges. A minimal causal slice in the brickwork architecturecontains ∼n2 gates. (c) The staircase architecture, too, acts on a 1D qubit chain. The circuit consists of layers in which n − 1 gates act onconsecutive qubit pairs. A minimal causal slice contains n− 1 gates.

provide intuition. As the n-qubit unitaries form a space ofdimension 4n, dA ∈ [0, 4n]. The greater the dA, the morespace U(A) fills in the set of n-qubit unitaries. Our first tech-nical result lower-bounds sufficiently connected architecture’saccessible dimension:

Proposition 1. Let AT denote an architecture with R = TLgates. Assume that AT consists of causal slices of ≤ Lgates each. The architecture’s accessible dimension is lower-bounded as

dAT ≥ T =R

L. (2)

We can upper-bound dA, for an arbitrary architecture A,by counting parameters. To synopsize the argument in Ap-pendix B: Fifteen real parameters specify each 2-qubit uni-tary. Each qubit shared by two unitaries makes 3 parametersredundant. Hence

dA ≤ 9R+ 3n . (3)

The accessible dimension reaches its maximal value, 4n, af-ter a number of gates exponential in n. Similarly, the cir-cuit complexity reaches its maximal value after exponentiallymany gates. This parallel suggests dA as a proxy for the cir-cuit complexity. The next section rigorously justifies the useof dA as a proxy.

The proof of Theorem 1 revolves around the accessible di-mension dAT of a certain R-gate architecture AT . The mainidea is as follows. Let R′ be less than a linear fraction ofR. More specifically, let 9R′ + 3n < T = R/L. For everyR′-gate architecture A′, dA′ < dAT holds by a combinationof (2) and (3). Consequently, Appendix B shows, U(A′) haszero probability in U(AT ), according to the measure in Def-inition 2. Therefore, almost every unitary U ∈ U(AT ) hasa complexity greater than the greatest possible R′. Inequal-ity (1) follows.

Methods. Having overviewed the proof at a high level,we fill in the key mathematics. Three points need clarifying.First, we must rigorously define the accessible dimension, orthe dimension of U(A), which is not a manifold. Second, wemust prove Proposition 1. Finally, we must elucidate steps inthe proof of Theorem 1. We address these points using thetoolbox of algebraic geometry.

Architecture Image

Contraction map

Choices ofunitary gates

Figure 3. The R-gate architecture A is associated with a contractionmap FA. FA maps a list of input gates (a point in [SU(4)]×R) to ann-qubit unitary U in SU(2n). The unitary results from substitutingthe gates into the architecture. FA has an image U(A), which con-sists of the unitaries implementable with the architecture. A has anaccessible dimension, dA, equal to the dimension of U(A). Our coretechnical result is that dA grows linearly with R. To bridge this re-sult to complexity, consider an arbitrary architectureA′ formed fromfewer gates than a constant fraction of R. Such an architecture’s ac-cessible dimension satisfies dA′ < dA, we show. Therefore, everyunitary in U(A) has a complexity linear inR, except for a measure-0set. The proof relies on algebraic geometry. A key concept is therank of FA at a point. The rank counts the local degrees of freedomin the image (orange arrows).

We associate with every R-gate architecture A a contrac-tion map FA : SU(4)×R → SU(2n). This function mapsa list of gates to an n-qubit unitary. The unitary results fromsubstituting the gates into the architectureA (Fig. 3). The mapcontracts every edge (qubit) shared by two vertices (gates) inA.

The image of FA is the set U(A) of unitaries imple-mentable with the architectureA. U(A) is a semialgebraic set,consisting of the solutions to a finite set of polynomial equa-tions and inequalities over the real numbers (see Appendix Afor a review). That U(A) is a semialgebraic set follows fromthe Tarski-Seidenberg principle, a deep result in semialge-braic geometry (Appendix A). A semialgebraic set’s dimen-sion quantifies the degrees of freedom needed to describe theset locally. More precisely, a semialgebraic set decomposesinto manifolds. The greatest dimension of any such manifoldequals the semialgebraic set’s dimension. More restricted than

4

a semialgebraic set is an algebraic set, which consists of thesolutions to a finite set of polynomial equations.

Just as the contraction map’s image will prove useful,so will the map’s rank, defined as follows. Let x =(U1, U2, . . . , UR) ∈ SU(4)×R denote an input into FA, suchthat the Uj denote two-qubit gates. The map’s rank at x isthe rank of a matrix that approximates FA linearly around x(the rank of the map’s Jacobian at x). The rank is low at x ifperturbing x can influence the n-qubit unitary only along fewdirections in SU(2n).

Crucially, we prove that FA has the same rank throughoutthe domain, except on a measure-zero set, where FA has alesser rank. The greater, “dominating” rank is the dimensionof U(A). To formalize this result, let Er denote the locusof points at which FA has a rank of r ≥ 0. Let E<r =⋃r′<r Er′ denote the set of points where FA has a lesser rank.

Let rmax denote the maximum rank achieved by FA at anypoint x. We prove the following lemma in Appendix B, usingthe dimension theory of real algebraic sets.

Lemma 1 (Low-rank locus). The low-rank locusE<rmax is analgebraic set of measure 0 and so is closed (in the Lie-grouptopology). Equivalently, Ermax is an open set of measure 1.Consequently, dA = rmax.

Lemma 1 guarantees that the contraction map’s rank equalsthe accessible dimension dA almost everywhere in U(A).

We now turn to the proof of Proposition 1. The rank r ofFA at each point x lower-bounds rmax, by definition. For anarchitectureAT of T causal slices, we identify an x at which ris lower-bounded by a quantity that grows linearly withR (thenumber of gates in the architecture AT ). We demonstrate thepoint’s existence by constructing circuits from Clifford gates.

Consider a choice x = (U1, U2, . . . , UR) ≡ (Uj)j of uni-tary gates. Perturbing a Uj amounts to appending an infinites-imal unitary: Uj 7→ Uj = eiεHUj . The H denotes a 2-qubitHermitian operator, and ε ∈ R. H can be written as a linearcombination of 2-qubit Pauli strings Sk. (An n-qubit Paulistring is a tensor product of n single-site operators, each ofwhich is a Pauli operator [X , Y , or Z] or the identity, 11.The 4n n-qubit Pauli strings form a basis for the space of n-qubit Hermitian operators.) Consider perturbing each gate Ujusing a combination of all 15 nontrivial 2-qubit Pauli strings[Fig. 4(a)]: x = (Uj)j 7→ x = (exp(i

∑15k=1 εj,kSk)Uj)j ,

wherein εj,k ∈ R. The perturbation x 7→ x causes a pertur-bation U = FAT (x) 7→ U = FAT (x) of the image underFAT . The latter perturbation is, to first order, ∂εj,k U

∣∣εj,k=0

.This derivative can be expressed as the original circuit withthe Pauli string Sk inserted immediately after the gate Uj[Fig. 4(b)].

The rank of FAT at x is the number of parametersεj,k needed to parameterize a general perturbation of U =FAT (x) within the image set U(AT ). To lower-bound therank of FAT at a point x, we need only show that ≥ r param-eters εj,k perturb FAT (x) in independent directions. To do so,we express the derivative as

∂εj,kFAT (x)

∣∣εj,k=0

= Kj,kFAT (x) , (4)

wherein Kj,k denotes a Hermitian operator [Fig. 4(c)]. Kj,k

results from conjugating Sk, the Pauli string inserted into thecircuit after gate Uj , with the later gates. The physical signifi-cance ofKj,k follows from perturbing the gateUj in the direc-tion Sk by an infinitesimal amount εj,k. The image FAT (x)is consequently perturbed, in SU(2n), in the direction Kj,k.

We choose for the gates Uj to be Clifford operators. (TheClifford operators are the operators that map the Pauli stringsto the Pauli strings, to within a phase, via conjugation. For ev-ery Clifford operator C and Pauli operator P , CPC† equalsa phase times a Pauli string [5, 6, 13, 14, 22].) As a result,the operators Kj,k are Pauli strings (up to a phase). Two Paulistrings are linearly independent if and only if they differ. ForClifford circuits, therefore, we can easily verify whether per-turbations of x cause independent perturbation directions inSU(2n): We need only show that the resulting operators Kj,k

are distinct.We apply that fact to prove Proposition 1, using the follow-

ing observation. Consider any Pauli string P and any causalslice of any architecture. We can insert Clifford gates intothe slice such that two operations are equivalent: (i) operatingon the input qubits with P before the extended slice and (ii)operating with the extended slice, then with a one-qubit Z.Figure 4(d) depicts the equivalence, which follows from thestructure of causal slices. We can iteratively construct a Clif-ford unitary that reduces the Pauli string’s weight until pro-ducing a single-qubit operator. See Appendix B for details.

We now prove Proposition 1 by recursion. Consider an R′-gate architecture AT ′ formed from T ′ causal slices of ≤ Lgates each. Let x′ denote a list of Clifford gates slotted intoAT ′ . Assume that FAT ′ (x′) has a rank ≥ T ′. Consider ap-pending a causal slice to AT ′ . The resulting architecture cor-responds to a contraction map whose rank is ≥ T ′ + 1, weshow.

