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

Random product states at high temperature equilibrate exponentially well

Yichen Huang (黄溢辰) yichenhuang@fas.harvard.edu Department of Physics, Harvard University, Cambridge, Massachusetts 02138, USA 0000-0002-8496-9251
Abstract

We prove that for all but a measure zero set of local Hamiltonians, starting from random product states at sufficiently high but finite temperature, with overwhelming probability expectation values of observables equilibrate such that at sufficiently long times, fluctuations around the stationary value are exponentially small in the system size.

1 Introduction

Equilibration is the process where the value of a dynamic property becomes almost time independent in the sense of staying very close to a particular value [1]. If the stationary value matches the thermal value (property of the Gibbs state with the same energy), then thermalization occurs. Thus, equilibration is a prerequisite for thermalization. A thermalizing system equilibrates by definition, but not vice versa. The most famous example of “equilibration without thermalization” is arguably many-body localization.

Not all quantum many-body systems equilibrate. A simple counterexample is a non-interacting multi-spin system whose Hamiltonian is a sum of on-site terms. If the initial state is a product state, each spin evolves (rotates) independently and thus never reaches equilibrium. However and of course, we expect that almost all quantum many-body systems (with local interactions) equilibrate. This can be rigorously established from the laws of quantum mechanics. It follows from a line of research over a long period of time [2, 3, 4, 5, 6, 7, 8, 9] that

Theorem 1 (informal).

For all but a measure zero set of local Hamiltonians, starting from a random product state, with overwhelming probability expectation values of observables equilibrate such that at sufficiently long times, fluctuations around the stationary value are exponentially small in the system size.

Remark.

This theorem has been extended to Floquet systems [10].

In the thermodynamic limit, the energy density of a random product state approaches the mean energy density of the Hamiltonian (average of all eigenvalues divided by the system size) with overwhelming probability. Thus, Theorem 1 is a statement at infinite temperature.

The main contribution of this paper is to extend Theorem 1 to sufficiently high but finite temperature. Random product states at finite temperature can be defined by either

  • a uniform distribution over a subset of product states whose energy corresponds to a given temperature (Definition 4) or

  • properly defining the probability of each product state so that their mixture is a thermal state, as done in Ref. [11] (Definition 5).

We prove analogues of Theorem 1 for both definitions above.

2 Preliminaries

We use standard asymptotic notation. Let f,g:++f,g:\mathbb{R}^{+}\to\mathbb{R}^{+} be two functions. One writes f(x)=O(g(x))f(x)=O(g(x)) if and only if there exist constants C,x0>0C,x_{0}>0 such that f(x)Cg(x)f(x)\leq Cg(x) for all x>x0x>x_{0}; f(x)=Ω(g(x))f(x)=\Omega(g(x)) if and only if there exist constants C,x0>0C,x_{0}>0 such that f(x)Cg(x)f(x)\geq Cg(x) for all x>x0x>x_{0}; f(x)=Θ(g(x))f(x)=\Theta(g(x)) if and only if there exist constants C1,C2,x0>0C_{1},C_{2},x_{0}>0 such that C1g(x)f(x)C2g(x)C_{1}g(x)\leq f(x)\leq C_{2}g(x) for all x>x0x>x_{0}; f(x)=o(g(x))f(x)=o(g(x)) if and only if for any constant C>0C>0 there exists a constant x0>0x_{0}>0 such that f(x)<Cg(x)f(x)<Cg(x) for all x>x0x>x_{0}.

Definition 1 (non-degenerate gap).

The spectrum {ej}\{e_{j}\} of a Hamiltonian has non-degenerate gaps if the differences {ejek}jk\{e_{j}-e_{k}\}_{j\neq k} are all distinct, i.e., for any jkj\neq k,

ejek=ejek(j=j)and(k=k).e_{j}-e_{k}=e_{j^{\prime}}-e_{k^{\prime}}\implies(j=j^{\prime})~{}\textnormal{and}~{}(k=k^{\prime}). (1)

Consider a system of NN qubits (spin-1/21/2’s) initialized in the state |ψ|\psi\rangle. Let {|j}j=12N\{|j\rangle\}_{j=1}^{2^{N}} be a complete set of eigenstates of the Hamiltonian HH.

