Emergent quantum state designs from individual many-body wavefunctions

Jordan S. Cotler,1, ∗ Daniel K. Mark,2, ∗ Hsin-Yuan Huang,3, ∗ Felipe Hern´andez,4 Joonhee Choi,3 Adam L. Shaw,3 Manuel Endres,3, † and Soonwon Choi2, 5, ‡ 1Society of Fellows, Harvard University, Cambridge, MA 02138, USA 2Center for Theoretical Physics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA 3California Institute of Technology, Pasadena, CA 91125, USA 4Department of Mathematics, Stanford University, Stanford, CA 94305, USA 5Department of Physics, University of California, Berkeley, CA 94720, USA Quantum chaos in many-body systems provides a bridge between statistical and quantum physics with strong predictive power. This framework is valuable for analyzing properties of complex quan- tum systems such as energy spectra and the dynamics of thermalization. While contemporary methods in quantum chaos often rely on random ensembles of quantum states and Hamiltonians, this is not reflective of most real-world systems. In this paper, we introduce a new perspective: across a wide range of examples, a single non-random quantum state is shown to encode universal and highly random quantum state ensembles. We characterize these ensembles using the notion of quantum state k-designs from quantum information theory and investigate their universality using a combination of analytic and numerical techniques. In particular, we establish that k-designs arise naturally from generic states as well as individual states associated with strongly interacting, time- independent Hamiltonian dynamics. Our results offer a new approach for studying quantum chaos and provide a practical method for sampling approximately uniformly random states; the latter has wide-ranging applications in quantum information science from tomography to benchmarking.

Analyzing the exact dynamics of general, strongly in- of pure states on a subsystem; this ensemble goes beyond teracting quantum many-body systems is intractable us- the reduced density matrix. Then we leverage statistical ing existing analytic and numerical tools. However, there tools to analyze the degree of randomness of the local en- is a widely used heuristic for understanding chaotic quan- semble, and in turn characterize the global wavefunction. tum dynamics: the eigenstates and eigenvalues of chaotic Concretely, an ensemble of states can be generated Hamiltonians have properties as if they were sampled from a single wavefunction by performing local measure- from a random ensemble. This heuristic leads one to ments over only part of the total system. We consider a leverage statistical approaches, such as random matrix many-body system partitioned into a subsystem A and theory [1,2], to address many physical problems and has its complement B. Performing local measurements on become a foundational principle in understanding chaos B, we obtain exponentially many different pure states and thermalization in quantum systems [3–6]. Exam- on A, each corresponding to a distinct measurement out- ples of this heuristic include Berry’s conjecture [7] and come on B. We call the set of pure states on A, along the eigenstate thermalization hypothesis (ETH) [8–10], with the associated measurement probabilities, the pro- which have been supported by an overwhelming amount jected ensemble on A (see Fig.1a,b); see also [14]. To of numerical evidence [11–13]. quantify the degree of randomness of an ensemble, we While these statistical approaches hinge on the pres- use a well-established notion from quantum information ence of random ensembles, realistic quantum systems theory, namely quantum state k-designs [15, 16]. An en- are often described by a fixed Hamiltonian. Accord- semble of pure states is said to form a quantum state k- ingly, energy eigenstates, eigenvalues, and states evolved design if the probability distribution over the constituent by Hamiltonian dynamics are deterministic functions of states is indistinguishable up to k-th moments from the a particular Hamiltonian without any randomness in- uniform distribution over the entire Hilbert space. There-

arXiv:2103.03536v1 [quant-ph] 5 Mar 2021 volved. Thus, it is of fundamental interest to establish fore, forming a higher k-design implies that the ensemble a connection between isolated quantum systems and the is more uniformly distributed over the Hilbert space. emergence of random ensembles that dictate their statis- Remarkably, using our new approach, we find that tical properties. approximate k-designs arise from a variety of natural In this paper, we present a new perspective on the quantum many-body states. We establish two theo- emergence of statistical behavior in chaotic quantum rems showing that the projected ensemble coming from many-body systems: instead of imagining that a phys- generic many-body quantum states forms an approxi- ical state is sampled from a random ensemble, we use a mate k-design as long as the size of B is sufficiently larger single many-body wavefunction to generate an ensemble than the size of A. Furthermore, a concurrent work [14] finds evidence that approximate k-designs emerge from projected ensembles in a Rydberg quantum simulator. ∗ These authors contributed equally to this work. We argue that this is a much more general phenomenon: † [email protected][email protected] we find strong numerical evidence across several mod- els that approximate k-designs arise from both quantum 2

FIG. 1. Emergence of a universal quantum state ensemble from a single many-body wavefunction. a, A subsystem B of a pure many-body wavefunction |Ψi is measured in a fixed local basis; the remaining unmeasured qubits in A are in a pure state that depends on the measurement outcome on B. b, For quantum systems consisting of qubits, the measurement on B samples random outcomes, each characterized as a bitstring zB (binary numbers in the blue/yellow arrows). Different measurement outcomes zB occur with probability p(zB ) and lead to distinct quantum states |ΨA(zB )i, forming the projected ensemble E = {p(zB ), |ΨA(zB )i}. Right panel: the ensemble of pure states |ΨA(zB )i (black arrows) are randomly distributed in the Hilbert space of A (black sphere), forming an approximate quantum state k-design. c, Examples of many- body wavefunctions whose projected ensembles form approximate quantum state designs include the typical output states of random unitary circuit evolution, quantum states obtained from quenched time evolution, and energy eigenstates of a chaotic Hamiltonian at infinite temperature. d, Illustration of minimal quantum state k-designs for a single qubit on the Bloch sphere. Forming higher k-designs requires a larger number of pair-wise non-orthogonal quantum states. In the limit of k → ∞, a k-design approaches the so-called Haar ensemble, which is the uniform distribution over all pure quantum states. states obtained by quenched time evolution and energy I. QUANTUM STATE DESIGNS FROM eigenstates of chaotic Hamiltonians (see Fig. 1c). In the PROJECTED ENSEMBLES former case, we find that the degree of randomness in a projected ensemble continues to grow even after con- Studying properties of random quantum states is use- ventional local thermalization has occurred. Specifically, ful because they often encode universal phenomena found we establish numerically that states evolved for a longer in nature. A standard approach is to consider the Haar amount of time form higher approximate k-designs. In ensemble, which is the uniform distribution over all pure the case of energy eigenstates, approximate k-designs states in a Hilbert space. While the Haar ensemble can be emerge from states in the middle of the energy spec- studied analytically using various statistical tools [20], it trum corresponding to effective infinite temperature. For is extremely challenging to experimentally realize, as do- finite-temperature eigenstates we observe that the pro- ing so requires an exponential amount of resources (such jected ensembles converge to a universal ensemble that as the number of quantum gates or experimental opera- smoothly varies with respect to the energy density. tions) [21]. Our findings suggest that for a wide range of physically Instead, ensembles that form quantum state designs relevant many-body states, projected ensembles exhibit a are considered because they mimic the Haar ensemble universal form of randomness. This allows us to quantify and can be efficiently realized in physical systems [22– the chaotic nature of Hamiltonian dynamics and to study 27]. To build intuition for quantum state designs, let the growth of complex, nonlocal correlations between a us consider quantum state designs for a single qubit subsystem and its complement beyond the conventional (see Fig. 1d). For example, the ensemble of single- paradigm of quantum thermalization [8–11, 13, 17, 18]. qubit states {|0i, |1i} with equal probabilities has the Post-thermalization physics in quantum many-body sys- same (first moment) as the Haar ensemble: ρ(1) = tems exhibits interesting quantitative and qualitative dif- 1 2 (|0ih0| + |1ih1|) = EΨ∼Haar[|ΨihΨ|], where EΨ∼E de- ferences from its classical counterpart (e.g. [19]), and the notes averaging |Ψi over the ensemble E. We say that current work clarifies and extends this paradigm. Fur- such a two-state ensemble forms a 1-design. However, the thermore, our work presents a simple protocol to effi- two-state ensemble does not form a 2-design because its ciently produce an ensemble of random pure states from 1 second moment 2 (|0ih0| ⊗ |0ih0| + |1ih1| ⊗ |1ih1|) differs fixed, time-independent Hamiltonian dynamics that does from that of the Haar ensemble. To form a 2-design for not require highly fine-tuned controls. a single qubit, one can use four distinct non-orthogonal quantum states uniformly spread over the Bloch sphere (Fig. 1d). In general, for an ensemble to form a higher order design, it must be supported over a larger number of states [28]. 3

Formally, given an ensemble E = {pi, |Ψii} consisting matrix, since the density matrix is the first moment of the of a set of wavefunctions |Ψii and their probabilities pi, ensemble. We will call |Ψi the generator state of EΨ,A. one can test whether the ensemble forms an approximate We note that Refs [29, 30] use a related construction to k-design by explicitly comparing its k-th moment define localizable entanglement. We find that projected ensembles form approximate (k) h ⊗ki X ⊗k ρE = E (|ΨihΨ|) = pi (|ΨiihΨi|) (1) k-designs for many physically relevant many-body wave- Ψ∼E i functions, including special, analytically soluble cases. (k) Specifically, it can be shown that a generator state for against that of the Haar ensemble ρHaar (which is de- an ε-approximate k-design can be efficiently prepared by fined by a continuous probability distribution over pure a simple protocol, in which an initial product state is states). More precisely, we say an ensemble E is an ε- evolved by a Hamiltonian over a short time duration. approximate quantum state k-design if The Hamiltonian can be time-independent and only has

(k) (k) nearest-neighbor Ising interactions. This is possible be- ρE − ρHaar ≤ ε , (2) 1 cause a universal resource state for measurement-based quantum computation can be prepared within a con- where k·k denotes the trace norm. This definition 1 stant time independent of system size [31, 32]. While that the k-th moment of E is nearly indistinguishable this example is fine-tuned, we show that, in fact, generic from the k-th moment of the Haar ensemble up to a small many-body wavefunctions are good generator states for error ε. It can be shown that an ε-approximate k-design ε-approximate k-designs: is also an ε-approximate j-design for any j < k, and accordingly larger values of k indicate that an ensemble Theorem 1. Let |Ψi be chosen uniformly at random looks more uniformly random. from the Hilbert space H. The ensemble EA,Ψ forms an Unlike the Haar ensemble, approximate designs arise ε-approximate k-design with probability at least 1 − δ if in physical settings [22–27]. A canonical example is ran- 1  1  dom unitary circuits [22, 23, 25], where a set of random NB = Ω k NA + log ε + log log δ . (6) two-qubit unitary gates in a certain geometric arrange- ment are sequentially applied to simple initial states (for Here Ω(·) denotes a lower bound up to a constant mul- an example, see Fig. 1c). Then the ensemble of resulting tiplicative factor and subleading terms. This theorem is states over different choices of unitary gates forms an ε- proved in AppendixC, and establishes that all but a tiny 2NB approximate k-design as long as the depth of the circuit is fraction (of order ∼ 1/2 ) of the states in the Hilbert sufficiently large, scaling polynomially in k, log(1/ε), and space are generator states for approximate k-designs if the number of qubits N [22]. Similarly, there are a num- NB is asymptotically larger than k times NA. However, a ber of proposals to generate approximate designs based quantum state randomly sampled from the entire Hilbert on time-dependent local Hamiltonian evolution [24, 27]. space is not so physical since such a state is extremely In these examples, states are sampled from approxi- difficult to produce experimentally [21]. To this end, we mate k-designs by realizing many distinct, highly engi- also present the following theorem: neered quantum evolutions. By contrast, we show that Theorem 2. Let |Ψi be a state sampled from an ensem- approximate k-designs arise naturally from the projected ble on H that forms an ε0-approximate k0-design. Then ensemble of a single many-body wavefunction. More pre- the projected ensemble EA,Ψ forms an ε-approximate k- cisely, consider a many-body wavefunction |Ψi in the design with probability at least 1 − δ if Hilbert space H for a bipartite system consisting of NA qubits in A and N qubits in B. The projected ensem- 1  B NB = Ω kNA + log εδ , (7) ble for A is generated by performing projective measure- 0 1  k = Ω k NB + log εδ , (8) ments on all NB qubits in B in a local basis {|zBi}, where NB NB 1  1  a bitstring zB ∈ {0, 1} enumerates over all 2 mea- log ε0 = Ω kNB NB + log εδ , (9) surement outcomes. After the measurement, with prob- 1  NA = Ω log(NB) + log(k) + log log εδ . (10) ability p(zB) the system is described by the normalized wavefunction |ΨA(zB)i ⊗ |zBi, where The proof is given in AppendixC and relies on higher- 1 order concentration of measure results [33] as well as a p(zB) := hΨ| ( A ⊗ |zBi hzB|) |Ψi (3) polynomial approximation technique used in quantum al-  p |ΨA(zB)i := 1A ⊗ hzB| |Ψi/ p(zB) (4) gorithms for solving linear systems [34]. This theorem shows that if the generator state is complex enough, in 1 and A is the identity operator acting on qubits in A. the sense that it is a typical state from an ε0-approximate This defines a projected ensemble: k0-design for small ε0 and large k0 [35], then the projected E := {p(z ) , |Ψ (z )i} . (5) ensemble will well-approximate a quantum k-design. Ψ,A B A B Theorem2 has implications for ongoing experiments: The ensemble consists of 2NB states which are generally if a single sample of an approximate design is experi- not pairwise orthogonal. The projected ensemble con- mentally realized, our theorem states that it can be used tains strictly more information than the reduced density to generate ensembles forming approximate designs on 4 a b 10 k=1 k=2 k=3 k=4