By assumption, we can perturb x′ such that its imageFAT ′ (x′) is perturbed in ≥ T ′ independent directions inSU(2n). These directions can be represented by Pauli oper-ators K ′jm,km , wherein m = 1, 2, . . . , T ′, by Eq. (4). Let Pdenote any Pauli operator absent from K ′jm,km. We can ap-pend to AT ′ a causal slice, forming an architecture AT ′+1 ofT ′+1 causal slices. We design the new causal slice from Clif-ford gates such that two operations are equivalent: (i) apply-ing P to the input qubits before the extended slice and (ii) ap-plying the extended slice, then a single-site Z. We denoteby x′′ the list of gates in x′ augmented with the gates in theextended slice. Conjugating the K ′jm,km with the new sliceyields operators K ′′jm,km , for m = 1, 2, . . . , T ′. They rep-resent the directions in which the image FAT ′+1(x′′) is per-turbed by the original perturbations of AT ′ . The K ′′jm,km arestill linearly independent Pauli operators. Also, the K ′′jm,kmand the single-site Z form an independent set, because P isnot in K ′jm,km. Meanwhile, the single-site Z is a direc-tion in which the slice’s final gate can be perturbed. The op-erators Kjm,km , augmented with the single-site Z, thereforespan T ′ + 1 independent directions along which FAT ′+1(x′′)can be perturbed. Therefore, T ′ + 1 lower-bounds the rank ofFAT ′+1 .

We apply the above argument recursively, starting from

5

(a) (b)

(c)(e)

1

2j j

causal slices

causal slices

(d)

Figure 4. Our core technical result is a lower bound on the accessible dimension (see Fig. 3). We prove this bound using a construction basedon Clifford circuits. (a) Each gate Uj is perturbed with a unitary eiεj,kSk , generated by a 2-qubit Pauli operator Sk and parameterized withan infinitesimal εj,k ∈ R. Perturbing the gate perturbs the n-qubit unitary, turning U into U ≈ U . (b) A key quantity is the derivative of Uwith respect to a parameter εj,k, evaluated at U . Taking this derivative is equivalent to inserting the Pauli string Sk immediately after the gateUj . (c) The derivative depicted in panel (b) is equivalent to following the circuit with a Hermitian operator Kj,k [Eq. (4)]. The operator Kj,k

results from conjugating Sk with the gates after Uj . If the circuit consists of Clifford gates, thenKj,k is a Pauli string, since Clifford gates mapthe Pauli strings to Pauli strings. Therefore, a perturbation of Uj in the direction of Sk results in a perturbation of the resulting unitary U inthe direction of Kj,k in SU(2n). (d) The following is true of every causal slice and every Pauli string P (leftmost green squares): The slice’sgates can be chosen to be Cliffords that map P to a single-site Z. The Clifford gates first map P to a Pauli string that acts nontrivially onfewer qubits (pale green squares), then to a Pauli string on fewer qubits, and so on until the Pauli string dwindles to one qubit (rightmost greensquare). (e) Our lower bound is proven by recursion. Consider an architecture AT ′ , formed from T ′ causal slices, whose accessible dimensionis ≥ T ′. There exist gates U1, U2, . . . , UR′ such that that T ′ linearly independent Pauli operators K′jm,km (wherein m = 1, 2, . . . , T ′) resultfrom perturbing the gates, as described in (a)–(c). Consider a Pauli operator P that is not in K′jm,km. We can append to AT ′ a causal slice,formed from Clifford gates, that maps P to a single-site Z, as depicted in panel (d). This Z is an important direction in SU(2n): Considerperturbing the causal slice’s final gate via the procedure in (a)–(c). The image U(AT ′) is perturbed, as a result, in the direction Z. Thus,T ′ + 1 linearly independent Pauli operators (the operators K′jm,km and P ) result from perturbing gates in the extended circuit. Therefore, theextended circuit’s accessible dimension is ≥ T ′ + 1.

an architecture that contains no gates. The following resultemerges: Consider any architecture AT that consists of Tcausal slices. At some point x, the map FAT has a rank lower-bounded by T . Lemma 1 ensures that the same bound appliesto dAT .

To conclude the proof of Theorem 1, we address an archi-tecture A′ whose accessible dimension satisfies dA′ < dAT .Consider sampling a random circuit with the architecture AT .We must show that the circuit has a zero probability of im-plementing a unitary in U(A′). To prove this claim, we in-voke the constant-rank theorem: Consider any map whoserank is constant locally—in any open neighborhood of anypoint in the domain. In that neighborhood, the map is equiva-lent to a projector, up to a diffeomorphism. We can apply theconstant-rank theorem to the contraction map: FAT has a con-stant rank throughout Ermax

, by Lemma 1. Therefore, FATacts locally as a projector throughout Ermax

—and so through-out SU(4)×R, except on a measure-0 region, by Lemma 1.Consider mapping an image back, through a projector, to apreimage. Suppose that the image forms a subset of dimen-sion lower than the whole range’s dimension. The backward-mapping just adds degrees of freedom to the image. There-fore, the preimage locally has a dimension less than the do-main’s dimension. Hence the preimage is of measure 0 in the

domain. We use the unitary group’s compactness to elevatethis local statement to the global statement in Theorem 1.

Extensions. This section concerns two extensions of Theo-rem 1, to random architectures and to approximate implemen-tations of unitaries. A natural question is whether the com-plexity grows linearly if the architecture is drawn randomly,as in Ref. [8]. We answer this question affirmatively in Ap-pendix D. A bound similar to Ineq. (1) holds with a probabilityexponentially (in system size) close to one:

Corollary 1 (Randomized architectures). Consider drawingan n-qubit unitary U according to the following probabilitydistribution: Choose a qubit j uniformly randomly. Apply aHaar-random two-qubit gate to qubits j and j + 1. Performthis process R times. With high probability, the unitary imple-mented has a high complexity: For all α ∈ [0, 1),

Pr

(Cu(U) ≥ α R

9n(n− 1)2− n

3

)≥ 1− 1

1− α(n−1)e−n .

(5)

We prove this corollary in Appendix D. The key is, if weplace many gates at random locations, many causal sliceslikely emerge. No such bound can hold with probability one:There is a nonzero probability of drawing “trivial” architec-tures, in which, e.g., the same two qubits undergo all the gates.

6

Our second extension of Theorem 1 centers on exact circuitcomplexity. The theorem accommodates no errors in the im-plementation of a unitary, but unitaries are not implementedperfectly in reality. Therefore, the exact complexity merits anextension to approximations. The complexity grows linearlyin time, we prove, even if the circuit approximates the targetunitary U with a small error. We quantify the error using theFrobenius norm, ‖X‖F = [tr(X†X)]1/2. The next corollarycan be synopsized as follows: For any δ > 0, we can finda ε > 0 such that an ε-robust version of complexity growslinearly with circuit size.

Corollary 2 (Slightly robust circuit complexity). Let U de-note the n-qubit unitary implemented by any random quantumcircuit in any architectureAT that satisfies the assumptions inTheorem 1. Let U ′ denote the n-qubit unitary implemented byany circuit ofR′ ≤ R/(9L)−n/3 gates. For every δ ∈ (0, 1],there exists an ε := ε(AT , δ) > 0 such that ||U − U ′||F ≥ ε,with probability 1− δ, unless R/L > 4n − 1.

We prove this corollary in Appendix E, using the proof of The-orem 1. We cannot control the dependence of ε on δ or onAT .Therefore, we cannot control the dependence of ε on n. Ac-cordingly, the error tolerance of Theorem 1 may be extremelysmall.

Discussion. We have proven a prominent physics conjec-ture proposed by Brown and Susskind for random quantumcircuits [12, 56]: A local random circuit’s quantum complex-ity grows linearly in the number of gates until reaching a valueexponential in the system size. To prove this conjecture, weintroduced a novel technique for bounding complexity. Theproof rests on our connecting the quantum complexity to theaccessible dimension, the dimension of the set of unitaries im-plementable with a given architecture (arrangement of gates).Our core technical contribution is a lower bound on the acces-sible dimension. The bound rests on techniques from differ-ential topology and algebraic geometry.

To the best of our knowledge, Theorem 1 is the first rigor-ous demonstration of the linear growth of random qubit cir-cuits’ complexities for exponentially long times. The boundholds until the complexity reaches Cu(U) = Ω(4n)—thescaling, up to polynomial factors, of the greatest complexityachievable by any n-qubit unitary [39]. A hurdle has stymiedattempts to prove that local random circuits’ quantum com-plexity grows linearly: Most physical properties (describedwith, e.g., local observables or correlation functions) reachfixed values in times subexponential in the system size. Onemust progress beyond such properties to prove that the com-plexity grows linearly at superpolynomial times. We over-come this hurdle by identifying the accessible dimension as aproxy for the complexity.

Theorem 1 complements another rigorous insight aboutcomplexity growth. In Ref. [9], the linear growth of complex-ity is proven in the limit of large local dimension q and for astrong notion of quantum circuit complexity, with help fromRef. [28]. Furthermore, depth-T random qubit circuits havecomplexities that scale as Ω(T 1/11) until T = exp(Ω(n)) [8,9]. The complexity scales the same way for other types ofrandom unitary evolutions, such as a continuous-time evolu-

tion under a stochastically fluctuating Hamiltonian [48]. Fi-nally, Ref. [9] addresses bounds on convergence to unitarydesigns [8, 24, 25, 28, 36, 48], translating these bounds intoresults about circuit complexity. Theorem 1 is neither strongernor weaker than the results of Ref. [9], which govern a moreoperational notion of complexity—how easily U |0n〉〈0n|U†can be distinguished from the maximally mixed state.