Definition 2 (effective dimension).

The effective dimension of |ψ|\psi\rangle is defined as

1Dψeff=j=12N|j|ψ|4.\frac{1}{D^{\textnormal{eff}}_{\psi}}=\sum_{j=1}^{2^{N}}\big{|}\langle j|\psi\rangle\big{|}^{4}. (2)

The time-averaged expectation value and fluctuation of an operator BB are

B¯:=limT1T0Tψ|B(t)|ψdt,B(t):=eiHtBeiHt,\displaystyle\bar{B}:=\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}\langle\psi|B(t)|\psi\rangle\,\mathrm{d}t,\quad B(t):=e^{iHt}Be^{-iHt}, (3)
ΔB:=limT1T0T|ψ|B(t)|ψB¯|2dt.\displaystyle\Delta B:=\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}\big{|}\langle\psi|B(t)|\psi\rangle-\bar{B}\big{|}^{2}\,\mathrm{d}t. (4)

Let AA be a subsystem of LL qubits and A¯\bar{A} be the rest of the system. Let

ψA(t)=trA¯(eiHt|ψψ|eiHt),ψ¯A=limT1T0TψA(t)dt\psi_{A}(t)=\operatorname{tr}_{\bar{A}}(e^{-iHt}|\psi\rangle\langle\psi|e^{iHt}),\quad\bar{\psi}_{A}=\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}\psi_{A}(t)\,\mathrm{d}t (5)

be the reduced density matrix of subsystem AA at time tt and its time average.

Theorem 2 ([3, 4, 5]).

If the spectrum of HH has non-degenerate gaps, then

ΔBB2/Dψeff,limT1T0TψA(t)ψ¯A1dt2L/Dψeff.\Delta B\leq\|B\|^{2}/D^{\textnormal{eff}}_{\psi},\quad\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}\|\psi_{A}(t)-\bar{\psi}_{A}\|_{1}\,\mathrm{d}t\leq 2^{L}/\sqrt{D^{\textnormal{eff}}_{\psi}}. (6)
Theorem 3 ([9]).

In the space of local Hamiltonians, the non-degenerate gap condition (1) is satisfied almost everywhere.

Let |x±,|y±,|z±|x^{\pm}\rangle,|y^{\pm}\rangle,|z^{\pm}\rangle be the eigenvectors of the Pauli matrices

σx=(0110),σy=(0ii0),σz=(1001)\sigma^{x}=\begin{pmatrix}0&1\\ 1&0\end{pmatrix},\quad\sigma^{y}=\begin{pmatrix}0&-i\\ i&0\end{pmatrix},\quad\sigma^{z}=\begin{pmatrix}1&0\\ 0&-1\end{pmatrix} (7)

with eigenvalues ±1\pm 1, respectively.

Definition 3.

Let \mathcal{E} be the uniform distribution over the set

S:={|x±,|y±,|z±}N.S:=\{|x^{\pm}\rangle,|y^{\pm}\rangle,|z^{\pm}\rangle\}^{\otimes N}. (8)
Theorem 4.
𝔼|ψ1Dψeff(2/3)N,Pr|ψ(Dψeff=eΩ(N))=1eΩ(N).\operatorname*{\mathbb{E}}_{|\psi\rangle\sim\mathcal{E}}\frac{1}{D^{\textnormal{eff}}_{\psi}}\leq(2/3)^{N},\quad\Pr_{|\psi\rangle\sim\mathcal{E}}(D^{\textnormal{eff}}_{\psi}=e^{\Omega(N)})=1-e^{-\Omega(N)}. (9)

This theorem can be proved in the same way as Lemma 5 in Ref. [8]. It is a special case of Lemma 4, whose full proof is given below.

Theorem 1 follows from Theorems 2, 3, 4.

3 Results

To extend Theorem 1 to finte temperature, it suffices to prove an analogue of Theorem 4 for random product states at finite temperature.