10

10

10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 Time Time Time Time c d 10 20 10

10 10

Design time Design 10 0 1 2 3 4 9 13 17 21 9 13 17 21 9 13 17 21 9 13 17 21 k NB NB NB NB

FIG. 2. Emergent quantum state designs from chaotic time evolution. a, Quenched dynamics under a time- independent Hamiltonian H starting from an initial product state, |0i⊗N , for a N-qubit system. The system is partitioned into two subsystems A and B with size NA and NB , respectively. At time t, a projective measurement in the local z-basis is (k) performed on subsystem B, resulting in a specific outcome zB of length NB . b, Trace distances ∆ between the k-th moments of the Haar ensemble and projected ensembles for an NA = 3 subsystem as a function of evolution time for various total system sizes N = 12, 14,... 24 (darker colors for increasing N). Dashed lines are a phenomenological power-law fit, yielding a scaling (k) −1.2 (k) of ∆ ∼ t . c, Design time τk defined by the evolution time to achieve a trace distance of ∆ = 0.02. We find that longer time evolution is required to form a higher approximate k-design. d, Late-time trace distances at t ≈ 103. For a mixed-field chaotic Hamiltonian (circles), the late-time trace distances exhibit an exponential scaling with NB while they remain nearly constant for an integrable, non-ergodic Hamiltonian (crosses, time traces not shown). Dashed lines represent the trace distances (k) NB ∆em between the k-th moments of the Haar ensemble and a finite set of 2 states sampled from the Haar ensemble on A. its subsystems. Moreover, this protocol is hardware- interacting fermions via the Jordan-Wigner transforma- efficient since it does not require fine-tuned controls to tion, leading to non-ergodic dynamics. However, for any produce many different quantum states, and could lead non-zero longitudinal field (hx 6= 0), the Hamiltonian is to various useful applications in quantum information sci- ergodic, and its eigenvalues and eigenvectors are expected ence [22, 36–40]. to have properties consistent with ETH predictions. This At a conceptual level, Theorem2 establishes that a has been explicitly checked for a specific parameter set large class of states which can be efficiently prepared are (hx, hy,J) = (0.8090, 0.9045, 1) [41] which we adopt for good generators of k-designs. This raises the possibility our study. We note that we do not find a qualitative that, even in natural chaotic quantum systems, approxi- differences in our results when using nearby parameter mate k-designs may arise from projected ensembles. values. We first consider the many-body state |Ψ(t)i = −iHQIMFt e |Ψ0i resulting from time evolution of the ini- II. QUANTUM STATE DESIGNS FROM ⊗N tial state |Ψ0i = |0i (Fig. 2a). Here |0ij and |1ij CHAOTIC SYSTEMS z are the eigenstates of σj with eigenvalues +1 and −1, respectively. The initial product state, |Ψ0i, has zero Motivated by the above results, we numerically investi- energy expectation value with respect to HQIMF, corre- gate projected ensembles that arise from chaotic Hamil- sponding to the total energy of an infinite temperature tonian dynamics. We start with the paradigmatic ex- state. This implies that local subsystems will relax to ample of the 1D quantum Ising spin system with mixed an infinite-temperature ensemble after a local thermal- fields (QIMF), described by the Hamiltonian ization time [8–10]. At any time t, the projected ensemble for a subsystem N N N−1 x X x y X y X x x A is obtained by simulating projective measurements on HQIMF = h σ + h σ + J σ σ , (11) j j j j+1 the rest of the NB = N − NA qubits in the local z- j=1 j=1 j=1 basis (Fig. 2a).1 In order to check if the projected en- µ semble forms an approximate k-design, we compare the where N is the number of spins, σj with µ = x, y, z are the Pauli operators for a spin at site j, J is the strength x y of Ising interactions, and h and h are the strengths 1 We note that our measurement z-basis is orthogonal to the of the longitudinal and transverse fields, respectively. In Hamiltonian; this choice is made to ensure that the measurement the absence of the longitudinal field (hx = 0), the Hamil- outcomes are not explicitly correlated with the total energy in tonian can be mapped to an integrable model of non- subsystem A. 5

(k) a k-th moment of the ensemble, ρE in Eq. (1), to that 1 k=1 k=2 k=3 of the uniform ensemble ρ(k) using the trace distance Haar (k) 1 (k) (k) ∆ = 2 ρE − ρHaar . For a fixed subsystem size of 1 0.1 (k) NA = 3, we numerically compute the trace distance ∆ up to k = 4 as a function of time for various NB (Fig. 2b). -1 0 1 -1 0 1 -1 0 1 In all cases, ∆(k) decays in time, following a phenomeno- E/N E/N E/N logical power-law scaling ∆(k) ∼ 1/t1.2, until it saturates b to a value that is governed by finite size effects. We note 0.5 that the extracted scaling exponent is non-universal and (k) Hamiltonian-dependent. The saturation value of ∆ 0.1 decreases exponentially with N√B (circles, Fig. 2d), ex- hibiting the scaling ∆(k) ∼ 1/ 2NB . If the power-law scaling persists at larger system sizes then ∆(k) may de- 7 9 11 13 7 9 11 13 7 9 11 13 NB NB NB crease over an exponentially long time scale. (k) As a comparison, we also present ∆em which is the FIG. 3. Emergent quantum state designs from energy trace distance between the k-th moments of the Haar eigenstates. a, Trace distances between the k-th moments ensemble and the empirical Haar ensemble consisting of of the Haar ensemble and a projected ensemble for an NA = 3 2NB states sampled uniformly at random on A (dashed subsystem generated from the energy eigenstates of a Hamil- lines, Fig. 2d). Since the empirical ensemble asymptoti- tonian. Results are presented as a function of energy density cally approaches the Haar ensemble in the limit of infinite E/N for a total system size of N = 14. b, Trace distances for (k) the projected ensembles obtained from eigenstates near zero samples, ∆em is determined only by statistical fluctua- energy corresponding to infinite temperature. The distances tions associated with having a finite number of quantum exhibit an exponential decay as a function of system size. The states. Remarkably, we find that the projected ensem- points are evaluated for 100 eigenstates near zero energy, and ble obtained from the quench dynamics shows a trace the error bars denote their . For compari- (k) distance almost identical to that of the empirical ensem- son, dashed lines represent the trace distances ∆em between ble of the same size, suggesting that the former is as the Haar ensemble and an empirical ensemble of 2NB states uniformly random as the latter. By contrast, repeating sampled from the Haar ensemble. similar calculations for the integrable model (hx = 0), we observe qualitatively different behavior where the trace as a function of system size for k = 1, 2, 3. However, the distance to the Haar ensemble is much larger than in decay rate is slightly slower than in the case of quenched the non-integrable case (crosses, Fig. 2d). Furthermore, dynamics (Fig. 3b). there is no appreciable dependence on system size. This In addition to the QIMF, we have also studied two is expected, since integrable systems do not locally ther- other ergodic Hamiltonian models in order to corroborate malize and instead relax to a generalized Gibbs ensem- the universality of our findings (see AppendixB). Specif- ble, and hence will not form 1-designs at effective infinite ically, we consider a system of random, all-to-all coupled temperature [42]. spin-1/2 particles as well as hard-core bosons with ran- Given the emergence of k-designs in asymptotic dom all-to-all hoppings with particle number conserva- regimes for a chaotic Hamiltonian, it is natural to ask how tion. In both cases, we observe excellent convergence of long it takes for a subsystem to achieve an approximate projected ensembles to approximate k-designs. In the design up to a small, fixed precision. To this end, we in- latter case, the measurement outcomes in B and corre- (k) sponding pure quantum states A are strongly correlated troduce a design time τk defined as the time at which ∆ becomes smaller than a certain fixed threshold ε. For a owing to the particle number conservation symmetry; chosen threshold ε = 0.02, we find that the formation of hence a na¨ıve approach based on Eq. (5) does not lead to higher k-designs requires longer time evolution (Fig. 2c). approximate k-designs. We instead introduce symmetry- This observation is consistent with the idea that typical resolved projected ensembles by grouping certain subsets quantum states from higher k-designs are more complex of measurement outcomes from B (see AppendixB for and hence more difficult to prepare [35, 43, 44]. details); this does lead to approximate k-designs. Next we investigate the properties of energy eigen- For chaotic Hamiltonians, projected ensembles forming states. We repeat a similar analysis as above by replacing k = 1 designs can be anticipated from the standard pic- the time-evolved state |Ψ(t)i with an energy eigenstate ture of quantum thermalization since the first moment of (k) a projected ensemble simply corresponds to the reduced |Eii of HQIMF. Figure 3 shows the trace distance ∆ as a function of energy E for a projected ensemble gener- density matrix of a subsystem. The reduced density ma- i trix approaching the first moment of the Haar ensemble, ated from |Eii. We find a sharp dip at zero energy den- sity corresponding to infinite temperature, which signi- i.e. the maximally mixed state, follows from local ther- fies the emergence of approximate quantum state designs malization at infinite temperature. However, the conver- (Fig. 3a). At zero energy, ∆(k) decreases exponentially gence of higher moments k ≥ 2 of the projected ensem- ble to higher k-designs is nontrivial and surprising. Such 6 a 0.2 0.8 b with system size (Fig. 4b inset). These observations sug- gest that the second moment of the projected ensemble 1 1 is indeed universal even at finite temperatures.