Our proof hinges on our architectures’ connectedness—onan architecture’s consisting of causal slices. One might find analternative proof technique to bypass this assumption. How-ever, a notion related to causal slices appears in the switchbackeffect in holography [12, 53, 57]. Consider producing a stateby inputting |0n〉 into a generic quantum circuit. Now, con-sider sequentially applying the gates’ inverses, starting withthe last gate’s inverse. The state’s complexity is expected todecrease until reaching zero. Suppose, however, that a qubitis perturbed during this process. The state’s complexity is nolonger expected to decrease to zero. Instead, the complexitywill start increasing again. The delay between the perturba-tion and the complexity’s return to increasing is the switch-back effect. It is of the order of the time necessary for theperturbation to grow, in the Heisenberg picture, to reach thefull system size. A causal slice is defined strikingly similarly:The slice enables a qubit, if time is reversed, to affect all thequbits at the causal slice’s input. This similarity indicates thatour definition of a causal slice, used here as a proof technique,might have broader relevance.

Outlook. Our main result governs exact circuit complexity.Corollary 2 generalizes the result to approximate circuit com-plexity. There, the complexity depends on our tolerance of theerror in the implemented unitary. Yet, the error tolerance canbe uncontrollably small in Corollary 2. The main challengein extending our results to approximate complexity is, the ac-cessible dimension crudely characterizes the set of unitariesimplementable with a given architecture. Consider attemptingto enlarge this set to include all the n-qubit unitaries that lieclose to the set in some norm. The enlarged set’s dimensionis 4n. The reason is, the enlargement happens in all direc-tions of SU(2n). Therefore, our argument does not work asfor the exact complexity. Extending our results to approxima-tions therefore offers an opportunity for future work. Approx-imations may also illuminate random circuits as instrumentsfor identifying quantum advantages [3, 37]; they would showthat a polynomial-size quantum circuit cannot be compressedsubstantially while achieving a good approximation.

Another opportunity involves Nielsen’s geometric notionof complexity [18, 38, 40, 41], which underlies Brown andSusskind’s conjecture. Nielsen’s notion originated as a toolfor lower-bounding the circuit complexity and opens up thetoolbox of Riemannian geometry. Nielsen’s complexity ge-ometry is related to circuit complexity conceptually and quan-titatively [18, 41]. Yet our results do not immediately lower-bound the cost. The reason is, the exact circuit complexitydoes not lower-bound Nielsen’s cost. Furthermore, the costinvolves how close gates lie to the identity operation (see Ap-pendix F). The definition of exact circuit complexity does not,instead counting the gates needed to implement U . Conse-quently, attempting to use our lower bound on discretized

7

paths in Nielsen’s geometry would yield a bound that dependson the discretization’s resolution.

These observations motivate an uplifting of the presentwork to robust notions of quantum circuit complexity (see,e.g., Ref. [9]). A possible uplifting might look as follows.Let A denote an R-gate architecture, and let A′ denote anR′-gate architecture. Suppose that the accessible dimensionsobey dA′ < dA. Our result relied on how a unitary imple-mented with A has a zero chance of occupying the set U(A′),which has smaller dimension than U(A). Consider enlargingU(A′) to include the unitaries that lie ε-close, for some ε > 0.If U(A′) is sufficiently smooth and well-behaved, we expectthe enlarged set’s volume, intersected with U(A), to scale as∼ εdA−dA′ . Furthermore, suppose that unitaries implementedwith A are distributed sufficiently evenly in U(A) [rather thanbeing concentrated close to U(A′)]. All the unitaries in U(A)except a small fraction ∼ εdA−dA′ would not be able to lie inU(A′). We expect, therefore, that all the unitaries in U(A)except a fraction ∼ εdA−dA′ have ε-approximate complexitiesgreater than R′.

While our results focus on circuits for qubits, we anticipateno significant challenges in extending our results to circuitsfor qudits of finite dimension D > 2. The complexity lowerbound might scale unfavorably in D, however. The reason is,our complexity bound’s coefficient stems from the dimensionof the group SU(4), which would be replaced with SU(D2).As the relevant group’s dimension grows, the coefficient—andthe lower bound—shrinks. However, we expect the resultingbound to scale linearly with the number of gates. The mainchallenge, we anticipate, is to generalize our Clifford-circuitconstruction, outlined in Fig. 4(e). An alternative approachwould be to embed each qudit in log2(D) qubits and considergates that act jointly on log2(D) qubits.

The most profound implications of our work are expectedto arise in the holographic context surrounding the Brown-Susskind conjecture. There, random quantum circuits are con-jectured to serve as proxies for chaotic quantum dynamicsgenerated by local time-independent Hamiltonians [35]. Ref-erence [27] introduced this conjecture into black-hole physics,and Ref. [55] discussed the conjecture in the context of holog-raphy. A motivation for invoking random circuits is, randomcircuits can be analyzed more easily than time-independent–Hamiltonian dynamics. Time-independent–Hamiltonian dy-

namics are believed to be mimicked also by time-fluctuatingHamiltonians [48] and by random ensembles of Hamilto-nians. Random ensembles have enabled progress in thestudy of scrambling, or the spread of local perturbationsthrough many-body entanglement, via the Sachdev-Ye-Kitaevmodel [12, 32, 51]. Studying random circuits’ complexityis expected to illuminate the complexity of chaotic Hamil-tonian dynamics similarly. Furthermore, complexity partici-pates in analogies with thermodynamics, such as a second lawof quantum complexity [12]. Our techniques can be leveragedto construct an associated resource theory of complexity [20].

In the context of holography, thermofield double states’complexities have attracted recent interest [15, 33, 54, 55].Thermofield double states are pure bipartite quantum statesfor which each subsystem’s reduced state is thermal. In thecontext of holography, thermofield double states are dual toeternal black holes in anti-de-Sitter space [33]. Such a blackhole’s geometry consists of two sides connected by a worm-hole, or Einstein-Rosen bridge. The wormhole’s volumegrows for a time exponential in the number of degrees of free-dom of the boundary theory [12, 55]. As discussed above, ran-dom quantum circuits are expected to capture the (presumedHamiltonian) dynamics behind the horizon. If they do, thegrowth of the wormhole’s volume is conjectured to match thegrowth of the boundary state’s complexity [12, 52, 55]; bothare expected to reach a value exponentially large in the num-ber of degrees of freedom. Our results govern the randomcircuit that serves as a proxy for the dynamics behind the hori-zon. That random circuit’s complexity, our results show strik-ingly, indeed grows to exponentially large values. We hopethat the present work, by innovating machinery for addressingcomplexity, stimulates further quantitative studies of hologra-phy, scrambling, and chaotic quantum dynamics.

Acknowledgements. We thank Péter Varjú and Aram Har-row for discussions that inspired this approach to complex-ity growth. N. Y. H. thanks Shira Chapman, Michael Walter,and the other organizers of the 2020 Lorentz Center workshop“Complexity: From quantum information to black holes” forinspiration. This work has been funded by the DFG (EI519/14-1, CRC 183, for which this is an inter-node Berlin-Cologne project) and by an NSF grant for the Institute forTheoretical Atomic, Molecular, and Optical Physics at Har-vard University and the Smithsonian Astrophysical Observa-tory. Administrative support was provided by the MIT CTP.

Appendix A: Algebraic and semialgebraic sets

For convenience, we review elementary aspects of algebraic geometry over the real numbers. We apply these properties in theproof of Theorem 1. Ref. [4] contains a more comprehensive treatment.

Definition 4 (Algebraic sets). A subset V ⊆ Rm is called an algebraic set, or an algebraic variety, if, for a set of polynomialsfjj ,

V = x ∈ Rm|fj(x) = 0. (A1)

A subset V ′ ⊆ V is called an algebraic subset if V ′ is an algebraic set. We call a subset W ⊆ Rm a semialgebraic set if, for setsfjj and gkk of polynomials,

W = x ∈ Rm|fj(x) = 0, gk ≤ 0. (A2)

8

A natural topology on algebraic sets is the Zariski topology.

Definition 5 (Zariski topology). Let V denote an algebraic set. The Zariski topology is the unique topology whose closed setsare the algebraic subsets of V .

A traditional definition of “dimension” for algebraic sets involves irreducible sets.

Definition 6 (Irreducible sets). Let X denote a topological space. X is called irreducible if it is not the union of two properclosed subsets.

Definition 7 (Dimension of algebraic sets). Let V be an algebraic set that is irreducible with respect to the Zariski topology.The dimension of V is the maximal length d of any chain V0 ⊂ V1 ⊂ · · · ⊂ Vd of distinct nonempty irreducible algebraic subsetsof V .

The relevant algebraic sets in the proof of Theorem 1 are SU(4)×R and SU(2n). Our interest in semialgebraic sets stems fromthe following principle. In the following, we refer to a function F : Rn → Rm as a polynomial map if its entries are polynomialsin the entries of its input.

Theorem 2 (Tarski-Seidenberg principle). Let F : Rn → Rm be a polynomial map. If W is a semialgebraic set, so is F (W ).

The Tarski-Seidenberg principle applies to the map that contracts sets of quantum gates. This application is important for us,because it provides a natural notion of dimension for the contraction map’s image.

All semialgebraic sets (and hence all algebraic sets) decompose into smooth manifolds.

Theorem 3 (Stratification of semialgebraic sets). If W is a semialgebraic set, then W =⋃Ni=jMj , wherein each Mj denotes a

smooth manifold. If W is an algebraic set of dimension d in the sense of Definition 7, then maxjdim(Mj) = d.