Consider a constant-dimensional hypercubic lattice of NN sites, where each lattice site has a qubit. We expand the Hamiltonian

H=P𝒫NcPPH=\sum_{P\in\mathcal{P}_{N}}c_{P}P (10)

in the Pauli basis, where 𝒫N\mathcal{P}_{N} is the set of Pauli operators on NN qubits. Assume without loss of generality that cI=0c_{I}=0 so that trH=0\operatorname{tr}H=0. The support suppP\operatorname{supp}P of a Pauli operator is the set of qubits that PP acts non-trivially on. Suppose that HH is local and extensive in that

  • cP=0c_{P}=0 if the diameter of suppP\operatorname{supp}P is greater than a certain constant rr.

  • For any hypercubic region RR of side length rr, there is at least one term in HH with cP0c_{P}\neq 0 such that suppPR\operatorname{supp}P\subseteq R.

  • |cP|=Θ(1)|c_{P}|=\Theta(1) if cP0c_{P}\neq 0.

Let

ρβ:=eβH/tr(eβH),E(β):=tr(ρβH).\rho_{\beta}:=e^{-\beta H}/\operatorname{tr}(e^{-\beta H}),\quad E(\beta):=\operatorname{tr}(\rho_{\beta}H). (11)

be a thermal state and its energy at inverse temperature β\beta.

Definition 4.

Let βmc\mathcal{E}^{\textnormal{mc}}_{\beta} be the uniform distribution over the set

Sβ:={|ψS:|ψ|H|ψE(β)|=o(N)}.S_{\beta}:=\big{\{}|\psi\rangle\in S:\big{|}\langle\psi|H|\psi\rangle-E(\beta)\big{|}=o(N)\big{\}}. (12)
Theorem 5.

There is a constant C>0C>0 such that for any |β|C|\beta|\leq C,

𝔼|ψβmc1Dψeff=eΩ(N),Pr|ψβmc(Dψeff=eΩ(N))=1eΩ(N).\operatorname*{\mathbb{E}}_{|\psi\rangle\sim\mathcal{E}^{\textnormal{mc}}_{\beta}}\frac{1}{D^{\textnormal{eff}}_{\psi}}=e^{-\Omega(N)},\quad\Pr_{|\psi\rangle\sim\mathcal{E}^{\textnormal{mc}}_{\beta}}(D^{\textnormal{eff}}_{\psi}=e^{\Omega(N)})=1-e^{-\Omega(N)}. (13)

Let I2I_{2} be the identity matrix of order 22 and

S+:={|x±,|y±,|z±,I2/2}N.S^{+}:=\{|x^{\pm}\rangle,|y^{\pm}\rangle,|z^{\pm}\rangle,I_{2}/2\}^{\otimes N}. (14)
Theorem 6 (Theorem 1.5 in Ref. [11]).

There is a constant βc>0\beta_{c}>0 such that for any |β|βc|\beta|\leq\beta_{c}, ρβ\rho_{\beta} is separable:

ρβ=ρS+p(ρ)ρ,\rho_{\beta}=\sum_{\rho\in S^{+}}p(\rho)\rho, (15)

where p:S+[0,1]p:S^{+}\to[0,1] is a probability distribution over S+S^{+}.

Corollary 1 (Theorem 1.1 in Ref. [11]).

For any |β|βc|\beta|\leq\beta_{c}, ρβ\rho_{\beta} can be written as a distribution over product states in SS.

This corollary can be regarded as a definition of random product states at sufficiently high temperature [12].

We now explicitly derive Corollary 1 from Theorem 6. Let n(ρ)n(\rho) be the number of “I2/2I_{2}/2” in ρS+\rho\in S^{+} so that ψ|ρ|ψ2n(ρ)\langle\psi|\rho|\psi\rangle\leq 2^{-n(\rho)} for any |ψS|\psi\rangle\in S. Let

Sρ:={|ψS:ψ|ρ|ψ=2n(ρ)},S_{\rho}:=\{|\psi\rangle\in S:\langle\psi|\rho|\psi\rangle=2^{-n(\rho)}\}, (16)

i.e., |ψSρ|\psi\rangle\in S_{\rho} if and only if |ψ|\psi\rangle and ρ\rho are the same on qubits where ρ\rho is not I2/2I_{2}/2. Thus,

|Sρ|=6n(ρ),ρ=1|Sρ||ψSρ|ψψ|.|S_{\rho}|=6^{n(\rho)},\quad\rho=\frac{1}{|S_{\rho}|}\sum_{|\psi\rangle\in S_{\rho}}|\psi\rangle\langle\psi|. (17)
Definition 5.