III. DISCUSSION AND OUTLOOK

j 0 E /N 0.1 0.1 The random quantum state ensembles considered in this paper are qualitatively different from conventional 6 8 10 12 -1 N ones in quantum statistical mechanics, such as the mi- B crocanonical and canonical ensembles. These latter en- -1 0 1 0 0.4 0.8 1.2 E /N sembles are fully specified by their corresponding density i |Ei -Ej |/N matrices (their first moment) and are used to evaluate FIG. 4. Universal ensemble at finite temperature. a, expectation values of observables. Trace distances between the second moments of projected en- Our formalism concerns more general statistical prop- sembles generated from a pair of energy eigenstates at Ei and erties (such as higher moments) of an ensemble with a Ej for NA = 3 subsystems. We plot the pairwise distances large number of pure states that are generally pairwise (2) ∆ij for every pair of eigenstates |Eii, |Ej i, computed for sys- non-orthogonal. As such, a projected ensemble encodes tem size N = 11. The distances are minimized when Ei ≈ Ej . additional information about a subsystem. For example, The plot suggests the existence of a universal ensemble that one can ask the following information-theoretic question: depends smoothly on the energy density. The black dashed how much information (i.e., classical bits) is required in line indicates a cut defined by (Ei + Ej )/N = −0.6. b, Trace order to specify the full wavefunction of a random sam- distances plotted as a function of energy density difference, ple from a projected ensemble? If the projected ensemble |Ei − Ej |/N, along the black dashed line in a for various were to be the Haar ensemble, an exponential number of systems sizes N = 9, 10,..., 15 (darker colors for increasing bits would be required. By contrast, for an ensemble N). Inset: the distances at zero energy difference Ei = Ej decay exponentially with system size. The trace distances which is uniformly distributed only over computational from the projected ensemble are comparable to those from basis states of A, a linear number of bits suffice to specify (2) a sample. Notice that the former and latter ensembles the finite-size ensemble, ∆em , of the empirical Haar distribu- tion (dashed line). Such an exponential scaling suggests the produce the same density matrix, namely the maximally existence of a universal random ensemble at finite tempera- mixed state. Therefore, projected ensembles provide a tures. novel framework to analyze the information content as- sociated with quantum states of a subsystem and their convergence cannot be explained by ETH alone, and sug- relation with the remainder of the system. gests a new form of emergent randomness beyond the Our findings open up a number of new directions in conventional framework of quantum thermalization. quantum chaos, thermalization, and quantum informa- A natural next step is to generalize quantum state de- tion. In particular, our numerical results demonstrate signs (or Haar ensembles) to a finite-temperature setting. that for an initial product state evolved by a chaotic However, we are unaware of any appropriate analogue of Hamiltonian, the largest k-design attained by the pro- the Haar ensemble at finite temperature. Such an en- jected ensemble grows as a function of time; moreover, semble, if it exists, would generally depend on the sys- this growth persists significantly past the local thermal- tem Hamiltonian, and its first moment should (approxi- ization timescale. While it is not presently clear how mately) be a thermal state. While explicitly identifying long this growth will persist, the observed growth is properties of such an ensemble is an interesting future clearly diagnostic of the sustained development of non- direction, here we find numerical evidence that such an local correlations after local thermalization has occurred ensemble exists. In Figure 4, we compute projected en- (e.g. [19, 45]). There may even be connections to the sembles for all energy eigenstates of H and present QIMF quantum complexity of a state evolving by chaotic dy- the pairwise distances namics, which likewise grows long after thermalization 1 has occurred [35, 46]. It would also be interesting to (2) (2) (2) ∆ij = ρi − ρj , (12) fully generalize the above to projected ensembles at fi- 2 1 nite temperature, in the presence of symmetries, (quasi- where ρ(2) denotes the second moment of a projected en- )integrability, or strong disorder resulting in localiza- i tion [11–13]. semble generated from an eigenstate |Eii. We find that In quantum information science, quantum states de- ∆(2) is a smooth function of energy up to small fluctua- ij signs are valuable resources in many applications. Our tions, suggesting that the projected ensembles smoothly (2) work establishes projected ensembles as a practical way vary with energy as well (Fig. 4a). Also, ∆ij is min- of sampling states from approximate designs using nat- imized when the energy difference |Ei − Ej| is small ural Hamiltonian evolutions of existing quantum simula- (2) (Fig. 4b), and in this regime ∆ij decreases exponentially tors without fine control. Further, our work could lead to 7 novel experimental quantum tomography protocols [40], tum simulators. cryptographic protocols for hiding information [22], the Acknowledgments. We thank Adam Bouland, Fer- design of unforgeable quantum encryption [36], and also nando Brand˜ao,Aram Harrow, Wen Wei Ho, Nicholas new methods for quantum device verification [39, 47]. Hunter-Jones, Anand Natarajan, and Hannes Pichler Indeed, a parallel work [14] uses projected ensembles to for valuable discussions. This work was partly sup- devise and implement a novel benchmarking protocol. ported by the Institute for Quantum Information and Matter, an NSF Physics Frontiers Center (NSF Grant Finally, an important question at the intersection of PHY- 1733907), the NSF CAREER award (1753386), the computer science and quantum many-body physics is AFOSR YIP (FA9550-19-1-0044), the DARPA ONISQ whether the computational complexity of simulating nat- program (W911NF2010021), the Army Research Office ural chaotic dynamics is beyond the capability of contem- MURI program (W911NF2010136), and the NSF QLCI porary classical computers. In other words, can quan- program (2016245). JSC is supported by a Junior Fel- tum supremacy tests be performed using a fixed chaotic lowship from the Harvard Society of Fellows, as well as Hamiltonian with analog quantum simulators? Exist- in part by the Department of Energy under grant DE- ing sampling-based quantum supremacy protocols heav- SC0007870. HH is supported by the J. Yang & Fam- ily rely on certain statistical and computational proper- ily Foundation. FH is supported by the Fannie & John ties of state ensembles formed by applying random uni- Hertz Foundation. JC acknowledges support from the taries to a fixed state [48–50]. The projected ensemble IQIM postdoctoral fellowship. ALS acknowledges sup- emerging from generic quantum dynamics may also pos- port from the Eddleman Quantum graduate fellowship. sess the requisite properties and our work could lead to a SC acknowledges support from the Miller Institute for new approach to quantum supremacy using analog quan- Basic Research in Science.

[1] Eugene P Wigner, “Characteristic vectors of bordered malization to statistical mechanics and thermodynam- matrices with infinite dimensions i,” in The Collected ics,” Advances in Physics 65, 239–362 (2016). Works of Eugene Paul Wigner (Springer, 1993) pp. 524– [13] Dmitry A. Abanin, Ehud Altman, Immanuel Bloch, and 540. Maksym Serbyn, “Colloquium : Many-body localization, [2] Freeman J Dyson, “Statistical theory of the energy levels thermalization, and entanglement,” Reviews of Modern of complex systems. i,” Journal of Mathematical Physics Physics 91, 021001 (2019). 3, 140–156 (1962). [14] Joonhee Choi, Adam L. Shaw, Ivaylo S. Madjarov, [3] Sandu Popescu, Anthony J Short, and Andreas Win- Xin Xie, Jacob Covey, Jordan Cotler, Daniel K. Mark, ter, “Entanglement and the foundations of statistical me- Hsin-Yuan Huang, Anant Kale, Hannes Pichler, Fer- chanics,” Nature Physics 2, 754–758 (2006). nando G.S.L Brand˜ao,Soonwon Choi, and Manuel En- [4] Peter Reimann, “Foundation of statistical mechanics un- dres, “Emergent Randomness and Benchmarking from der experimentally realistic conditions,” Physical Review Many-Body Quantum Chaos,” arXiv, same day (2021). Letters 101, 190403 (2008). [15] Joseph M Renes, Robin Blume-Kohout, Andrew J Scott, [5] Noah Linden, Sandu Popescu, Anthony J Short, and An- and Carlton M Caves, “Symmetric informationally com- dreas Winter, “Quantum mechanical evolution towards plete quantum measurements,” Journal of Mathematical thermal equilibrium,” Physical Review E 79, 061103 Physics 45, 2171–2180 (2004). (2009). [16] Andris Ambainis and Joseph Emerson, “Quantum t- [6] Anthony J Short and Terence C Farrelly, “Quantum equi- designs: t-wise independence in the quantum world,” in libration in finite time,” New Journal of Physics 14, Twenty-Second Annual IEEE Conference on Computa- 013063 (2012). tional Complexity (CCC’07) (IEEE, 2007) pp. 129–140. [7] M V Berry, “Regular and irregular semiclassical wave- [17] Adam M Kaufman, M Eric Tai, Alexander Lukin, functions,” Journal of Physics A: Mathematical and Gen- Matthew Rispoli, Robert Schittko, Philipp M Preiss, and eral 10, 2083–2091 (1977). Markus Greiner, “Quantum thermalization through en- [8] J. M. Deutsch, “Quantum statistical mechanics in a tanglement in an isolated many-body system,” Science closed system,” Physical Review A 43, 2046–2049 (1991). 353, 794–800 (2016). [9] Mark Srednicki, “Chaos and quantum thermalization,” [18] Masahito Ueda, “Quantum equilibration, thermalization Physical Review E 50, 888–901 (1994). and prethermalization in ultracold atoms,” Nature Re- [10] Marcos Rigol, Vanja Dunjko, and Maxim Olshanii, views Physics 2, 669–681 (2020). “Thermalization and its mechanism for generic isolated [19] Jordan Cotler, Nicholas Hunter-Jones, and Daniel quantum systems,” Nature 452, 854–858 (2008). Ranard, “Fluctuations of subsystem entropies at late [11] Rahul Nandkishore and David A. Huse, “Many-Body times,” arXiv:2010.11922 (2020). Localization and Thermalization in Quantum Statisti- [20] Aram W. Harrow, “The church of the symmetric sub- cal Mechanics,” Annual Review of Condensed Matter space,” arXiv:1308.6595 (2013). Physics 6, 15–38 (2015). [21] Emanuel Knill, “Approximation by quantum circuits,” [12] Luca D’Alessio, Yariv Kafri, Anatoli Polkovnikov, and arXiv:quant-ph/9508006 (1995). Marcos Rigol, “From quantum chaos and eigenstate ther- [22] Fernando GSL Brand˜ao,Aram W Harrow, and Michal 8