This maxjdim(Mj) does not depend on the decomposition chosen. Therefore, this maximum is defined as the dimension ofW . Two interrelated semialgebraic sets have interrelated dimensions.

Lemma 2 (Dimension of an image). Let F : Rn → Rm be a polynomial map. If W is a dimension-d semialgebraic set, F (W )is of dimension ≤ d.

The bound follows from combining the results of Ref. [4, Prop. 2.8.7] with the results of Ref. [4, Prop. 2.8.6]. (Ref. [4] invokesa semialgebraic mapping, which encompasses polynomial maps.)

Appendix B: Proof of the main theorem and lemmata

In this appendix, we prove Lemma 1, Lemma 3, and the main theorem. The proofs rely on the topics reviewed in Appendix A,as well as the following notation and concepts. In differential geometry, the rank of FA at the point x = (U1, U2, . . . , UR)is defined as the rank of the derivative DxF

A. Mapping lists of gates to unitaries, F is a complicated object. We can moreeasily characterize a map from real numbers to real numbers. Related is a map from Hermitian operators to Hermitian operators:n qubits occupy a state represented by a 2n × 2n Hermitian operator, which has (2n)2 = 4n real parameters. Therefore, forconvenience, we shift focus from unitaries to their Hermitian generators. We construct a map whose domain is the algebrasu(4)×R ' R15R that generates SU(4)×R. The range is the set of n-qubit Hermitian operators, su(2n) ' R4n . We constructsuch a map from three steps, depicted by the dashed lines in Fig. 5.

The first step is a chart, a diffeomorphism that maps one manifold to another invertibly. Our chart acts on the algebra su(4)×R

that generates SU(4)×R. To define the chart, we parameterize an element H of the jth copy of su(4):

H =∑

α,β∈1,X,Y,Z(α,β) 6=(1,1)

λj,α,β α⊗ β, (B1)

wherein λj,α,β ∈ R. For each point x = (U1, U2, . . . , UR) ∈ SU(4)×R, we define the local exponential chart exp×Rx :su(4)×R → SU(4)×R as exp×Rx (H1, . . . ,HR) := (eiH1U1, . . . , e

iHRUR), and we define the analogous expU : su(2n) →SU(2n) as expU (H) := eiHU . These charts are standard for matrix Lie groups. Both are locally invertible in small neighbour-hoods around x and U by a standard result in Lie group theory [26]. The three-part map, represented by the dashed lines inFig. 5, has the form exp−1

FA(x)FA exp×Rx .

We now characterize the map’s derivative, to characterize the derivative of FA, to characterize the rank of FA. Denote by D0

the derivative evaluated where the Hermitian operators are set to zero, such that each chart reduces to the identity operation. Theimage of D0(exp−1

FA(x)FA exp×Rx ) is spanned by the operators

∂λj,A,B

(exp−1

FA(x)FA exp×Rx

) ∣∣∣0. (B2)

9

Subset of the domain,

SU(4)×R

Subset of the image,

SU(2n)

Subset of

realized as "#(4)×R ≃ ℝ15R ,(H1U1, …, HnUn)

Subset of

realized as "#(2n) ≃ ℝ4n,

HF(x)

Contraction map

Local chart, exp×R

x

Inverse of local chart

expF(x)

Figure 5. Three-part map used in the proof of Lemma 1. Hj denotes the jth two-qubit Hermitian operator, Uj denotes the jth two-qubitunitary, and H denotes an n-qubit Hermitian operator.

These operators have the form

UR . . . Uj+1PUj . . . U1, (B3)

wherein P denotes a two-qubit Pauli operator.We apply the setting above to prove the following lemma. To summarize the lemma: The low-rank locusE<rmax is of measure

0 or equals the whole domain, SU(4)×R.

Lemma 1 (Low-rank locus). The low-rank locus E<rmax is an algebraic set of measure 0 and so is closed (in the Lie-grouptopology). Equivalently, Ermax is an open set of measure 1. Consequently, dA = rmax.

Proof. Consider representing an operator (B3) as a matrix relative to an arbitrary tensor-product basis. To identify the matrix’sform, we imagine representing the unitaries in SU(4)×R as matrices relative to the corresponding tensor-product basis on C2 ⊗C2. Combining the unitary matrices’ elements polynomially yields the matrix elements of (B3).DxF

A has the same rank as D0(exp−1FA(x)

FA exp×Rx ), because exp×Rx and expFA(x) are local charts [31]. Recall thatE<rmax

denotes the locus of points, in SU(4)×R, where FA has a rank < rmax. Equivalently, by the invertible-matrix theorem,E<rmax

consists of the points where certain minors of D0(exp−1 FA exp×Rx )—the determinants of certain collections ofrmax×rmax matrix elements—vanish. The determinants’ vanishing implies a set of equations polynomial in the matrix elementsof D0(exp−1 FA exp×Rx )—and so, by the last paragraph, polynomial in the entries of matrices in SU(4)×R. SU(4)×R is areal algebraic set, being the set of operators that satisfy the polynomial equations equivalent to UU† = 1 and detU = 1. Thus,by Definition 4, the points of rank < r form an algebraic subset of SU(4)×R.

We can now invoke properties of algebraic subsets, reviewed in Appendix A. First, we prove that SU(4)×R is irreducible inthe Zariski topology. The Zariski topology of SU(4)×R is coarser than the topology inherited from (C4×4)×R, identified withR32R. As SU(4)×R is connected in the finer topology, so is SU(4)×R connected in the Zariski topology. This connectednessimplies that SU(4)×R is irreducible, as SU(4)×R is an algebraic group [34, Summary 1.36]. Being irreducible, SU(4)×R hasa dimension defined through Definition 7. If the low-rank locus E<rmax

is not all of SU(4)×R, then it is, by Definition 7, alower-dimensional algebraic subset. Every dimension-N algebraic subset decomposes into a collection of submanifolds, eachof which has dimension ≤ N [4, Prop. 9.1.8]. As a proper submanifold has measure 0, E<rmax

has measure 0. As an algebraicsubset, E<rmax

is closed in the Lie-group topology.Finally, we prove that dA = rmax. In a small open neighborhood V of a point x ∈ Ermax

, the contraction map’s rank isconstant, by Lemma 1. By the constant-rank theorem [31, Thm 5.13], therefore, FAT acts locally as a projector throughoutErmax

—and so throughout SU(4)×R (except on a region of measure 0, by Lemma 1). The projector has a rank, like FAT , ofrmax. A rank-rmax projector has an image that is a dimension-rmax manifold. Hence rmax ≤ dA. The other direction, dA ≤rmax, follows directly from Sard’s theorem [50]. Let Xr denote the set of points where FA is rank-r. As FA is a smooth map,Sard’s theorem ensures that r upper-bounds the Hausdorff dimension of the image FA(Xr). As FA(SU(4) is a semialgebraicset, it stratifies into manifolds, by Theorem 3. Therefore, the Hausdorff dimension coincides with the semialgebraic set’sdimension.

Lemma 1, combined with the following lemma, implies Proposition 1.

Lemma 3 (Existence of a high-rank point). Let T ∈ Z>0 denote any nonnegative integer. Consider any architecture AT formedfrom T causal slices of L gates each. The map FAT has the greatest rank possible, rmax ≥ T .

10

Figure 6. Examples of partial derivatives ∂λj,α,β (exp−1FA(x)

FA exp×4x )

∣∣λj,α,β=0

that span the image of D0(expFA(x) FA exp×4x ).

Proof. Without loss of generality, we assume that all T causal slices have identical architectures. This assumption will simplifythe notation below. We can lift the assumption by complicating the notation.

Consider an arbitrary point x = (U1, U2, . . . , UR) ∈ SU(4)×R. For all x, the contraction map FAT has a derivative charac-terized, in the proof of Lemma 1, with local charts expFAT (x) and exp×Rx . The number of gates in AT is R = TL. The mapFAT has an image spanned by the partial derivatives ∂λj,α,β (exp−1

FAT (x)FAT exp×Rx )

∣∣λj,α,β=0

. Each partial derivative hasthe form

URUR−1 . . . Uj+1(α⊗ β)UjUj−2 . . . U1 (B4)

(Fig. 6). α and β denote Pauli operators; each acts nontrivially on just one of the two qubits on which Uj acts nontrivially. Weimplicitly pad operators with identities wherever necessary such that the operators act on the appropriate Hilbert space.

We aim to lower-bound the greatest rank possible, rmax, of the map FAT . To do so, we construct a point

xT =(C

(1)1 , . . . , C

(L)1︸ ︷︷ ︸

L gates

, . . . , C(1)T , . . . , C

(L)T︸ ︷︷ ︸

L gates

)∈ SU(4)×R . (B5)

We will choose for theC(i)j ’s to be Clifford gates. A gate’s subscript, j, labels the slice to which the gate belongs. The superscript,

i, labels the gate’s position within the slice. The gates constitute a slice as C(L)j C

(L−1)j . . . C

(1)j =: Cj . Our construction of Cj

relies on a property of an arbitrary Pauli operator Qj : We can choose the Clifford gates C(i)j such that slice Cj maps Qj to a Z

on qubit t: CjQjC†j = Zt ≡ 1⊗(t−1) ⊗ Z ⊗ 1⊗(n−t). We now show how the existence of such a Clifford unitary Cj implies

Lemma 3. Afterward, we show to construct Cj .Let us choose the Pauli strings Qj that guide our construction of the Clifford slice Cj . We choose the Qj’s inductively over

T such that (CTCT−1 . . . Cj)Qj(Cj−1Cj−1 . . . C1)1≤j≤T is linearly independent. We start with an arbitrary Pauli string Q1.The form of Q1 guides our construction of C1. Second, we choose for Q2 to be an arbitrary Pauli string 6= C1Q1C

†1 . Q2 guides

our construction of C2. Third, we choose for Q3 to be an arbitrary Pauli string outside spanC1C2Q1C†2C†1 , C2Q2C

†2. This

Q3 guides our construction of C3. After T steps, we have constructed all the Qj’s and Cj’s. If T ≤ 4n− 1, enough Pauli stringsexist that, at each step, a Pauli string lies outside the relevant span.