The proof of Theorem 6 is constructive. Let pp be the probability distribution constructed in Ref. [11]. We sample and obtain ρS+\rho\in S^{+} with probability p(ρ)p(\rho). Then, we sample uniformly from SρS_{\rho}. This results in a probability distribution βcan\mathcal{E}^{\textnormal{can}}_{\beta} over SS so that

ρβ=𝔼|ψβcan|ψψ|.\rho_{\beta}=\operatorname*{\mathbb{E}}_{|\psi\rangle\sim\mathcal{E}^{\textnormal{can}}_{\beta}}|\psi\rangle\langle\psi|. (18)
Remark.

βcan\mathcal{E}^{\textnormal{can}}_{\beta} is defined only for |β|βc|\beta|\leq\beta_{c} where the construction of pp in Ref. [11] is provably valid.

Theorem 7.
𝔼|ψβcan1Dψeff=eΩ(N),Pr|ψβcan(Dψeff=eΩ(N))=1eΩ(N).\operatorname*{\mathbb{E}}_{|\psi\rangle\sim\mathcal{E}^{\textnormal{can}}_{\beta}}\frac{1}{D^{\textnormal{eff}}_{\psi}}=e^{-\Omega(N)},\quad\Pr_{|\psi\rangle\sim\mathcal{E}^{\textnormal{can}}_{\beta}}(D^{\textnormal{eff}}_{\psi}=e^{\Omega(N)})=1-e^{-\Omega(N)}. (19)

4 Proofs

4.1 Proof of Theorem 5

Lemma 1.
min|ψSψ|Hψ=Ω(N).\min_{|\psi\rangle\in S}\langle\psi|H|\psi\rangle=-\Omega(N). (20)
Proof.

We construct hypercubic regions R1,R2,,RmR_{1},R_{2},\ldots,R_{m} of side length rr such that the distance between any two of them is greater than rr. Since r=O(1)r=O(1), we can have m=Ω(N)m=\Omega(N).

For each RjR_{j}, there is at least one Pauli operator, denoted by PjP_{j}, with cPj0c_{P_{j}}\neq 0 such that suppPjRj\operatorname{supp}P_{j}\subseteq R_{j}. If there is more than one such Pauli operator, PjP_{j} with the smallest |suppPj||\operatorname{supp}P_{j}| is chosen. Let |ϕj|\phi_{j}\rangle be a pure product state of qubits in suppPj\operatorname{supp}P_{j} such that ϕj|cPjPj|ϕj=|cPj|\langle\phi_{j}|c_{P_{j}}P_{j}|\phi_{j}\rangle=-|c_{P_{j}}|.

Let ρS+\rho\in S^{+} be the tensor product of |ϕ1,|ϕ2,,|ϕm|\phi_{1}\rangle,|\phi_{2}\rangle,\ldots,|\phi_{m}\rangle and the maximally mixed state on the rest of the system so that tr(ρP)0\operatorname{tr}(\rho P)\neq 0 only if

suppPj=1msuppPj.\operatorname{supp}P\subseteq\bigcup_{j=1}^{m}\operatorname{supp}P_{j}. (21)

Since the supports of P1,P2,,PmP_{1},P_{2},\ldots,P_{m} are pairwise separated by distance rr,

tr(ρH)=j=1m|cPj|=Ω(N).\operatorname{tr}(\rho H)=-\sum_{j=1}^{m}|c_{P_{j}}|=-\Omega(N). (22)

Since there exists |ψSρS|\psi\rangle\in S_{\rho}\subseteq S such that ψ|H|ψtr(ρH)\langle\psi|H|\psi\rangle\leq\operatorname{tr}(\rho H), we complete the proof. ∎

The Hamiltonian (10) can be written as

H=iHi.H=\sum_{i}H_{i}. (23)