Horodecki, “Local random quantum circuits are approx- Review A 100, 032328 (2019). imate polynomial-designs,” Communications in Mathe- [39] Andreas Elben, Benoˆıt Vermersch, Rick van Bijnen, matical Physics 346, 397–434 (2016). Christian Kokail, Tiff Brydges, Christine Maier, Manoj K [23] Aram W. Harrow and Richard A. Low, “Random Quan- Joshi, Rainer Blatt, Christian F Roos, and Peter tum Circuits are Approximate 2-designs,” Communica- Zoller, “Cross-platform verification of intermediate scale tions in Mathematical Physics 291, 257–302 (2009). quantum devices,” Physical Review Letters 124, 010504 [24] Yoshifumi Nakata, Christoph Hirche, Masato Koashi, (2020). and Andreas Winter, “Efficient quantum pseudorandom- [40] Hsin-Yuan Huang, Richard Kueng, and John Preskill, ness with nearly time-independent hamiltonian dynam- “Predicting many properties of a quantum system from ics,” Physical Review X 7, 021006 (2017). very few measurements,” Nature Physics 16, 1050–1057 [25] Aram Harrow and Saeed Mehraban, “Approximate (2020). unitary t-designs by short random quantum cir- [41] Hyungwon Kim, Tatsuhiko N Ikeda, and David A Huse, cuits using nearest-neighbor and long-range gates,” “Testing whether all eigenstates obey the eigenstate ther- arXiv:1809.06957 (2018). malization hypothesis,” Physical Review E 90, 052105 [26] Jonas Haferkamp, Felipe Montealegre-Mora, Markus (2014). Heinrich, Jens Eisert, David Gross, and Ingo Roth, [42] Lev Vidmar and Marcos Rigol, “Generalized Gibbs en- “Quantum homeopathy works: Efficient unitary designs semble in integrable lattice models,” Journal of Statis- with a system-size independent number of non-clifford tical Mechanics: Theory and Experiment 2016, 064007 gates,” arXiv:2002.09524 (2020). (2016). [27] Tom Farshi, Daniele Toniolo, Carlos E Gonz´alez-Guill´en, [43] Daniel A Roberts and Beni Yoshida, “Chaos and com- Alvaro´ M Alhambra, and Lluis Masanes, “Time- plexity by design,” Journal of High Energy Physics 2017, periodic dynamics generates pseudo-random unitaries,” 121 (2017). arXiv:2007.03339 (2020). [44] Jordan Cotler, Nicholas Hunter-Jones, Junyu Liu, and [28] Philippe Delsarte, Jean-Marie Goethals, and Johan Ja- Beni Yoshida, “Chaos, complexity, and random matri- cob Seidel, “Spherical codes and designs,” in Geometry ces,” Journal of High Energy Physics 2017, 48 (2017). and Combinatorics (Elsevier, 1991) pp. 68–93. [45] Jordan Cotler and Nicholas Hunter-Jones, “Spectral de- [29] Frank Verstraete, Markus Popp, and J Ignacio Cirac, coupling in many-body quantum chaos,” Journal of High “Entanglement versus correlations in spin systems,” Energy Physics 2020, 1–62 (2020). Physical Review Letters 92, 027901 (2004). [46] Adam R Brown and Leonard Susskind, “Second law of [30] Markus Popp, Frank Verstraete, Miguel Angel Mart´ın- quantum complexity,” Physical Review D 97, 086015 Delgado, and J Ignacio Cirac, “Localizable entangle- (2018). ment,” Physical Review A 71, 042306 (2005). [47] Jose Carrasco, Andreas Elben, Christian Kokail, Barbara [31] Robert Raussendorf and Hans J. Briegel, “A one-way Kraus, and Peter Zoller, “Theoretical and experimental quantum computer,” Physical Review Letters 86, 5188– perspectives of quantum verification,” PRX Quantum 2, 5191 (2001). 010102 (2021). [32] Rawad Mezher, Joe Ghalbouni, Joseph Dgheim, and [48] Scott Aaronson and Alex Arkhipov, “The computational Damian Markham, “Efficient quantum pseudorandom- complexity of linear optics,” in Proceedings of the forty- ness with simple graph states,” Physical Review A 97, third annual ACM symposium on Theory of computing 022333 (2018). (2011) pp. 333–342. [33] Sergey G Bobkov, Friedrich G¨otze, and Holger Sambale, [49] Juan Bermejo-Vega, Dominik Hangleiter, Martin “Higher order concentration of measure,” Communica- Schwarz, Robert Raussendorf, and Jens Eisert, “Ar- tions in Contemporary Mathematics 21, 1850043 (2019). chitectures for quantum simulation showing a quantum [34] Andrew M Childs, Robin Kothari, and Rolando D speedup,” Physical Review X 8, 021010 (2018). Somma, “Quantum algorithm for systems of linear equa- [50] Adam Bouland, Bill Fefferman, Chinmay Nirkhe, and tions with exponentially improved dependence on pre- Umesh Vazirani, “On the complexity and verification of cision,” SIAM Journal on Computing 46, 1920–1950 quantum random circuit sampling,” Nature Physics 15, (2017). 159–163 (2019). [35] Fernando GSL Brand˜ao,Wissam Chemissany, Nicholas Hunter-Jones, Richard Kueng, and John Preskill, “Mod- els of quantum complexity growth,” arXiv:1912.04297 (2019). [36] Gorjan Alagic, Tommaso Gagliardoni, and Christian Majenz, “Unforgeable quantum encryption,” in Annual International Conference on the Theory and Applications of Cryptographic Techniques (Springer, 2018) pp. 489– 519. [37] Christoph Dankert, Richard Cleve, Joseph Emerson, and Etera Livine, “Exact and approximate unitary 2-designs and their application to fidelity estimation,” Physical Re- view A 80, 012304 (2009). [38] Andrew W Cross, Lev S Bishop, Sarah Sheldon, Paul D Nation, and Jay M Gambetta, “Validating quantum computers using randomized model circuits,” Physical 9

Appendix A: Details of numerical simulations

1. k-th Moment of the Haar Ensemble

To quantify the degree of randomness of our projected ensembles, we computed the trace distance ∆(k) between the k-th moments of the projected ensemble and the Haar ensemble: 1 (k) (k) (k) ∆ = ρE − ρHaar . (A1) 2 1 (k) Here ρHaar is the k-th moment averaged over the Haar ensemble. For a Hilbert space H with dimension d, it has the form [20]:

(k)  ⊗k ρHaar = E (|ΨihΨ|) (A2) Ψ∼Haar(d) P PermH⊗k (π) = π∈Sk . (A3) d(d + 1) ··· (d + k − 1)

Here, Sk is the symmetric group on k elements, and π ∈ Sk is an element of the group. PermH⊗k is a representation ⊗k of Sk on H which permutes the tensor factors according to

PermH⊗k (π) |Ψ1i ⊗ · · · ⊗ |Ψki = |Ψπ−1(1)i ⊗ · · · ⊗ |Ψπ−1(k)i . (A4) 0 0 The inverses in the subscripts are chosen so that PermH⊗k (π)·PermH⊗k (π ) = PermH⊗k (π◦π ) (i.e., the representation is a homomorphism of Sk, and not an antihomomorphism). It is readily checked that Eq. (A2) can be written as P PermH⊗k (π) Π π∈Sk = k (A5) d(d + 1) ··· (d + k − 1) d+k−1 k ⊗k where Πk is simply the projector onto the symmetric subspace of H (that is, the invariant subspace under Sk); this d+k−1 subspace has dimension k .

2. Finite sampling error from the empirical Haar ensemble

NB In the main text, we constructed projected ensembles of size 2 , with NB the size of the complement subsystem. (k) (k) In Figs. 2d, 3b, and 4b, we compared the system size scaling of ∆ against the trace distance ∆em of the empirical Haar ensemble, an ensemble formed by sampling from the Haar ensemble 2NB times. In particular, we have

2NB (k) 1 X ⊗k ρem = (|ΨiihΨi|) , |Ψii ∼ Haar(d) (A6) 2NB i=1   (k) 1 (k) (k) ∆em = E kρem − ρHaark1 (A7) Ψ1,Ψ2,...∼Haar(d) 2

(k) This comparison is made to estimate the degree to which the error ∆ in our projected ensemble√ is due to its finite (k) (k) N size. We estimate ∆em for various values of NB and find that ∆em scales exponentially as 1/ 2 B .

Appendix B: Additional numerical simulations of ergodic Hamiltonian models

1. Random coupling model

In the main text we presented results for the ergodic QIMF model. Here we discuss an additional ergodic model: a spin-1/2 model with random all-to-all interactions,

N X X µ,ν µ ν H = Jij σi σj , (B1) i,j=1 µ,ν∈{x,y,z} i

a c

b d

FIG. 5. Emergent quantum state designs in random coupling and random hopping models. a, Time evolution and b, eigenstate data for the random all-to-all coupling model in Eq. (B1). The trace distances ∆(k) for k = 1, 2 and 3 are plotted in orange, red and purple respectively, for NA = 3 subsystems (a, left). Time evolution data for N = 21 is plotted, along with the system size scaling of the late-time ∆(k) for N = 8 to 21 (a, right). We also plot ∆(k) for each eigenstate for N = 14 (b, left), along with the system size scaling of the eigenstate data near E = 0 for N = 10 to 14 (b, right). c, Time evolution and d, eigenstate data for the random hopping model in Eq. (B2). Data is presented for NA = 5 subsystems, in z the Stot. = 0 sector. In c, time evolution data for N = 24 is plotted along with system size scaling for N = 10, 12, ..., 24. To illustrate the failure of projecting onto all strings zB , we plot the results of this na¨ıve procedure with NA = 3 in light colors and dashes. In d, eigenstate data for N = 16 is plotted, along with its system size scaling for N = 12, 14, 16 and 18. System size scaling is plotted against the effective NB , which is log2 of the number of strings zB which is post-selected. In each plot, the error bars are smaller than the marker sizes. µ,ν µ,ν where Jij are random variables drawn from i.i.d. normal distributions: Jij ∼ N(0, 1/N). Such random coupling models are paradigmatic examples of quantum chaotic systems. The of the couplings is chosen so that 2 N z z ⊗N Tr(H )/2 ∼ O(N). Our model does not have σi σj terms so that our initial state |Ψ0i = |0i has zero energy, and accordingly is regarded to be at infinite temperature. As shown in Fig. 5a, we see excellent convergence to k-designs for our projected ensembles constructed from (k) (k) time-evolved quenched states as well as eigenstates. The late-time ∆ values are very close to ∆em . The average (k) (k) ∆ values for infinite-temperature eigenstates are also close to ∆em , and show clear exponential decay (Fig. 5b). Notably, we do not average over disorder realizations: each data point and time series are computed with fixed disorder realizations. This provides additional support for our conjecture that ergodic Hamiltonian systems generate approximate k-designs via time-evolved states and eigenstates at infinite temperature.

2. Random hopping model with U(1) symmetry

Having established that chaotic models such as the QIMF and the all-to-all random coupling model provide projected ensembles which converge to k-designs, we next ask whether quantum models with symmetries do so as well. We specifically study a random hopping model

N X + x x y y − x y y x H = Jij σi σj + σi σj + Jij σi σj − σi σj (B2) i,j=1 i

± ± where Jij are random variables drawn from i.i.d. normal distributions: Jij ∼ N(0, 1/N). Like the all-to-all random coupling model, this model is also expected to exhibit chaos. However, the specific interaction terms are chosen such z P z that this model has a U(1) symmetry: there is conservation of the total magnetization Stot. = j Sj . The model can equivalently be viewed as describing hard-core bosons with random complex all-to-all hopping amplitudes. 11

Any bitstring in the z-basis is a suitable infinite-temperature quench state. For N even, we choose the initial z state |Ψ0i = |0101 ··· 01i which lies in the Stot. = 0 sector. Under time evolution by our chaotic Hamiltonian, the z wavefunction is ergodic in the Stot. = 0 Hilbert space. Na¨ıvely forming the projected ensemble, we do not find convergence to an approximate k-design. This is shown in the light curves in Fig. 5c. With increasing system size, ∆(k) for k = 1, 2, 3 saturate at a non-zero value. This is because of large correlations between the bitstrings zB and the projected state |ΨA(zB)i: if zB has total magnetization z sB, |ΨA(zB)i necessarily has total magnetization Stot. − sB due to the global conservation law. The Hilbert space HA naturally decomposes into a direct sum of multiple sectors enumerated by the magnetization sA. Accordingly, the projected ensemble now produces multiple ensembles, one for each sector. In order to properly account for the U(1) conservation law, instead of projecting onto all strings zB, we post-select a subset of all strings: bitstrings with fixed total magnetization, e.g. sB = 1/2. The projected states |ΨA(zB)i will also have fixed magnetization, and we then ask whether this ensemble forms a k-design in the subspace of z HA with magnetization sA = Stot. − sB. In our numerical examples, we present results with NA = 5 qubits and NA sA = −1/2, sB = 1/2. The relevant subspace of HA has dimension 10, far smaller than 2 = 32. Fig. 5c,d show the results of our symmetry resolution. We now see excellent convergence towards a k-design. As (k) (k) with the random coupling model, the late-time ∆ values are very close to ∆em . In order to make a fair comparison, (k) (k) we plot the late-time ∆ against an “effective NB”, defined as log2 of the number of post-selected strings zB. ∆em is also computed by sampling the same number of times. Using this post-selection procedure, we find excellent k-design convergence for projected ensembles from eigenstates near E = 0. Our results indicate that our basic approach remains valid for chaotic models with additional symmetries if these symmetries are properly addressed.