The operators (CTCT−1 . . . Cj)Qj(Cj−1Cj−2 . . . C1), for j ∈ [1, T ], are in the image of D0(exp−1FAT (xT )

FAT exp×(R)xT ):

∂λjL,11,Z

(exp−1

FAT (xT )FAT exp×RxT

) ∣∣∣0

= (CTCT−1 . . . Cj+1)(1t−1 ⊗ Zt ⊗ 1n−t)(CjCj−1 . . . C1)

= (CTCT−1 . . . Cj)Qj(Cj−1Cj−2 . . . C1).(B6)

We have assumed, without loss of generality, that each slice’s final gate acts on qubit t. For all j ∈ [1, T ], the operators(CTCT−1 . . . Cj)Qj(Cj−1Cj−2 . . . C1) are in the image of D0(exp−1

FAT (xT )FAT exp×RxT ) and are linearly independent.

Therefore, the rank of FAT at the point xT is ≥ T .In the remainder of this proof, we provide the missing link: We show that, for every Pauli string P , we can construct a causal

slice that implements a Clifford unitaryC = C(L)C(L−1) . . . C(1) such thatCPC† = Zt. We drop subscripts because subscriptsindex slices and this prescription underlies all slices. By definition, each causal slice contains a qubit t to which each other qubitt′ connects via gates in the slice. The path from a given qubit t′ depends on t′, and multiple paths may connect a t′ to t. Also,one path may connect t to multiple qubits. We choose an arbitrary complete set of paths (which connect all the other qubits tot) that satisfies the merging property described below. To introduce the merging property, we denote by m the number of pathsin the set. Let p ∈ [1,m] index the paths. Path p contacts the qubits in the order ip,1 7→ ip,2 7→ . . . 7→ ip,lp = t, reachinglp ∈ [1, L + 1] qubits. We choose the paths such that they merge whenever they cross: If ip,j = ip′,j′ , then ip,j+k = ip′,j′+kfor all k ∈ 1, 2, . . . , lp − j=lp′ − j′. We choose for all the gates outside these paths to be identities. Next, we choose thenontrivial gates in terms of an arbitrary Pauli string.

11

Figure 7. Along every contraction is a redundant copy of the gauge group SU(2).

Let P =⊗n

j=1 Pj denote an arbitrary n-qubit Pauli string. Some Clifford unitary C maps P to a Pauli string that actsnontrivially on just one qubit (see Refs. [16, 59, 60] and [43]), which we choose to be t. We arbitrarily choose for the string’snontrivial single-qubit Pauli operator to be Z. Let ip,kp denote the first index j in ip,1 7→ ip,2 7→ . . . 7→ ip,lp for whichPj 6= 1. By definition, Pip,k ⊗ Pip,k+1

is a nontrivial Pauli string. There exists a two-local Clifford gate Cp,(0) that transformsPip,k ⊗ Pip,k+1

into a Z acting on qubit ip,k+1:

Cp,(0)(Pp,ik ⊗ Pp,ik+1)(Cp,(0))† = 1ik ⊗ Zik+1

. (B7)

Operating with Cp,(0) (padded with 1’s) on the whole string P yields another Pauli string:

Cp,(0)P (Cp,(0))† =

n⊗j=1

Pp,(1)j . (B8)

Let ip,` denote the first index j for which P p,(1)j is a nontrivial Pauli operator. Since

P(1)ip,k+1

⊗ P (1)ip,k+2

= Zip,k+1⊗ Pip,k+2

, (B9)

ip,` = ip,k+1. There exists a two-local Clifford gate Cp,(1) that shifts the Z down the path:

Cp,(1)(Pp,(1)ip,k+1

⊗ P p,(1)ip,k+2

)(Cp,(1))† = 11 ⊗ Zip,k+2

. (B10)

We perform this process—of shifting the Z down the path and leaving a 11 behind—for every path simultaneously. For example,if we begin with two equal-length paths, C(2) = Cp,(2)Cp

′,(2). This simultaneity is achievable until two paths merge. Wheneverpaths merge, we choose the next Clifford gate such that we proceed along the merged path. Every qubit is visited, and everypath ends at qubit t. Therefore, we have constructed a circuit that implements a Clifford operation C such that CPC† = Zt.Figure 4(d) depicts an example of this construction.

The foregoing proof has a surprising implication: A map’s rank is somewhat divorced from a circuit’s complexity. The rankof FAT at xT is at least T , which could be a large number. Yet, the contracted unitary corresponding to this circuit is Clifford.Hence the extended circuit’s complexity surpassed the original circuit’s complexity only a little—by, at most,O(n2/ log(n)) [2].

Finally, we combine Lemmata 1 and 3 to prove Theorem 1:

Theorem 1 (Linear growth of complexity). Let U denote a unitary implemented by a random quantum circuit in an architectureformed from T = R/L causal slices of L gates each. The unitary’s circuit complexity is lower-bounded as

Cu(U) ≥ R

9L− n

3, (1)

with unit probability, until the number of gates grows to T = R/L ≥ 4n − 1. The same bound holds for Cstate(U |0n〉), untilT = R/L ≥ 2n+1 − 1.

Proof of Theorem 1. We reuse the notation introduced in Lemmata 1 and 3. Examples include AT , an arbitrary architecture thatsatisfies the assumptions in Lemma 3 and that consists of R=TL gates. FAT denotes the corresponding contraction map. Ermax

denotes the locus of points at which FAT achieves its greatest rank, rmax. In a small open neighborhood V of a point x ∈ Ermax,

the contraction map’s rank is constant, by Lemma 1. By the constant-rank theorem [31, Thm 5.13], therefore, FAT acts locallyas a projector throughout Ermax

—and so throughout SU(4)×R (except on a region of measure 0, by Lemma 1). The projectorhas a rank, like FAT , of rmax. Therefore, in the open set V ⊆ SU(4)×R, FAT is equivalent, up to a diffeomorphism, to theprojection

(x1, . . . , xdim(SU(4)×R)) 7→ (x1, . . . , xrmax , 0, . . . , 0︸ ︷︷ ︸dim(SU(2n))−rmax

). (B11)

12

For simplicity of notation, we identify V with its image under the local diffeomorphism (we do not distinguish V from its imagenotationally).

The open subset V contains, itself, an open subset that decomposes as a product: V1 × V2 ⊆ V , such that x ∈ V1 × V2 and,as suggested by Eq. (B11),

V1 ⊆ Rrmax , and V2 ⊆ Rdim(SU(4)×R)−rmax . (B12)

(Again to simplify notation, we are equating the local sets Vj=1,2 with their images, under local charts, in Rm, for m ∈ Z>0.)From now on, V1 × V2 is the open subset of interest. The contraction map’s equivalence to a projector, in V1 × V2, will help uscompare high-depth circuits with low-depth circuits: Consider a circuit whose contraction map takes some local neighborhoodto an image of some dimension. How does the dimension differ between high-depth circuits and low-depth circuits? We start byupper-bounding the dimension for low-depth circuits.

We have been discussing an R-gate architecture AT . Consider any smaller architecture A′ of R′ < R gates. A′ is encodedin a contraction map FA

′whose domain is SU(4)×R

′. As explained in the proof of Lemma 1, FA

′is a polynomial map.

Therefore, FA′

has a property prescribed by the Tarski-Seidenberg principle [4] (Theorem 2): The image FA′(SU(4)×R

′) is a

semialgebraic set of dimension ≤ dim(SU(4)×R′) = R′ dim(SU(4)) = 15R′.

We can strengthen this bound: Consider contracting two gates that share a qubit. The shared qubit may undergo a one-qubitgate specified by three parameters (one parameter per one-qubit Pauli). The one-qubit gate can serve as part of the first two-qubit gate or as part of the second two-qubit gate; which does not affect the contraction. Hence the contraction contains 3fewer parameters than expected [44]. Let us classify the shared qubit as an input of the second two-qubit gate. A two-qubitgate in a circuit’s bulk accepts two input qubits outputted by earlier gates. So we might expect an R′-gate circuit to havedim(FA

′(SU(4)×R

′)) ≤ 15R′ − 2× 3R′ = 9R′. But the first n/2 gates [the leftmost vertical line of gates in Fig. 2(a)] receive

their input qubits from no earlier gates. So we must restore 3 × 2 parameters for each of the n/2 initial gates, or restore 3nparameters total [45]

dim(FA′(SU(4)×R

′)) ≤ 9R′ + 3n. (B13)

We have upper-bounded the dimension for low-depth circuits.We now lower-bound the corresponding dimension for high-depth circuits. We can do so by lower-bounding the greatest

possible rank, rmax, of a high-depth architecture’s contraction map FAT : In an open neighborhood of x ∈ SU(4)×R, FAT isequivalent to a projector, which has some rank. The neighborhood’s image, under the projector, is a manifold. The manifold’sdimension equals the projector’s rank. Therefore, we bound the rank to bound the dimension.