The sum is over lattice sites. Each term HiH_{i} is traceless trHi=0\operatorname{tr}H_{i}=0 and has bounded operator norm Hi=O(1)\|H_{i}\|=O(1) and is supported in a constant-radius neighborhood of site ii. Let d(i,j)d(i,j) be the distance between two lattice sites i,ji,j. At sufficiently high temperature, the thermal state (11) has exponential decay of correlations [13, 14, 15]. Therefore,

dE(β)dβ=tr2(ρβH)tr(ρβH2)=i,jtr(ρβHi)tr(ρβHj)tr(ρβHiHj)=i,jO(eΩ(d(i,j)))=O(N)E(β)=O(βN).\frac{\mathrm{d}E(\beta)}{\mathrm{d}\beta}=\operatorname{tr}^{2}(\rho_{\beta}H)-\operatorname{tr}(\rho_{\beta}H^{2})=\sum_{i,j}\operatorname{tr}(\rho_{\beta}H_{i})\operatorname{tr}(\rho_{\beta}H_{j})-\operatorname{tr}(\rho_{\beta}H_{i}H_{j})=-\sum_{i,j}O(e^{-\Omega(d(i,j))})\\ =-O(N)\implies E(\beta)=-O(\beta N). (24)
Lemma 2.

There is a constant c>0c>0 such that

|Sβ|=Ω(6N(1c|β|)),β[1/c,1/c].|S_{\beta}|=\Omega(6^{N(1-c|\beta|)}),\quad\forall\beta\in[-1/c,1/c]. (25)
Proof.

Assume without loss of generality that β0\beta\geq 0. If cc is sufficiently large, β\beta is sufficiently small so that (24) applies. We divide the lattice into two hyperrectangular regions AA and A¯\bar{A}. By applying Lemma 1 to region AA, we can have a product state of AA such that the energy of AA is <E(β)<E(\beta), as long as AA has O(βN)O(\beta N) sites for a sufficiently large constant hidden in the Big-O notation. Therefore, there exists a product state of AA such that the energy of AA is E(β)±O(1)E(\beta)\pm O(1). Combining this product state of AA with almost any product state of A¯\bar{A} results in a product state whose total energy is E(β)±o(N)E(\beta)\pm o(N). Thus, |Sβ|=Ω(6|A¯|)|S_{\beta}|=\Omega(6^{|\bar{A}|}), where |A¯||\bar{A}| is the number of sites in A¯\bar{A}. ∎

We are ready to prove Theorem 5:

𝔼|ψβmc1Dψeff=1|Sβ||ψSβ1Dψeff1|Sβ||ψS1Dψeff=|S||Sβ|𝔼|ψ1Dψeff(2/3)N×O(6cN|β|)=eΩ(N)\operatorname*{\mathbb{E}}_{|\psi\rangle\sim\mathcal{E}^{\textnormal{mc}}_{\beta}}\frac{1}{D^{\textnormal{eff}}_{\psi}}=\frac{1}{|S_{\beta}|}\sum_{|\psi\rangle\in S_{\beta}}\frac{1}{D^{\textnormal{eff}}_{\psi}}\leq\frac{1}{|S_{\beta}|}\sum_{|\psi\rangle\in S}\frac{1}{D^{\textnormal{eff}}_{\psi}}=\frac{|S|}{|S_{\beta}|}\operatorname*{\mathbb{E}}_{|\psi\rangle\sim\mathcal{E}}\frac{1}{D^{\textnormal{eff}}_{\psi}}\leq(2/3)^{N}\times O(6^{cN|\beta|})=e^{-\Omega(N)} (26)

if |β|15c|\beta|\leq\frac{1}{5c}.

4.2 Proof of Theorem 7

Lemma 3.
ρS+:n(ρ)=Ω(N)p(ρ)=1eΩ(N).\sum_{\rho\in S^{+}:n(\rho)=\Omega(N)}p(\rho)=1-e^{-\Omega(N)}. (27)
Proof.

A scheme for sampling ρS+\rho\in S^{+} with probability p(ρ)p(\rho) is given in Ref. [11]. The scheme has NN iterations. In each iteration, it samples the state of a new qubit from {|x±,|y±,|z±,I2/2}\{|x^{\pm}\rangle,|y^{\pm}\rangle,|z^{\pm}\rangle,I_{2}/2\}. It suffices to prove that the probability of sampling I2/2I_{2}/2 is always at least a positive constant. Then, the lemma follows from the Chernoff bound.