Appendix C: Proof of main theorems

1. Proof sketch for the main theorems

We begin with a sketch of the proof of Theorem1. We will use streamlined notation for clarity. Recall that for a generator state |Φi on H, the projected ensemble EΦ,A = {pz , |Φzi} has

pz = hΦ|Pz|Φi (C1)  √ |Φzi = 1A ⊗ hz|B |Φi/ pz . (C2)

It will be convenient to define the unnormalized state

 NB |Φezi := 1A ⊗ hz|B |Φi for z ∈ {0, 1} (C3) √ so that pz = hΦez|Φezi and |Φzi = |Φezi/ pz. We note that in our new notation, we can write

h ⊗ki E |ΨihΨ| (C4) Ψ∼EΦ,A X ⊗k = pz |ΦzihΦz| (C5) z∈{0,1}NB ⊗k X |ΦezihΦez| = ≡ A(|Φi) . (C6) hΦ |Φ ik−1 z∈{0,1}NB ez ez

The key to establishing Theorem1 is understanding the random tensor A(|Φi) when |Φi is a Haar-random state. Here, the Haar-random state is a normalized vector chosen uniformly at random in d = dAdB complex dimensions, N NA NB where d = 2 , dA = 2 , dB = 2 are the dimensions of the Hilbert space with N, NA, and NB qubits respectively. Two ingredients are needed to understand A(|Φi): 1. The expectation value of A(|Φi) over the Haar ensemble. 2. The concentration of A(|Φi) around its expectation. This will tell us that with high probability, A(|Φi) is close to its expectation. 12

Let us discuss these two ingredients in more detail. Using the fact that the uniform measure (i.e., the Haar measure) on the complex sphere is invariant under any unitary rotations, we can show that pz and |ΦzihΦz| are independent random variables. Furthermore, |ΦzihΦz| is a uniform random vector from a dA-dimensional complex sphere. Hence, we have ⊗k E A(|Φi) = E |ΨihΨ| (C7) Φ∼Haar(d) Ψ∼Haar(dA) which is the k-th moment of the uniform ensemble. This means that the expectation of the k-th moment of the projected ensemble of a randomly selected many-body wavefunction |Φi reproduces the k-th moment of the Haar ensemble in dA complex dimensions. Hence the expectation EΦ∼Haar(d) A(|Φi) is exactly equal to the desired quantity. To understand the concentration of A(|Φi) around its expectation, we recall the well-known result that the uni- form distribution over a high-dimensional sphere has a very sharp concentration. For any Lipschitz function on the sphere, the fluctuation around its expectation value is small and the probability of having a large fluctuation decays exponentially. This is known as Levy’s lemma and allows us to upper bound the probability that A(|Φi) is far from its expectation. Together with the expectation identity, we have    d ε2  Prob A(|Φi) − (|ΨihΨ|)⊗k ≥ ε ≤ 2d2k exp − B . (C8) E A 3 4k Φ∼Haar(d) Ψ∼Haar(dA) 1 18π (2k − 1) dA The asymptotic relation in Theorem1 follows immediately from the above probabilistic statement. The detailed proof of Theorem1 is given in AppendixC3. The proof of Theorem2 is very different from Theorem1. First of all, the expectation of A(|Φi) in Theorem1 is computed using the invariance property of the Haar measure, which does not hold for a measure that only forms a state design. Furthermore, Levy’s lemma only holds for the Haar measure, so we cannot resort to Levy’s lemma to control its statistical fluctuations. To prove Theorem2, we make use of a technique used in the context of solving linear systems on a quantum computer [34]. The key idea is to add a modulating function that approximates A(|Φi) by a polynomial function in |Φi. In particular, ⊗k X |ΦezihΦez| A(|Φi) = (C9) hΦ |Φ ik−1 z∈{0,1}NB ez ez ⊗k X |ΦezihΦez| ≈ B(|Φi) = µk,b(dBhΦez|Φezi), (C10) hΦ |Φ ik−1 z∈{0,1}NB ez ez where the modulating function is given by

2(k−1) b µk,b(s) = (1 − (1 − s ) ) for b even. (C11)

Using the binomial expansion, we can check that B(|Φi) is a polynomial function. Furthermore, µk,b(s) is very close to one when s = dBhΦez|Φezi is around one. Taking b larger allows us to better approximate the constant function µk,b(s) = 1, which corresponds to the target expression A(|Φi). A visualization of µk,b(s) can be found in Fig.6a. If |Φi is sampled from the uniform measure on the quantum state space, then s = dBhΦez|Φezi will concentrate around one; see Fig.6b. So for |Φi sampled from the uniform measure, we can show that A(|Φi) ≈ B(|Φi). However, |Φi is sampled from a quantum state design, so there is no guarantee that s = dBhΦez|Φezi should concentrate around one. To address this, we utilize the following result proved in Lemma5

kA(|Φi) − B(|Φi)k1 ≤ R(|Φi), (C12) where R(|Φi) is a polynomial function given by X    hΦez|Φezi 1 − µk,b dBhΦez|Φezi . (C13) z∈{0,1}NB Hence, the approximation error can be upper bounded by a polynomial function that we have better control over. After introducing the key quantities, the proof proceeds by bounding the error between (i) A(|Φi) where |Φi is sampled from a state design, and (ii) the expectation of A(|Ψi) where |Ψi is from the Haar measure. We use B(|Φi) as an intermediate point of comparison:

E A(|Φi) − E A(|Ψi) (C14) Φ∼(ε0,k0) design Ψ∼Haar 1 13

a b

FIG. 6. Visualization of the modulating function and the concentration of dB hΦez|Φezi. a, The modulating function µk,b(s) defined in Eq. (C11) for varying values of b from 2 to 16. Darker colors correspond to higher b. We fix k = 2 in the NA plot. b, The concentration of s = dB hΦez|Φezi for the Hilbert space dimension of subsystem A, i.e., dA = 2 = 16, when |Φi is sampled uniformly at random from the complex sphere. We sample 10000 different vectors |Φi to generate the histogram. The black solid line is a kernel density estimation fit of the histogram. If dA is larger, then s is more concentrated.

≤ E kA(|Φi) − B(|Φi)k1 + E B(|Φi) − E A(|Ψi) . (C15) (ε0,k0) design (ε0,k0) design Haar 1 The first term can be upper bounded by R(|Φi) and we can use the fact that state designs approximate the expectation of any polynomial function under the Haar measure. Accordingly,

E kA(|Φi) − B(|Φi)k1 ≤ E R(|Φi) ≈ E R(|Φi). (C16) (ε0,k0) design (ε0,k0) design Haar

In the second term, we can apply a similar idea by upper bounding the 1-norm with the 2-norm and utilizing the fact that B(|Φi) − EHaar A(|Ψi) is a polynomial function in |Φi. This gives s 2 k E B(|Φi) − E A(|Ψi) ≤ dA E B(|Φi) − E A(|Ψi) (C17) (ε0,k0) design Haar 1 (ε0,k0) design Haar 2 r 2 k ≈ dA E B(|Φi) − E A(|Ψi) . (C18) Haar Haar 2 We then approximate B(|Φi) by A(|Φi) in the above expression, where the error can be upper bounded by R(|Φi)2. This gives the approximate relation s  2 k 2 E B(|Φi) − E A(|Ψi) . dA E R(|Φi) + A(|Φi) − E A(|Ψi) . (C19) (ε0,k0) design Haar 1 Haar Haar 2

Using these steps, all our expressions contain only expectations over the Haar measure. More precisely,

E A(|Φi) − E A(|Ψi) (C20) Φ∼(ε0,k0) design Ψ∼Haar 1 r 2 q k 2 k . E R(|Φi) + dA E R(|Φi) + dA E A(|Φi) − E A(|Ψi) . (C21) Haar Haar Haar Haar 2 The first two terms are small because

1. µk,b(s) is close to one when s is close to one.

2. If |Φi is sampled from the Haar measure, then s = dBhΦez|Φezi will concentrate around one. See Figure6 for visualizations. The third term is small due to Theorem1. Finally, we can resort to Markov’s inequality to show that with a probability of at least 0.99, we have that kA(|Φi) − EΨ∼Haar A(|Ψi)k1 is small. The full proof of Theorem2 is given in AppendixC4. 14

2. Useful identities

We collect here several useful identities that we will leverage throughout the remainder of the Appendix. The first identity is P Perm ⊗k (π) h ⊗ki π∈Sk HA E (|ΦezihΦez|) = . (C22) Φ∼Haar(d) d(d + 1) ··· (d + k − 1) Taking the trace, we find

h ki dA(dA + 1) ··· (dA + k − 1) E hΦez|Φezi = . (C23) Φ∼Haar(d) d(d + 1) ··· (d + k − 1) The related identity

h 2ki k! dA(dA + 1) ··· (dA + k − 1) E |hΦez|Φeyi| = for x 6= y (C24) Φ∼Haar(d) d(d + 1) ··· (d + 2k − 1) will likewise be useful.

3. Proof of Theorem1

The main ingredient to prove Theorem1 is Levy’s lemma. Levy’s lemma states that if a function is Lipschitz continuous, then it will concentrate around it’s expectation value if the variable is sampled uniformly from a high- dimensional complex sphere.

2d−1 Lemma 1 (Levy’s Lemma). Let f : S → R satisfying |f(v) − f(w)| ≤ ηkv − wk2 . Then for any δ ≥ 0, we have    2  2d δ ProbΦ∼Haar(d) f(Φ) − E [f(Ψ)] ≥ δ ≤ 2 exp − . (C25) Ψ∼Haar(d) 9π3η2 Before applying Levy’s lemma, we will need to define a function of interest. In particular, we will consider individual P ⊗k ⊗k entries in the main quantity z∈{0,1}NB pz |ΦzihΦz| . Let {|ii}i∈{0,1}tNA be the standard basis on HA , and write Nt (k) (k) Nt (k) 2d−1 |ii = k=1 |i i where each |i i ∈ HA. Similarly write |ji = k=1 |j i. Consider the function fij : S → R defined by

 ⊗k  X |ΦezihΦez| fij(|Φi) = hi|   |ji (C26) hΦ |Φ ik−1 z∈{0,1}NB ez ez Qk (`) (`) X hi |ΦezihΦez|j i = `=1 . (C27) hΦ |Φ ik−1 z∈{0,1}NB ez ez

A nice property of this function is that it is Lipschitz continuous and hence Levy’s lemma can be applied. Lemma 2 (Lipschitz constant). We have

d fij(|Φi) ≤ 2(2k − 1) = η. (C28) d|Φi 2

Proof. Since our fij is differentiable, we can choose any η such that

d η ≥ fij(|Φi) . (C29) d|Φi 2 We have

k Q hi(`)|Φ ihΦ |j(`)i d X X `6=`0 ez ez d (`0) (`0) fij(|Φi) = hi |ΦezihΦez|j i d|Φi hΦ |Φ ik−1 d|Φi 2 `0=1 z∈{0,1}NB ez ez 15