Augmenting an architecture with T L-gate causal slices increases the contraction map’s greatest possible rank, rmax, by≥ T − 1. Therefore, for a circuit of R = TL gates in the architecture AT , we have constructed a point of rank T = R/L.Therefore,

rmax ≥ R/L. (B14)

We have lower-bounded the dimension of the image of a high-depth architecture’s contraction map [the rank in Ineq. (B14)]and have upper-bounded the analogous dimension for a low-depth architecture [Ineq. (B13)]. The high-depth-architecture di-mension upper-bounds the low-depth-architecture dimension,

dim(FA

′(SU(4)×R

′))< rmax, (B15)

if

9R′ + 3n < rmax, (B16)

by Ineq. (B13). Furthermore, by Ineqs. (B13) and (B14), Ineq. (B15) holds if 9R′ + 3n < R/L, or R′ < R9L −

n3 .

R′ <R

9L− n

3. (B17)

holds. We have upper-bounded the short circuit’s gate count in terms of the deep circuit’s gate count.Let us show that, if Ineq. (B17) holds, the short circuits form a set of measure 0 in SU(4)×R [46]. We will begin with a point

x ∈ Ermax; apply the short-architecture contraction map FA

′; and follow with the deep-architecture contraction map’s inverse,

(FAT )−1. The result takes up little space in SU(4)×R, we will see.To make this argument rigorous, we recall the small open neighborhood V1 × V2 of x ∈ Ermax

. In V1 × V2, FA′(SU(4)×R

′)

has the preimage, under FAT , of(FAT |V1×V2

)−1 (FA

′(SU(4)×R

′))'[FA

′(SU(4)×R

′) ∩ V1

]× V2. (B18)

13

The ' represents our identification of the map FA with its representation in local charts. By the proof of Lemma 1,FA

′(SU(4)×R

′) is a semialgebraic set. Therefore, by Theorem 3, FA

′(SU(4)×R

′) is a union of smooth manifolds. Each

manifold is of dimension≤ 9R′+3n, by Theorem 3 and Ineq. (B13). By Eq. (B12), V2 is of dimension dim(SU(4)×R)−rmax.Therefore, [FA

′(SU(4)×R

′) ∩ V1] × V2 consists of manifolds of dimension ≤ 9R′ + 3n + dim(SU(4)×R) − rmax. Using

Ineq. (B16), we can cancel the 9R′ + 3n with the −rmax, at the cost of loosening the bound: [FA′(SU(4)×R

′) ∩ V1] × V2

consists of manifolds of dimension < dim(SU(4)×R). As a collection of manifolds of submaximal dimension, the unitariesimplemented by short circuits satisfying (B17), restricted to a small open neighborhood V1, form a set of measure 0 [47].

Let us extend this conclusion about n-qubit unitaries—about images of maps FA′—to a conclusion about preimages—about

lists of gates. By Lemma 1, Ermaxis of measure 1. Therefore, for every ε > 0, there exists a compact subset K ⊆ Ermax

ofmeasure 1−ε. SinceK is compact, for any cover ofK by open subsets, a finite subcover exists. The foregoing paragraph showsthat, restricted to each open set in this finite subcover, the preimage of the unitaries reached by lower-depth circuits is of measure0. Therefore, the preimage of the R′-gate, architecture-A′ circuits is of measure ≤ ε. Since ε > 0 is arbitrary, the preimage is ofmeasure 0. The foregoing argument holds for each architecture A′ of R′ gates. Hence each preimage forms a set of measure 0.The total measure is subadditive. So the union of the preimages, over all architectures with ≤ R′ gates, is of measure 0.

We have proven the circuit-complexity claim posited in Theorem 1. The state-complexity claim follows from tweaks to theproof. The tweaks are described in Appendix C.

Appendix C: Proof of the linear growth of state complexity

At the end of Appendix B, we proven part of Theorem 1—that circuit complexity grows linearly with the number of gates.Here, we prove rest of the theorem—that state complexity grows linearly. We need only tweak the proof presented in Appendix B.

Consider instead of the contraction map FAT , the map that contracts a list of gates, forming an architecture-AT circuit, andapplies the circuit to |0n〉, to get

GAT : SU(4)×R → S2×2n−1 ⊆ C2n . (C1)

The argument works the same as in Appendix B, with one exception: The derivative DxGAT has an image that does

not contain 4n − 1 nontrivial linearly independent Pauli operators. Rather, the image contains the computational basisiκ|x〉x∈0,1n,κ∈0,1 formed by applying tensor products of Z,X and Y to |0n〉. (We denote the imaginary number

√−1 by

i.) The proof of Lemma 1 ports over without modification, as GAT is a polynomial map between algebraic sets.The proof of Lemma 3 changes slightly. We must prove the existence of a point x ∈ SU(4)×R at which GAT has a rank at

least linear in the circuit depth. The only difference in the proof is, we must choose the operators Qj inductively such that thestates (CTCT−1 . . . Cj)Qj(Cj−1Cj−2 . . . C1)|0n〉 are linearly independent. Such a choice is possible if T < 2 × 2n − 1, thenumber of real parameters in a pure n-qubit state vector.

Appendix D: Randomized architectures

From Theorem 1 follows a bound on the complexity of a doubly random circuit: Not only the gates, but also the gates’positions, are drawn randomly. This model features in Ref. [8]. Our proof focuses on nearest-neighbor gates, but other models(such as all-to-all interactions) yield similar results.

Corollary 1 (Randomized architectures). Consider drawing an n-qubit unitary U according to the following probability distri-bution: Choose a qubit j uniformly randomly. Apply a Haar-random two-qubit gate to qubits j and j + 1. Perform this processR times. With high probability, the unitary implemented has a high complexity: For all α ∈ [0, 1),

Pr

(Cu(U) ≥ α R

9n(n− 1)2− n

3

)≥ 1− 1

1− α(n− 1)e−n . (5)

Proof. The proof relies on the following strategy: We consider constructing slices randomly to form a circuit. If the slicescontain enough gates, we show, many of the slices are causal. This result enables us to apply Theorem 1 to bound the circuit’scomplexity.

Consider drawing L gates’ positions uniformly randomly. For each gate, the probability of drawing position (j, j + 1)is 1/(n − 1). The probability that no gates act at position (j, j + 1) is (1 − 1/(n − 1))L. Let us choose for each sliceto contain L = n(n − 1)2 gates. Define a binary random variable Ij as follows: If one of the gates drawn during steps(j − 1)n(n− 1), (j − 1)n(n− 1) + 1, . . . , jn(n− 1) acts at (j, j + 1), then Ij = 1. Otherwise, Ij = 0. With high probability,

14

gates act at all positions:

p := Pr

n−1∧j=1

(Ij = 1)

=

(1−

(1− 1

n− 1

)n(n−1))n−1≥(1− e−n

)n−1 ≥ 1− (n− 1)e−n. (D1)

We have invoked the inverse Bernoulli inequality and the Bernoulli inequality. We will use this inequality to characterize causalslices.

Consider drawing T L-gate slices randomly, as described in the corollary. Denote by X the number of slices in which at leastone position is bereft of gates: For some j, Ij = 0. With high probability, X is small: For all a ∈ (0, T ],

Pr(X ≥ a) ≤ T (1− p)a

≤ T (n− 1) e−L/n/a , (D2)

by Markov’s inequality. Let us choose for the threshold to be a = (1− α)T . With overwhelming probability, αT slices satisfy∧j(Ij = 1) and so contain gates that act at all positions (j, j + 1) in increasing order. Therefore, these slices contain a staircase

architecture and are so causal. Therefore, a slight variation on Theorem 1 governs the αT × L = αR gates that form the causalslices. Strictly speaking, Theorem 1 governs only consecutive causal slices. In contrast, extra gates may separate the causal sliceshere. However, the extra gates can only increase the contraction map’s image. Therefore, the additional (1−α)TL gates cannotdecrease the accessible dimension dAT . Therefore, the bound from Theorem 1 holds. With probability ≥ 1− 1

1−α (n− 1) e−n

over the choice of architecture,

Cu(U) ≥ R− (1− α)R

9n2(n− 1)− n

3, (D3)

with probability one over the choice of gates. This bound is equivalent to Ineq. (5).

Appendix E: Proof of Corollary 2

Corollary 2 extends Theorem 1 to accommodate errors in the target unitary’s implementation. We prove Corollary 2 bydrawing on the proof of Theorem 1 and reusing notation therein.

Corollary 2 (Slightly robust circuit complexity). Let U denote the n-qubit unitary implemented by any random quantum circuitin any architectureAT that satisfies the assumptions in Theorem 1. Let U ′ denote the n-qubit unitary implemented by any circuitofR′ ≤ R/(9L)−n/3 gates. For every δ ∈ (0, 1], there exists an ε := ε(AT , δ) > 0 such that ||U −U ′||F ≥ ε, with probability1− δ, unless R/L > 4n − 1.