The procedure for sampling the state of a new qubit is Algorithm 6.3 in Ref. [11]. Please refer to that algorithm while reading the rest of the proof. In that algorithm, jj is the index of the qubit being sampled in the current step; E,E1,E2E,E_{1},E_{2} are Pauli operators (not energy). In step 3, they sample from seven cases, each of which has a constant probability. The probability may be adjusted in Step 8, but the adjustment is only a positive constant factor. EE^{\prime} is a Pauli operator computed from E,E1,E2E,E_{1},E_{2}. The formula for computing EE^{\prime} is different in different cases. Ej{I2,σx,σy,σz}E^{\prime}_{j}\in\{I_{2},\sigma^{x},\sigma^{y},\sigma^{z}\} is the restriction of EE^{\prime} to qubit jj and determines which state to sample for qubit jj.

Let E0,j,E1,j,E2,jE_{0,j},E_{1,j},E_{2,j} be the restriction of E,E1,E2E,E_{1},E_{2} to qubit jj, respectively. EjE^{\prime}_{j} can be calculated from E0,j,E1,j,E2,jE_{0,j},E_{1,j},E_{2,j} using the formulas in Step 3. By enumerating (E0,j,E1,j,E2,j){I2,σx,σy,σz}3(E_{0,j},E_{1,j},E_{2,j})\in\{I_{2},\sigma^{x},\sigma^{y},\sigma^{z}\}^{3}, we observe that among the seven cases, at least one of them leads to Ej=I2E^{\prime}_{j}=I_{2}. Hence, the probability of sampling I2/2I_{2}/2 is always at least a positive constant. ∎

Lemma 4.

For any ρS+\rho\in S^{+},

1|Sρ||ψSρ1Dψeff(2/3)n(ρ).\frac{1}{|S_{\rho}|}\sum_{|\psi\rangle\in S_{\rho}}\frac{1}{D^{\textnormal{eff}}_{\psi}}\leq(2/3)^{n(\rho)}. (28)
Proof.

Let AA be the set of qubits where ρ\rho is a pure state (not I2/2I_{2}/2) and A¯\bar{A} be the rest of the system. Then, states in SρS_{\rho} can be parameterized as

|ϕAjA¯|ϕj,|ϕj{|x±,|y±,|z±},|\phi\rangle_{A}\otimes\bigotimes_{j\in\bar{A}}|\phi_{j}\rangle,\quad|\phi_{j}\rangle\in\{|x^{\pm}\rangle,|y^{\pm}\rangle,|z^{\pm}\rangle\}, (29)

where |ϕA|\phi\rangle_{A} is a fixed pure product state of subsystem AA determined by ρ\rho.

It can be directly verified that

(|x+x+|)2+(|xx|)2+(|y+y+|)2+(|yy|)2+(|z+z+|)2+(|zz|)2=2Πsym,(|x^{+}\rangle\langle x^{+}|)^{\otimes 2}+(|x^{-}\rangle\langle x^{-}|)^{\otimes 2}+(|y^{+}\rangle\langle y^{+}|)^{\otimes 2}+(|y^{-}\rangle\langle y^{-}|)^{\otimes 2}+(|z^{+}\rangle\langle z^{+}|)^{\otimes 2}+(|z^{-}\rangle\langle z^{-}|)^{\otimes 2}\\ =2\Pi_{\textnormal{sym}}, (30)

where Πsym=Πsym2\Pi_{\textnormal{sym}}=\Pi_{\textnormal{sym}}^{2} is the projector onto the symmetric subspace of two qubits. Therefore,