Qk (`) (`) X `=1hi |ΦezihΦez|j i d − (k − 1) hΦez|Φezi hΦ |Φ ik d|Φi z∈{0,1}NB ez ez 2

k Q hi(`)|Φ ihΦ |j(`)i X X `6=`0 ez ez d (`0) (`0) ≤ hi |ΦezihΦez|j i hΦ |Φ ik−1 d|Φi `0=1 z∈{0,1}NB ez ez 2 Qk (`) (`) X `=1hi |ΦezihΦez|j i + 2(k − 1) Pz|Φi hΦ |Φ ik z∈{0,1}NB ez ez 2 (C30) where we have simply explicitly evaluated the derivatives and used the triangle inequality for the 2-norm. Let us bound each of the term terms on the right-hand side in turn. For the first one, we have

k Q hi(`)|Φ ihΦ |j(`)i X X `6=`0 ez ez d (`0) (`0) hi |ΦezihΦez|j i hΦ |Φ ik−1 d|Φi `0=1 z∈{0,1}NB ez ez 2

k Q hi(`)|Φ ihΦ |j(`)i X X `6=`0 ez ez  (`0) (`0) (`0) (`0)   = |j ihi | + |i ihj | ⊗ |zihz|B |Φi . (C31) hΦ |Φ ik−1 `0=1 z∈{0,1}NB ez ez 2

(`) (`) Q 0 0 0 0 `6=`0 hi |Φez ihΦez |j i (` ) (` ) (` ) (` ) 0 0 Writing b` ,x = k−1 and M` = |j ihi | + |i ihj | (which is Hermitian), the above evaluates to hΦez |Φez i

1/2  k  X X ∗  †   bp0,xb`0,xhΦ| Mp0 M`0 ⊗ |zihz|B |Φi `0,p0=1 z∈{0,1}NB 1/2  k  X X ∗  †  ≤  |bp0,x| |b`0,x| hΦ| |Mp0 M`0 | ⊗ |zihz|B |Φi (C32) `0,p0=1 z∈{0,1}NB √ where |A| := A†A. But since    ⊗(k−1) ! ⊗(k−1)  O (`) (`) |ΦezihΦez|  O (`) (`) |ΦezihΦez| |b`0,x| = tr  |j ihi | · ≤ |j ihi | = 1 , k−1 k−1  `6=`0 hΦez|Φezi  `6=`0 hΦez|Φezi 2 2 Eq. (C32) is less than or equal to  1/2  1/2 k   t   X X † X † 1  hΦ| |Mp0 M`0 | ⊗ |zihz|B |Φi =  hΦ| |Mp0 M`0 | ⊗ B |Φi `0,p0=1 z∈{0,1}NB `0,p0=1 1/2  t  X † ≤  kMp0 M`0 k∞ . (C33) `0,p0=1

† Since kMp0 M`0 k∞ ≤ 4, the above is ≤ 2t. Qk (`) (`) `=1hi |Φez ihΦez |j i Now we bound the term in the last line of Eq. (C30). Letting cz = k , the term can be written as hΦez |Φez i  1/2 X 2 2(k − 1)  |cz| hΦ|Pz|Φi . (C34) z∈{0,1}NB Using the bound

( k ! ⊗k !) k ⊗k O (`) (`) |ΦezihΦez| O (`) (`) |ΦezihΦez| |cz| = tr |j ihi | · ≤ |j ihi | = 1 , hΦ |Φ ik hΦ |Φ ik `=1 ez ez `=1 2 ez ez 2 16 we can upper bound Eq. (C34) by

 1/2 X 2(k − 1)  hΦ|Pz|Φi = 2(k − 1) . (C35) z∈{0,1}NB

Putting our bounds together, we have

d fij(|Φi) ≤ 2t + 2(k − 1) = 2(2k − 1) , (C36) d|Φi 2 which is the desired result.

While the above bound on the Lipschitz constant, along with Levy’s lemma, guarantee that each function fij concentrates around the its expectation value, the expectation value of fij has not yet been computed. The following lemma shows that the expectation value of fij is closely connected to the quantum state k-design on subsystem A. Lemma 3 (Expectation value identity). We have the identity   ⊗k X (|ΦezihΦez|)  ⊗k E  k−1  = E (|ΨihΨ|) . (C37) Φ∼Haar(d) Ψ∼Haar(dA) z∈{0,1}NB hΦez|Φezi

Therefore the expectation of fij is given by

 ⊗k E [fij(|Φi)] = hi| E (|ΨihΨ|) |ji (C38) Φ∼Haar(d) Ψ∼Haar(dA)

Proof. Observe that

⊗k (|ΦezihΦez|) ⊗k k−1 = (|ΦzihΦz|) hΦez|Φezi . (C39) hΦez|Φezi

Furthermore, from a technical lemma given below, we have that |Φzi is independent from hΦez|Φezi. Using this fact and the linearity of expectation, we compute   ⊗k X (|ΦezihΦez|) X  ⊗k h i E   = E (|ΦzihΦz|) E hΦez|Φezi Φ∼Haar(d) k−1 Φ∼Haar(d) Φ∼Haar(d) z∈{0,1}NB hΦez|Φezi z∈{0,1}NB   (C40)  ⊗k X 1 = E (|ΨihΨ|)   Ψ∼Haar(dA) dB z∈{0,1}NB as desired.

During the evaluation of the expectation value, we used a property that the normalized state |ΦzihΦz| and the corresponding probability pz are independent random variables. This is proven in the following lemma.

Lemma 4. If |Φi is a Haar-random state on H, then the random variables |ΦzihΦz| and pz are independent.

Proof. Define the map Pz(|Φi) := hΦ|Pz|Φi, and let Rz be the map taking |Φi to the normalized state |Φzi,

Rz(|Φi) := trB((1A ⊗ |zihz|B) · |ΦihΦ|) /Pz(|Φi) . (C41)

Let UA be a Haar-random unitary operator on HA and let U = UA ⊗1B. Because U is unitary and |Φi is Haar-random, the state U |Φi is also Haar-random. Also note that

† Pz(U |Φi) = Pz(|Φi) , Rz(U |Φi) = UARz(|Φi)UA (C42) 17 where the equivalence is in the sense of random variables. Therefore, for any functions F and G,

E [F (Rz(|Φi)) G(Pz(|Φi)] = E [F (Rz(U |Φi)) G(Pz(|Φi)] Φ∼Haar(d) Φ∼Haar(d) UA∼U(dA) h † i = E F (UARz(|Φi)UA) G(Pz(|Φi) (C43) Φ∼Haar(d) UA∼U(dA)

= E [F (|Ψi hΨ|)] E [G(Pz(|Φi)] . Ψ∼Haar(dA) Φ∼Haar(d)

To pass from the second line to the third, we used the fact that for any state |Ψi on HA, the state UA |Ψi is an independent Haar-random state on HA. This shows that Rz(|Φi) and Pz(|Φi) are independent, as desired. With the above lemmas, we now proceed with the proof of Theorem1.

Proof of Theorem1. Let us consider the function fij(|Φi) defined in Eq. (C26). Because we have bound the Lipschitz constant η ≤ 4t − 2, we can leverage Levy’s lemma given in Lemma1 to obtain

   2  d ε ProbΦ∼Haar(d) fij(|Φi) − E [fij(|Φi)] ≥ ε ≤ 2 exp − . (C44) Φ∼Haar(d) 18π3(2k − 1)

2k Performing a union bound and rescaling ε → ε/dA , we have  ε   d ε2  Prob f (|Φi) − [f (|Φi)] ≥ , for some i, j ≤ 2d2k exp − . (C45) Φ∼Haar(d) ij E ij 2k A 3 4k Φ∼Haar(d) dA 18π (2k − 1) dA

By comparing to the definition of fij in Eq. (C26) and using Lemma3 to obtain the expectation value of fij, the above is equivalent to the following concentration inequality

    ⊗k ⊗k X |ΦezihΦez| X |ΦezihΦez| ProbΦ∼Haar(d) − E   ≥ ε  k−1 Φ∼Haar(d) k−1  N hΦez|Φezi N hΦez|Φezi z∈{0,1} B z∈{0,1} B entrywise, 1  d ε2  ≤ 2d2k exp − (C46) A 3 4k 18π (2k − 1) dA P where kAkentrywise, 1 = i,j |Aij|. Finally, using kAkentrywise, 1 ≥ kAk1 and applying Lemma3, we find   ⊗k  2  X |ΦezihΦez| dB ε Prob − (|ΨihΨ|)⊗k ≥ ε ≤ 2d2k exp − . Φ∼Haar(d) E  A 3 4k k−1 Ψ∼Haar(dA) 18π (2k − 1) d N hΦez|Φezi A z∈{0,1} B 1 (C47)

We can see that if we have 18π3(2k − 1)d4k d ≥ A (2t log(d ) + log(2/δ)) , (C48) B ε2 A

NA then the ensemble EΦ,A forms an ε-approximate t-design with probability at least 1 − δ. Recall that dA = 2 , dB = 2NB , hence taking a logarithm on both side of the above condition gives rise to

 1 1 N = Ω kN + log + log log , (C49) B A ε δ which is the result stated in the main text. 18

4. Proof of Theorem2

For convenience we restate Theorem2 in the following. Theorem 3 (Restatement of Theorem 2). Let |Ψi be a state chosen from an ensemble on H that forms an ε0- 0 approximate k -design. Then the projected ensemble EA,Ψ forms an ε-approximate k-design with probability at least 1 − δ if

1  NB = Ω kNA + log εδ , (C50) 0 1  k = Ω k NB + log εδ , (C51) 1 1  log( ε0 ) = Ω kNB NB + log εδ , (C52) 1  NA = Ω log(NB) + log(k) + log log εδ . (C53) In the following subsections, we will begin with a discussion of a generalized Levy’s lemma that provides sharper concentration for quadratic functions. We will then give a general structure of the proof and provide the detailed proof of several technical lemmas afterwards.

a. Higher order concentration

In the proof of Theorem1, Levy’s lemma plays a crucial role in establishing the desired statement. Levy’s lemma tells us about a sharp concentration when the random variables are sampled from a high-dimensional sphere. We will continue to utilize concentration on high-dimensional spheres for the proof of Theorem2. However, the original statement of Levy’s lemma does not provide good bounds for Theorem2. We will instead make use of a higher order variant of Levy’s lemma. In this section we recall a higher order variant of Levy’s lemma that was established in [33], and use it to provide a simple proof of a concentration inequality for the quantity hΦez|Φezi. This higher order concentration results provided in [33] is a useful tool for obtaining concentration inequalities for polynomial functions on measures satisfying a log-Sobolev inequality, and may prove useful in other problems arising in quantum information theory. For this work we only need the following result from [33]:

Theorem 4 (Theorem1.13 in [33]) . Let f be a C2-smooth function on an open neighborhood of the sphere SN−1 with R N−1 f dσN−1 = 0. If S Z 2 1 k∇fk2 dσN−1 ≤ (C54) N−1 N S N−1 and k Hess f(θ)k∞ ≤ 1 for all θ ∈ S then Z exp((N − 1)|f|/(8e)) dσN−1 ≤ 2. (C55) N−1 S As a corollary, we obtain the following result for quadratic forms.

N N 1 Corollary 1 (Concentration for quadratic forms). Let Q : R → R be a linear map satisfying kQk2 ≤ 2 and 1 kQk∞ ≤ 2 , and let Z T q0 = θ Qθ dσN−1(θ). N−1 S Then Z T exp((N − 1)(θ Qθ − q0)/(8e)) dσN−1 ≤ 2. (C56) N−1 S T Proof. Let f(x) = x Qx − a0. Then ∇f(x) = 2Qx, so Z Z 2 T T k∇fk2 dσN−1 = 4 θ Q Qθ dσN−1(θ). (C57) N−1 N−1 S S 19

N T 2 Let {uj}j=1 be an orthonormal basis of eigenvectors of Q Q with eigenvalues λj . Then the latter integral is equal to Z X 2 2 4 X 2 −1 2 4 λj (θ · uj) dσN−1(θ) = λj = 4n kQk2. (C58) N−1 N j S j

If kQk2 ≤ 1/2 then f satisfies the condition (C54). Moreover Hess f = 2Q, so k Hess fk∞ = 2kMk∞. We can therefore apply Theorem4 to f to obtain (C56). We only need to use Corollary (1) in the case that Q is an orthogonal projection.