Proof of Corollary 2. The proof of Theorem 1 can be modified to show that, for every δ > 0, there exists an open set B ⊆SU(2n) that contains FA

′(SU(4)×R

′), such that the preimage (FAT )−1(B) is small—of measure ≤ δ. The modification is as

follows. For every δ′ > 0, there exists a measure-(1 − δ′) compact subset K of Ermax. As K is compact, there exists a finite

cover of K that has the following properties: K is in the union ∪jV j of subsets V j . On the V j , the contraction map FAT isequivalent to a projector, up to a local diffeomorphism. As in the proof of Theorem 1, we can assume, without loss of generality,that V j = V j1 × V

j2 . The V j1 and V j2 are defined analogously to the V1 and V2 in the proof of Theorem 1. For each V j , there

exists an open neighborhood W j of FA′(SU(4)×R

′) ∩ V j1 such that W j has an arbitrarily small measure δ′′j > 0. Therefore,

B := ∪jW j has a preimage of measure ≤ δ′ +∑j δ′′j = δ. Each of the summands, though positive, can be arbitrarily small.

The Frobenius norm induces a metric dF on SU(4)×R. In terms of dF, we define the function

dF(. , FAT (SU(4)×R) \B

): FA

′(SU(4)×R

′)→ R≥0. (E1)

This function is continuous, and FA′(SU(4)×R

′) is compact. Therefore, the function achieves its infimum at a point xmin ∈

FA′(SU(4)×R

′). Therefore, the minimal distance to FAT (SU(4)×R) \ B is dF(xmin, F (SU(4)×R) \ B). Since B is open,

FA(SU(4)×R) \B is closed and so compact. By the same argument,

ε(AT , δ) := dF(xmin, F (SU(4)×R) \B

)= infy∈F (SU(4)×R)\B

dF(xmin, y) = dF(xmin, ymin) > 0. (E2)

We have identified an ε > 0 that satisfies Corollary 2.

15

Appendix F: Notions of circuit complexity

As circuit complexity is a widely popular concept, there is a zoo of quantities that measure it. We prove our main theorem forthe straightforward definition of exact circuit implementation—the clearest and historically first notion of a circuit complexity—and for a version of approximate circuit complexity (Corollary 2) with an uncontrollably small error. In this appendix, we brieflymention other notions of complexity, partially to review other notions and partially to place the main text’s findings in a widercontext. Let U ∈ SU(2n) denote a unitary. Ref. [41] discusses notions of approximate circuit complexity.

Definition 8 (Approximate circuit complexity). The approximate circuit complexity Cu(U, η) is the least number of 2-localgates, arranged in any architecture, that implements U up to an error η > 0 in operator norm ||.||.

This definition is similar in mindset to the above (slightly) robust definition of a circuit complexity. For any pair U,U ′ ∈ SU(2n)of circuits,

1

2n||U − U ′||F ≤ ||U − U ′|| ≤ ||U − U ′||F. (F1)

A widely used proxy for quantum circuit complexity—one that is increasingly seen as a complexity measure in its own right—isNielsen’s geometric approach to circuit and state complexity [18, 38, 41]. This approach applies geometric reasoning to circuitcomplexity and led to many intuitive insights, including Brown and Susskind’s conjectures about the circuit complexity’s behav-ior under random evolution. To connect to cost functions as considered in Nielsen’s framework, consider 2-local Hamiltonianterms H1, H2, . . . ,Hm in the Lie algebra su(2n) of traceless Hermitian matrices, normalized as ‖Hj‖ = 1 for j = 1, 2, . . . ,m.Consider generating a given unitary, by means of a control system, following Schrödinger’s equation:

d

dtU(t) = −iH(t)U(t), wherein H(t) =

m∑j=1

hj(t)Hj . (F2)

The control function [0, τ ] → Rm is defined as t 7→ (h1(t), . . . , hm(t)) and satisfies U(0) = 1. That is, a quantum circuitresults from time-dependent control. In practice, not all of Rm reflects meaningful control parameters; merely a control regionR ⊂ Rm does. With each parameterized curve is associated a cost function c : R → R, so that the entire cost of a unitaryU ∈SU(2n) becomes

C(U) := infT, t 7→H(t)

∫ τ

0

dt c(H(t)). (F3)

We take the infimum over all time intervals [0, τ ] and over all control functions t 7→ H(t) such that the control parameters are inR for all t ∈ [0, τ ] and scuh that U(τ) = U . Several cost functions are meaningful and have been discussed in the literature. Acommon choice is

cp(H(t)) =

m∑j=1

hj(t)p

1/p

. (F4)

In particular, c2 gives rise to a sub-Riemannian metric. For the resulting cost C2(U), Ref. [40] establishes a connection betweenthe approximate circuit complexity and the cost: Any bound on the approximate circuit complexity, with an approximation errorbounded from below independently of the system size, immediately implies a lower bound on the cost.

Theorem 4 (Approximate circuit complexity and cost [40]). For every integer n, every U ∈ SU(2n) and every η > 0,

Cu(U, η) ≤ c C2(U)3n6

η2. (F5)

The quantity on the right-hand side can, in turn, be upper-bounded: C2(U) ≤ C1(U). This C1 has a simple interpretation interms of a weighted gate complexity [19].

Definition 9 (Weighted circuit complexities). Let U ∈ SU(2n) denote a unitary. The weighted circuit complexity Cw(U) equalsthe sum of the weights of 2-local gates, arranged in any architecture, that implement U , wherein each gate Uj is weighted by itsstrength W (Uj), defined through

W (U) := inf‖h‖ : U = eih

. (F6)

16

The weighted circuit complexity Cw(U) turns out to equal the cost C1(U) for a given unitary. We can grasp this result byTrotter-approximating the time-dependent parameterized curve in the definition of C1(U).

Lemma 4 (Weighted circuit complexity and cost). If n denotes an integer and U ∈ SU(2n), then

Cw(U) = C1(U). (F7)

Therefore, the weighted circuit complexity grows like the cost C1. By implication, the circuit complexity’s growth willbe reflected by a notion of circuit complexity that weighs the quantum gates according to their strengths. Again, once themain text’s approximate circuit complexity is established with an n-independent approximation error, one finds bounds on theweighted circuit complexity, as well.

The last important notion of circuit complexity that has arisen in the recent literature is that of Ref. [9]. Denote byGa ⊂SU(22n) the set of 2n-qubit unitary circuits comprised of ≤ a elementary quantum gates, wherein the first n qubitsform the actual system and the next n qubits form a memory. Let Mb denote the class of all two-outcome measurements,defined on 2n qubits, that require quantum circuits whose implementation requires ≤ b elementary quantum gates. Define

β(r, U) := maximize∣∣tr (M

[U ⊗ 1]|φ〉〈φ|[U ⊗ 1]† − [1/2n ⊗ tr1(|φ〉〈φ|)])∣∣ , (F8)

:= subject to M ∈Mb, |φ〉 = V |02n〉, V ∈ Ga, r = a+ b. (F9)

In terms of this quantity, Ref. [9] defined strong unitary complexity.

Definition 10 (Strong unitary complexity [9]). Let r ∈ R and δ ∈ (0, 1). A unitary U ∈ SU(2n) has strong unitary complexity≤ r if

β(r, U) ≥ 1− 1

22n− δ, (F10)

denoted by C(U, δ) ≥ r.

While seemingly technically involved, the definition is operational. The definition is also more stringent and demanding thanmore-traditional definitions of approximate circuit complexity. To concretize this statement, we denote the diamond norm by‖.‖ [58].

Lemma 5 (Implications of strong unitary complexity [9]). Suppose that U ∈ U(2n) obeys C(U, δ) ≥ r+ 1 for some δ ∈ (0, 1),r ∈ R, arbitrary measurement procedures that include the Bell measurement. Then

minCu(V )≤r

1

2‖U − V‖ >

√δ . (F11)

That is, it is impossible to accurately approximate U with circuits V of < r elementary quantum gates.

U and V denote the unitary quantum channels defined by U(ρ) = UρU† and V(ρ) = V ρV †. The diamond norm between themis

1

2‖U − V‖ =

1

2supρ‖(U ⊗ 1)ρ(U ⊗ 1)† − (V ⊗ 1)ρ(V ⊗ 1)†‖1 (F12)

≤ 1

2supρ‖[(U − V )⊗ 1]ρ(U ⊗ 1)†‖1 +

1

2supρ‖(U ⊗ 1)ρ[(U − V )⊗ 1]†‖1.

We have added and subtracted a term and have used the triangle inequality. Therefore,

1

2‖U − V‖ ≤

1

2‖U − V ‖∞

(supρ‖ρ(U ⊗ 1)†‖1 + sup

ρ‖(V ⊗ 1)ρ‖1

)≤ ‖U − V ‖∞, (F13)

as the operator norm is a weakly unitarily invariant norm. Therefore, C(U, δ) ≥ r+1 implies that C(U, δ) ≥ r. That is, the strongunitary complexity of Ref. [9] is tighter than approximate circuit complexity. A topic of future work will be the exploration ofthe growth of approximate notions of complexity with an approximation error independent of the system size.

[1] The run-time of the best known algorithms for the T -count [21]—deciding whether the optimal gate decomposition of a circuit presentedas a sequence of Clifford gates and T gates on n qubits involves ≤ m T gates—is O(Nmpoly(m,N)), with N := 2n.