Corollary 2 (Concentration for quadratic forms with projectors). Let V ⊂ RN be a subspace of RN with dimension m, and let PV be the orthogonal projection onto V . Then

 (N − 1)|kP θk2 − m/N| exp √V 2 ≤ 2 ,, (C59) EN−1 θ∼S 8e 2m and in particular, for θ sampled uniformly from the sphere,    2  1 −1/2 Probθ∼ N−1 |kPV θk − m/N| ≥ δ ≤ 2 exp − √ (N − 1)m δ . (C60) S 8e 2

2 −1/2 Proof. Observe that kPV k2 = m and kPV k∞ = 1, so the operator (2m) PV satisfies the conditions of Corollary1, from which (C59) follows. Then (C60) follows from Markov’s inequality applied to (C59).

Specializing to the case that |Φi is a Haar-random quantum state on H and PV is the projector Pz = 1A ⊗ |xi hx|, we deduce an exponential concentration inequality for hΦez|Φezi. Corollary 3 (Concentration for probability in a projected ensemble). We have the bound     1 1 1/2 ProbΦ∼Haar(d) hΦez|Φezi − ≥ δ ≤ 2 exp − √ dA dB δ . (C61) dB 8e 2

b. General structure of the proof

The proof is based on the idea of modulating the target expression, which is a rational function, with a high-degree polynomial to form a polynomial function. More precisely, we will be interested in the following two functions:

⊗k X |ΦezihΦez| A(|Φi) = , (C62) hΦ |Φ ik−1 z∈{0,1}NB ez ez  b ⊗k  2(k−1)! X |ΦezihΦez| dB B(|Φi) = 1 − 1 − hΦez|Φezi  , (C63) hΦ |Φ ik−1 r z∈{0,1}NB ez ez where r, b > 0 are parameters for tuning the approximation of polynomial function B(|Φi) to the target expression A(|Φi), which is a rational function. Because the 1 is cancelled in the binomial expansion

 2(k−1)!b b    2(k−1)p dB X b p dB 1 − 1 − hΦez|Φezi = (−1) hΦez|Φezi , (C64) r p r p=1 it is not hard to see that B(|Φi) is indeed a polynomial function in the real and imaginary parts of |Φi. We will consider |Φi to be sampled from an (ε0, k0)-design. We will use the basic Markov inequality to bound the concentration. A higher order concentration inequality can also be used but would require k0 to be higher. The central quantity in Markov inequality is the expectation value of the error

E A(|Φi) − E A(|Ψi) . (C65) Φ∼(ε0,k0) design Ψ∼Haar 1 20

We will use B(|Φi) as a surrogate to obtain an upper bound on the above quantity. This is because B(|Φi) is a polynomial function rather than a rational function in the real and imaginary parts of |Φi. A triangle inequality gives the following bound

E A(|Φi) − E A(|Ψi) (C66) Φ∼(ε0,k0) design Ψ∼Haar 1

≤ E kA(|Φi) − B(|Φi)k1 + E B(|Φi) − E A(|Ψi) . (C67) (ε0,k0) design (ε0,k0) design Ψ∼Haar 1 We can now analyze the two terms independently by using properties of quantum designs. For the first term, we will prove in Lemma5 that the following inequality holds.

 2(k−1)!b X dB kA(|Φi) − B(|Φi)k ≤ hΦez|Φezi 1 − hΦez|Φezi . (C68) 1 r z∈{0,1}NB

Therefore we will define a polynomial function

 2(k−1)!b X dB R(|Φi) = hΦez|Φezi 1 − hΦez|Φezi , (C69) r z∈{0,1}NB which is an upper bound on the approximation error between A(|Φi) and B(|Φi). Because a quantum design approx- imates any polynomial function, we have the following upper bound

E kA(|Φi) − B(|Φi)k1 ≤ E R(|Φi) (C70) (ε0,k0) design (ε0,k0) design 0 ≤ E R(|Φi) + ε × E1, (C71) Haar where the exact expression of E1 is given in Lemma6. We will also apply a similar philosophy for bounding the second term in Eq. (C67) by first upper bounding the term by a polynomial function that turns expectations over designs into expectations over the Haar measure: q k E B(|Φi) − E A(|Ψi) ≤ E dA B(|Φi) − E A(|Ψi) (C72) (ε0,k0) design Ψ∼Haar 1 (ε0,k0) design Ψ∼Haar 2 s 2 k ≤ E dA B(|Φi) − E A(|Ψi) (C73) (ε0,k0) design Ψ∼Haar 2 r 2 k 0 ≤ E dA B(|Φi) − E A(|Ψi) + ε × E2. (C74) Haar Ψ∼Haar 2 The first line follows from the relation between 1-norm and 2-norm. The second line applies Jensen’s inequality. The k 2 third line follows from the observation that dA kB(|Φi) − EΨ∼Haar A(|Ψi)k2 is a polynomial function in |Φi. Therefore, we can turn the expectation over a quantum state design to one over the Haar measure, which incurs a small error of 0 2 ε ×E2. The exact expression of E2 is given in Lemma7. We will now upper bound EHaar kB(|Φi) − EΨ∼Haar A(|Ψi)k2 by turning B(|Φi) back into A(|Φi) and incurring an additional error:

2 2 2 E B(|Φi) − E A(|Ψi) ≤ 2 E kB(|Φi) − A(|Φi)k2 + 2 E A(|Φi) − E A(|Ψi) (C75) Haar Ψ∼Haar 2 Haar Haar Ψ∼Haar 2 2 2 ≤ 2 E kB(|Φi) − A(|Φi)k1 + 2 E A(|Φi) − E A(|Ψi) (C76) Haar Haar Ψ∼Haar 2 2 2 ≤ 2 E R(|Φi) + 2 E A(|Φi) − E A(|Ψi) . (C77) Haar Haar Ψ∼Haar 2

The first line follows from the triangle inequality and (a + b)2 ≤ 2(a2 + b2), ∀a, b ∈ R. The second line follows from the relation between the 1-norm and 2-norm. The third line uses the inequality given in Eq. (C68), which is proved in Lemma5. We can now combine Eq.’s (C67), (C71), (C74), and (C77) to find

E A(|Φi) − E A(|Ψi) (C78) Φ∼(ε0,k0) design Ψ∼Haar 1 21

r 2 0 2 0 ≤ E R(|Φi) + ε × E1 + 2 E R(|Φi) + 2 E A(|Φi) − E A(|Ψi) + ε × E2 . (C79) Haar Haar Haar Ψ∼Haar 2 2 In Lemma8, we give upper bounds for EHaar R(|Φi) and EHaar R(|Φi) . Lemma8 is the key to Theorem2. The goodness of fit of the polynomial approximation B(|Φi) to the target expression A(|Φi) is reflected in upper bound 2 of EHaar R(|Φi) and EHaar R(|Φi) . By changing the polynomial approximation, one may likely obtain an improved statement of Theorem2. We leave open the choice of the optimal polynomial approximation, and focus mainly on the simple and tractable polynomial approximation given in Eq. (C63). The upper bounds for EHaar R(|Φi) and 2 EHaar R(|Φi) rely on the higher order variant of Levy’s lemma given in AppendixC4a. In Lemma9, we obtain an 2 0 0 upper bound on EHaar kA(|Φi) − EHaar A(|Φi)k2. Along with Lemma6 that bounds ε E1, Lemma7 that bounds ε E2, we can show that for any ε > 0 as long as we choose r = 2 and b to be an even integer with

b = Ω (NB + log(1/ε)) , (C80)

NA = Ω (log(NB) + log(k) + log log(1/ε)) , (C81) k0 = Ω (bk) , (C82) 0 log(1/ε ) = Ω (log(1/ε) + k(bNB + log k + NA)) , (C83)  1 N = Ω kN + log , (C84) B A ε we can obtain the following upper bounds

0 0 2 ε × E1 ≤ ε, ε × E2 ≤ ε , (C85) 2 2 E R(|Φi) ≤ ε, E R(|Φi) ≤ ε , (C86) Haar Haar 2 2 E A(|Φi) − E A(|Ψi) ≤ ε . (C87) Haar Ψ∼Haar 2 Together with Markov’s inequality, we have the following concentration result: √   2ε + 5ε2 ε Pr A(|Φi) − E A(|Ψi) ≥ ε˜ ≤ = 5 (C88) Φ∼(ε0,k0) design Ψ∼Haar 1 ε˜ ε˜ for any ε, ε˜ > 0. Using Lemma3 established in the proof of Theorem1, we have  ⊗k E A(|Ψi) = E (|ΨihΨ|) . (C89) Ψ∼Haar Ψ∼Haar(dA) Therefore the projected ensemble of a randomly sample state |Φi forms an (ε, k)-design on the local subsystem with NA dimension dA = 2 under probability at least 1 − δ as long as the following conditions hold:   1  N = Ω kN + log , (C90) B A εδ    1  k0 = Ω k N + log , (C91) B εδ    1  log(1/ε0) = Ω kN N + log , (C92) B B εδ   1  N = Ω log(N ) + log(k) + log log . (C93) A B εδ This concludes the proof of Theorem2.

c. Technical lemmas

Lemma 5 (Error bound on polynomial approximation). If b is even, then kA(|Φi) − B(|Φi)k1 ≤ R(|Φi). Proof. Recall the following definitions:

⊗k X |ΦezihΦez| A(|Φi) = , (C94) hΦ |Φ ik−1 z∈{0,1}NB ez ez 22

 b ⊗k  2(k−1)! X |ΦezihΦez| dB B(|Φi) = 1 − 1 − hΦez|Φezi  , (C95) hΦ |Φ ik−1 r z∈{0,1}NB ez ez  2(k−1)!b X dB R(|Φi) = hΦez|Φezi 1 − hΦez|Φezi . (C96) r z∈{0,1}NB

We note the definition of trace norm k·k1 for Hermitian matrices:

kXk1 = sup tr(OX). (C97) O:kOk∞≤1 Hence, we have

 b ⊗k  2(k−1)! X |ΦezihΦez| dB kA(|Φi) − B(|Φi)k1 = sup tr O 1 − hΦez|Φezi  (C98) O:kOk ≤1 hΦ |Φ ik−1 r ∞ z∈{0,1}NB ez ez     ⊗k ⊗k !b hΦz| O |Φzi  2(k−1) X e e dB = sup 1 − hΦez|Φezi (C99) O:kOk ≤1 hΦ |Φ ik−1 r ∞ z∈{0,1}NB ez ez b t  2(k−1)! X hΦez|Φezi dB ≤ sup 1 − hΦez|Φezi (C100) k−1 O:kOk ≤1 hΦz|Φzi r ∞ z∈{0,1}NB e e  2(k−1)!b X dB ≤ sup hΦez|Φezi 1 − hΦez|Φezi (C101) O:kOk ≤1 r ∞ z∈{0,1}NB = R(|Φi). (C102)

This concludes the proof. Lemma 6 (Quantum state design on the error bound). For k0 ≥ 2b(k − 1) + 1, we have

0 E R(|Φi) ≤ E R(|Φi) + ε E1, (C103) (ε0,k0) design Haar where the error term is given by !b d 2(k−1) E = 1 + B . (C104) 1 r

NB In particular, if we choose r = 2 ≤ dB and recall that dB = 2 , then for any ε > 0 as long as k0 = Ω (bk) , (C105) 0 log(1/ε ) = Ω (log(1/ε) + bkNB) , (C106)

0 we have the following upper bound on the error term: ε E1 ≤ ε. 0 Proof. Letting Mzz = 1A ⊗ |zihz|, for k ≥ ` we have

X ` X ` E hΦez|Φezi − E hΦez|Φezi (ε0,k0) design Haar z∈{0,1}NB z∈{0,1}NB      1⊗(k0−`) X ⊗` ⊗k0 ⊗k0 = tr  ⊗ Mzz  · E (|ΦihΦ|) − E (|ΦihΦ|)  (ε0,k0) design Haar z∈{0,1}NB

1⊗(k0−`) X ⊗` ⊗k0 ⊗k0 ≤ ⊗ Mzz E (|ΦihΦ|) − E (|ΦihΦ|) (ε0,k0) design Haar N 1 z∈{0,1} B ∞ 23

≤ ε0 (C107)

0 1⊗(k −`) P ⊗` where in the last inequality we have used ⊗ z∈{0,1}NB Mzz = 1. Upon examining Eq. (C69), we see ∞ that for k0 ≥ 2b(k − 1) + 1 we can leverage (C107) to achieve

 2(k−1)!b 0 dB E R(|Φi) ≤ E R(|Φi) + ε 1 + , (C108) (ε0,k0) design Haar r which is the desired bound. Lemma 7 (Quantum state design on the polynomial approximation). For k0 ≥ 4b(k − 1) + 2,

2 2 k k 0 E dA B(|Φi) − E [A(|Ψi)] ≤ E dA B(|Φi) − E [A(|Ψi)] + ε E2, (ε0,k0) design Ψ∼Haar 2 Haar Ψ∼Haar 2 where the error term is given by    !b  !b2 d 2(k−1) d 2(k−1) E = 2 k! −1 + 1 + B + dk −1 + 1 + B  . (C109) 2   r  A  r  

NB In particular, if we choose r = 2 ≤ dB and recall that dB = 2 , then for any ε > 0 as long as

k0 = Ω (bk) , (C110) 0 log(1/ε ) = Ω (log(1/ε) + k(bNB + log k + NA)) , (C111)

0 2 we have the following upper bound on the error term: ε E2 ≤ ε . Proof. Recall from Lemma3 that

⊗k ΠA,k E [A(|Ψi)] = E (|ΦihΦ|) = (C112) Ψ∼Haar(d) Φ∼Haar(d ) dA+k−1 A k

⊗k where in the last line, ΠA,k is the projector onto the symmetric subspace of HA . Then we can write 2 kB(Φ) − EΨ∼Haar[A(|Ψi)]k2 as 2 2 tr(B(Φ)) 1 2 B(Φ) − E [A(|Ψi)] = tr(B(Φ) ) − d +k−1 + d +k−1 . (C113) Ψ∼Haar 2 A  A  k k Using (C107), we have for k0 ≥ 2b(k − 1) + 1 that   ! !  2(k−1)!b k 2 tr(B(Φ)) k 2 tr(B(Φ)) 0 dB dA E − ≤ dA E − + ε · 2 k! −1 + 1 +  . (C114) (ε0,k0) design dA+k−1 Haar dA+k−1 r k k

k 2 To obtain a similar bound on dA E(ε0,k0) design tr(B(Φ) ), we need the following generalization of the inequality 0 in (C107). Letting Myz = 1A ⊗ |yihz|, for k ≥ 2(p + q) we have

X 2p q q X 2p q q E |hΦez|Φeyi| hΦez|Φezi hΦey|Φeyi − E |hΦez|Φeyi| hΦez|Φezi hΦey|Φeyi (ε0,t0) design Haar y,z∈{0,1}NB y,z∈{0,1}NB      1⊗(k0−2(p+q)) X ⊗p ⊗p ⊗q ⊗q ⊗k0 ⊗k0 = tr  ⊗ Myz ⊗ Myz ⊗ Mzz ⊗ Myy  · E (|ΦihΦ|) − E (|ΦihΦ|)  (ε0,k0) design Haar y,z∈{0,1}NB

1⊗(k0−2(p+q)) X ⊗p ⊗p ⊗q ⊗q ⊗k0 ⊗k0 ≤ ⊗ Myz ⊗ Myz ⊗ Mzz ⊗ Myy E (|ΦihΦ|) − E (|ΦihΦ|) (ε0,k0) design Haar N 1 x,y∈{0,1} B ∞ 24

≤ ε0 (C115)

0 1⊗(k −2(p+q)) P ⊗p ⊗p ⊗q ⊗q 0 where we have used that ⊗ y,z∈{0,1}NB Myz ⊗ Myz ⊗ Mzz ⊗ Myy = 1. Using (C115), for k ≥ ∞ 2t + 2(2b − 1)(k − 1) = 4b(k − 1) + 2 we have  2  2(k−1)!b k 2 k 2 0 k dB dA E tr(B(Φ) ) ≤ dA E tr(B(Φ) ) + ε · dA −1 + 1 +  . (C116) (ε0,k0) design Haar r

Putting together (C114) and (C116), we arrive at 2 2 k k E dA B(|Φi) − E [A(|Ψi)] ≤ E dA B(|Φi) − E [A(|Ψi)] (ε0,k0) design Ψ∼Haar 2 Haar Ψ∼Haar 2    !b  !b2 d 2(k−1) d 2(k−1) + ε0 2 k! −1 + 1 + B + dk −1 + 1 + B  (C117)   r  A  r   for k0 ≥ 4b(k − 1) + 2 as claimed.

1/4 Lemma 8 (Moment bounds on error function). For t ≥ 2, r = 1, b even, and dA ≥ 8t − 6,   dB 2b(k−1)+2 1 1/4 E R(|Φi) ≤ + 2dB exp − √ dA , (C118) Haar 2b 8e 2 2   2 dB 4b(k−1)+3 1 1/4 E R(|Φi) ≤ + 2dB exp − √ dA . (C119) Haar 22b 8e 2

NA NB In particular, if we fix an ε > 0 and recall that dA = 2 , dB = 2 , then as long as

b = Ω (NB + log(1/ε)) , (C120)

NA = Ω (log(NB) + log(k) + log log(1/ε)) , (C121) we have the following moment bounds on the error function

E R(|Φi) ≤ ε, (C122) Haar 2 2 E R(|Φi) ≤ ε . (C123) Haar Proof. Recall the following definitions with r = 1:

b X   2(k−1) R(|Φi) = hΦez|Φezi 1 − dBhΦez|Φezi . (C124) z∈{0,1}NB

We note that hΦez|Φezi ∈ [0, 1] and we define an associated function b  2(k−1) γ(s) = s 1 − (dBs) , ∀s ∈ [0, 1]. (C125)

1/4 1/4 For s ∈ [(1 − 1/dA )/dB, (1 + 1/dA )/dB], we can see that γ(s) ≥ 0 because b is even. We now proceed to upper 1/4 bound γ(s) in this domain. We have the following bound based on the condition that s ≤ (1 + 1/dA )/dB, !2(k−1) 1 2(k − 1)/d1/4 4(k − 1) (d s)2(k−1) − 1 ≤ 1 + − 1 ≤ A ≤ . (C126) B 1/4 1/4 1/4 dA 1 − (2t − 3)/dA dA 2(k−1) The first inequality follows from the monotonicity of (dBs) − 1. The second inequality follows from the fact that (1 + x)n ≤ 1 + nx/(1 − (n − 1)x), ∀z ∈ [−1, 1/(n − 1)], n > 1. The third inequality uses the condition that 1/4 dA ≥ 8t − 6 ≥ 2(2t − 3). We can also obtain another bound using s ≥ (1 − 1/dA )/dB, 2(k − 1) 1 − (d s)2(k−1) ≤ , (C127) B 1/4 dA 25

n 1/4 which follows from the fact that (1 + x) ≥ 1 + nx, ∀x ≥ −1, n ≥ 1. Together we have ∀s ∈ [(1 − 1/dA )/dB, (1 + 1/4 1/dA )/dB], the function γ(s) is bounded as follows: !b b 4(k − 1) 1 |γ(s)| ≤ (d s)2(k−1) − 1 ≤ ≤ . (C128) B 1/4 2b dA

1/4 The second inequality uses the fact that dA ≥ 8t − 6 ≥ 8(k − 1). If the variable s is not within that domain, but is in [0, 1], we have

2b(k−1) |γ(s)| ≤ dB . (C129) We consider the event G such that " # 1 − 1/d1/4 1 + 1/d1/4 A A NB hΦez|Φezi ∈ , , ∀z ∈ {0, 1} , (C130) dB dB which is equivalent to the event that

1 1 hΦ |Φ i − ≤ , ∀z ∈ {0, 1}NB . (C131) ez ez 1/4 dB dA dB We now utilize the concentration result given in Corollary3,     1 1 1/2 ProbΦ∼Haar(d) hΦez|Φezi − ≥ δ ≤ 2 exp − √ dA dB δ , (C132) dB 8e 2 to derive the probability upper bound for the complement of the event G   1 1/4 ProbΦ∼Haar(d)[ G did not happen ] ≤ 2dB exp − √ d , (C133) 8e 2 A which is obtained by taking the union bound. We can now proceed to upper bound the first and second moments of P R(|Φi) by noting that R(|Φi) = z∈{0,1}NB γ(hΦez|Φezi). When the event G happens, we can use Eq. (C128) to obtain

d |R(|Φi)| ≤ B . (C134) 2b If the event G did not happen, then using Eq. (C129), we have

2b(k−1)+1 |R(|Φi)| ≤ dB . (C135) Together for both m = 1, 2, we have  m m dB E R(|Φi) ≤ ProbΦ∼Haar(d)[ G happened ] (C136) Haar 2b m(2b(k−1)+1) + dB ProbΦ∼Haar(d)[ G did not happen ] (C137) dm  1  ≤ B + 2dm(2b(k−1)+1)+1 exp − √ d1/4 , (C138) 2bm B 8e 2 A where the last inequality uses the concentration result given in Eq. (C133). The asymptotic bounds can be obtained from the above result under suitable choices of the parameters given in the statement of this lemma. Lemma 9 (Second moment bound for projected ensemble from Haar measure).

" 2# 3 6k E A(|Φi) − E [A(|Ψi)] ≤ 36π (2k − 1)(dA /dB) . (C139) Φ∼Haar(d) Ψ∼Haar(d) 2 26

NA NB In particular, recalling that dA = 2 , dB = 2 , then for any ε > 0, as long as  1 N = Ω kN + log , (C140) B A ε we have the upper bound

" 2# 2 E A(|Φi) − E [A(|Ψi)] ≤ ε . (C141) Φ∼Haar(d) Ψ∼Haar(d) 2

Proof. Notice that Eq. (C47) implies " # 2  d y  Prob A(|Φi) − [A(|Ψi)] ≥ y ≤ 2d2k exp − B . (C142) Φ∼Haar(d) E A 3 4k Ψ∼Haar(d) 2 18π (2k − 1)dA

R ∞ Then we can use the fact that E[X] = 0 dx Prob[X ≥ x] for any positive X to obtain

" 2#

E A(|Φi) − E [A(|Ψi)] (C143) Φ∼Haar(d) Ψ∼Haar(d) 2 Z ∞ " 2 #

= dy ProbΦ∼Haar(d) A(|Φi) − E [A(|Ψi)] ≥ y (C144) −∞ Ψ∼Haar(d) 2 Z ∞  d y  ≤ dy 2d2k exp − B (C145) A 3 4k 0 18π (2k − 1)dA d6k ≤ 36π3(2k − 1) A , (C146) dB which establishes the bound.