17

[2] S. Aaronson and D. Gottesman. Improved simulation of stabilizer circuits. Phys. Rev. A, 70:052328, 2004.[3] F. Arute et al. Quantum supremacy using a programmable superconducting processor. Nature, 574:505–510, 2019.[4] J. Bochnak, M. Coste, and M.-F. Roy. Real algebraic geometry, volume 36. Springer Science & Business Media, 2013.[5] B. Bolt, T. G. Room, and G. E. Wall. On the Clifford collineation, transform and similarity groups. i. J. Austr. Math. Soc., 2:60–79, 1961.[6] B. Bolt, T. G. Room, and G. E. Wall. On the Clifford collineation, transform and similarity groups. II. J. Austr. Math. Soc., 2:80–96,

1961.[7] A. Bouland, B. Fefferman, and U. Vazirani. Computational pseudorandomness, the wormhole growth paradox, and constraints on the

AdS/CFT duality. arXiv:1910.14646, 2019.[8] F. G. S. L. Brandão, A. W. Harrow, and M. Horodecki. Local random quantum circuits are approximate polynomial-designs. Commun.

Math. Phys., 346:397–434, 2016.[9] F. G. S. L. Brandão, W. Chemissany, N. Hunter-Jones, R. Kueng, and J. Preskill. Models of quantum complexity growth.

arXiv:1912.04297, 2019.[10] A. R. Brown, D. A. Roberts, L. Susskind, B. Swingle, and Y. Zhao. Complexity, action, and black holes. Phys. Rev. D, 93(8):086006,

2016.[11] A. R. Brown, D. A. Roberts, L. Susskind, B. Swingle, and Y. Zhao. Holographic complexity equals bulk action? Phys. Rev. Lett.,

116:191301, 2016.[12] A. R. Brown and L. Susskind. Second law of quantum complexity. Phys. Rev. D, 97:086015, 2018.[13] A. R. Calderbank, E. M. Rains, P. M. Shor, and N. J. A. Sloane. Quantum error correction via codes over gf (4). IEEE Trans. Inf. Th.,

44:1369–1387, 1998.[14] A. R. Calderbank, E. M. Rains, P. W. Shor, and N. J. A. Sloane. Quantum error correction and orthogonal geometry. Phys. Rev. Lett.,

78:405, 1997.[15] S. Chapman, J. Eisert, L. Hackl, M. P. Heller, R. Jefferson, H. Marrochio, and R. C. Myers. Complexity and entanglement for thermofield

double states. SciPost Phys., 6:034, 2019.[16] R. Cleve, D. Leung, L. Liu, and C. Wang. Near-linear constructions of exact unitary 2-designs. Quant. Inf. Comp., 16:0721–0756, 2016.[17] C. Dankert, R. Cleve, J. Emerson, and E. Livine. Exact and approximate unitary 2-designs and their application to fidelity estimation.

Phys. Rev. A, 80:012304, 2009.[18] M. R. Dowling and M. A. Nielsen. The geometry of quantum computation. Quant. Inf. Comp., 8:861–899, 2008.[19] J. Eisert. Entangling power and quantum circuit complexity. arXiv:2104.03332, 2021.[20] N. Yunger Halpern et al., 2021. In preparation.[21] D. Gosset, V. Kliuchnikov, M. Mosca, and V. Russo. An algorithm for the t-count. Quant. Inf. Comp., 14:1277–1301, 2014.[22] D. Gottesman. Stabilizer codes and quantum error correction. quant-ph/9705052, 1997.[23] D. Gross, K. M. R. Audenaert, and J. Eisert. Evenly distributed unitaries: on the structure of unitary designs. J. Math. Phys., 48:052104,

2007.[24] J. Haferkamp and N. Hunter-Jones. Improved spectral gaps for random quantum circuits: large local dimensions and all-to-all interactions.

arXiv:2012.05259, 2020.[25] J. Haferkamp, F. Montealegre-Mora, M. Heinrich, J. Eisert, D. Gross, and I. Roth. Quantum homeopathy works: Efficient unitary designs

with a system-size independent number of non-Clifford gates. arXiv:2002.09524, 2020.[26] B. Hall. Lie groups, Lie algebras, and representations: an elementary introduction, volume 222. Springer, 2015.[27] P. Hayden and J. Preskill. Black holes as mirrors: quantum information in random subsystems. JHEP, 0709:120, 2007.[28] N. Hunter-Jones. Unitary designs from statistical mechanics in random quantum circuits. arXiv:1905.12053, 2019.[29] N. Khaneja and S. Glaser. Cartan decomposition of su (2n), constructive controllability of spin systems and universal quantum computing.

quant-ph/0010100, 2000.[30] E. Knill. Approximation by quantum circuits. quant-ph/9508006, 1995.[31] J. M. Lee. Smooth manifolds. In Introduction to smooth manifolds, pages 1–31. Springer, 2013.[32] J. Maldacena, S. H. Shenker, and D. Stanford. A bound on chaos. JHEP, 1608:106, 2016.[33] J. M. Maldacena. Eternal black holes in anti-de sitter. JHEP, 04:021, 2003.[34] J. S. Milne. Algebraic groups: the theory of group schemes of finite type over a field, volume 170. Cambridge University Press, 2017.[35] A. Nahum, S. Vijay, and J. Haah. Operator spreading in random unitary circuits. Phys. Rev. X, 8:021014, 2018.[36] Y. Nakata, C. Hirche, M. Koashi, and A. Winter. Efficient quantum pseudorandomness with nearly time-independent hamiltonian dy-

namics. Phys. Rev. X, 7:021006, 2017.[37] C. Neill, P. Roushan, K. Kechedzhi, S. Boixo, S. V. Isakov, V. Smelyanskiy, R. Barends, B. Burkett, Y. Chen, and Z. Chen. A blueprint

for demonstrating quantum supremacy with superconducting qubits. 2017.[38] M. A. Nielsen. A geometric approach to quantum circuit lower bounds. quant-ph/0502070, 2005.[39] M. A. Nielsen and I. L. Chuang. Quantum computation and quantum information. Cambridge Series on Information and the Natural

Sciences. Cambridge University Press, 2000.[40] M. A. Nielsen, M. R. Dowling, M. Gu, and A. C. Doherty. Optimal control, geometry, and quantum computing. Phys. Rev. A, 73:062323,

2006.[41] M. A. Nielsen, M. R. Dowling, M. Gu, and A. C. Doherty. Quantum computation as geometry. Science, 311:1133–1135, 2006.[42] A t-design is a probability distribution, over unitaries, whose first t moments equal the Haar measure’s moments [17, 23]. The Haar

measure is the unique unitarily invariant probability measure over a compact group.[43] Consider conjugating an arbitrary n-qubit Pauli operator P with a uniformly random Clifford operator C. The result, C†PC, is a

uniformly random n-qubit Pauli operator [16, 59, 60]. Therefore, for every initial Pauli operator P and every final Pauli operator, someClifford operator C maps one to the other.

[44] In other words, the contraction has a redundant copy of the gauge group SU(2): Every unitary U ∈ SU(4) decomposes as (U1 ⊗

18

U2)K(V1 ⊗ V2), wherein K = ei(aZ⊗Z+bY⊗Y+cX⊗X) and the Uj and the Vj denote single-qubit unitaries [29].[45] Technically, this bound on the dimension does not follow from Lemma 2 as it does for the bound dimFA

′(SU(4)×R

′) ≤ 15R′. This

is because the quotient space SU(4)×R′/SU(2)×(2R′−n) is not necessarily semialgebraic. This can however be resolved by envoking

Sard’s theorem [50], which asserts as a special case that the Hausdorff dimension of the image of a smooth map is bounded by thedimension of its domain. As a semialgebraic set’s dimension is just the highest dimension in it’s stratification it agrees with the Hausdorffdimension.

[46] Technically, the sets of few gates form a set of measure 0 in SU(4)×R. Circuits are not in SU(4)×R, as explained in Fig. 3. However, weexpected the sentence above to be more intuitive with “short circuits” instead of “sets of few gates.”.

[47] Again, by “short circuits,” we mean, “sets of few gates.” We replaced the latter phrase for ease of expression.[48] E. Onorati, O. Buerschaper, M. Kliesch, W. Brown, A. H. Werner, and J. Eisert. Mixing properties of stochastic quantum hamiltonians.

Commun. Math. Phys., 355:905, 2017.[49] D. A. Roberts and B. Yoshida. Chaos and complexity by design. JHEP, 2017:121, 2017.[50] A. Sard. Hausdorff measure of critical images on Banach manifolds. Am. J.f Math., 87:158–174, 1965.[51] Y. Sekino and L. Susskind. Fast scramblers. JHEP, 10:65, 2008.[52] D. Stanford and L. Susskind. Complexity and shock wave geometries. Phys. Rev. D, 90:126007, 2014.[53] Douglas Stanford and Leonard Susskind. Complexity and shock wave geometries. Phys. Rev. D, 90:126007, 2014.[54] L. Susskind. Entanglement is not enough. arXiv:1411.0690.[55] L. Susskind. Computational complexity and black hole horizons. Fort. Phys., 64:24–43, 2016.[56] L. Susskind. Black holes and complexity classes. arXiv:1802.02175, 2018.[57] Leonard Susskind and Ying Zhao. Switchbacks and the bridge to nowhere. August 2014.[58] J. Watrous. Semidefinite programs for completely bounded norms. Th. Comp., 5, 2009.[59] Z. Webb. The Clifford group forms a unitary 3-design. 2015. arXiv:1510.02769.[60] H. Zhu. Multi-qubit Clifford groups are unitary 3-designs. Phys. Rev. A, 96:062336, 2017.