|ψSρ|j|ψ|4=|ψSρj|2(|ψψ|)2|j2=j|2|ψSρ(|ψψ|)2|j2\displaystyle\sum_{|\psi\rangle\in S_{\rho}}\big{|}\langle j|\psi\rangle\big{|}^{4}=\sum_{|\psi\rangle\in S_{\rho}}\langle j|^{\otimes 2}(|\psi\rangle\langle\psi|)^{\otimes 2}|j\rangle^{\otimes 2}=\langle j|^{\otimes 2}\sum_{|\psi\rangle\in S_{\rho}}(|\psi\rangle\langle\psi|)^{\otimes 2}|j\rangle^{\otimes 2}
=j|2((|ϕAϕ|A)2jA¯|ϕj(|ϕjϕj|)2)|j2=j|2((|ϕAϕ|A)2(2Πsym)n(ρ))|j2\displaystyle=\langle j|^{\otimes 2}\big{(}(|\phi\rangle_{A}\langle\phi|_{A})^{\otimes 2}\otimes\bigotimes_{j\in\bar{A}}\sum_{|\phi_{j}\rangle}(|\phi_{j}\rangle\langle\phi_{j}|)^{\otimes 2}\big{)}|j\rangle^{\otimes 2}=\langle j|^{\otimes 2}\big{(}(|\phi\rangle_{A}\langle\phi|_{A})^{\otimes 2}\otimes(2\Pi_{\textnormal{sym}})^{\otimes n(\rho)}\big{)}|j\rangle^{\otimes 2}
2n(ρ)j|2((|ϕAϕ|A)2IA¯2)|j22n(ρ)j|(|ϕAϕ|AIA¯)|j,\displaystyle\leq 2^{n(\rho)}\langle j|^{\otimes 2}\big{(}(|\phi\rangle_{A}\langle\phi|_{A})^{\otimes 2}\otimes I_{\bar{A}}^{\otimes 2}\big{)}|j\rangle^{\otimes 2}\leq 2^{n(\rho)}\langle j|(|\phi\rangle_{A}\langle\phi|_{A}\otimes I_{\bar{A}})|j\rangle, (31)

where IA¯I_{\bar{A}} is the identity operator on A¯\bar{A}. Summing over jj,

1|Sρ||ψSρ1Dψeff16n(ρ)j2n(ρ)j|(|ϕAϕ|AIA¯)|j=(2/3)n(ρ).\frac{1}{|S_{\rho}|}\sum_{|\psi\rangle\in S_{\rho}}\frac{1}{D^{\textnormal{eff}}_{\psi}}\leq\frac{1}{6^{n(\rho)}}\sum_{j}2^{n(\rho)}\langle j|(|\phi\rangle_{A}\langle\phi|_{A}\otimes I_{\bar{A}})|j\rangle=(2/3)^{n(\rho)}. (32)

Theorem 7 follows directly from Lemmas 3, 4.

Acknowledgments

I would like to thank Aram W. Harrow for collaboration on closely related work [8]. This work was supported by the Army Research Office (grant no. W911NF-21-1-0262).

References

  • [1] Christian Gogolin and Jens Eisert “Equilibration, thermalisation, and the emergence of statistical mechanics in closed quantum systems” In Reports on Progress in Physics 79.5 IOP Publishing, 2016, pp. 056001 DOI: 10.1088/0034-4885/79/5/056001
  • [2] Hal Tasaki “From Quantum Dynamics to the Canonical Distribution: General Picture and a Rigorous Example” In Physical Review Letters 80.7 American Physical Society, 1998, pp. 1373–1376 DOI: 10.1103/PhysRevLett.80.1373
  • [3] Peter Reimann “Foundation of Statistical Mechanics under Experimentally Realistic Conditions” In Physical Review Letters 101.19 American Physical Society, 2008, pp. 190403 DOI: 10.1103/PhysRevLett.101.190403
  • [4] Noah Linden, Sandu Popescu, Anthony J. Short and Andreas Winter “Quantum mechanical evolution towards thermal equilibrium” In Physical Review E 79.6 American Physical Society, 2009, pp. 061103 DOI: 10.1103/PhysRevE.79.061103
  • [5] Anthony J Short “Equilibration of quantum systems and subsystems” In New Journal of Physics 13.5 IOP Publishing, 2011, pp. 053009 DOI: 10.1088/1367-2630/13/5/053009
  • [6] Terry Farrelly, Fernando G… Brandão and Marcus Cramer “Thermalization and Return to Equilibrium on Finite Quantum Lattice Systems” In Physical Review Letters 118.14 American Physical Society, 2017, pp. 140601 DOI: 10.1103/PhysRevLett.118.140601
  • [7] H. Wilming, M. Goihl, I. Roth and J. Eisert “Entanglement-Ergodic Quantum Systems Equilibrate Exponentially Well” In Physical Review Letters 123.20 American Physical Society, 2019, pp. 200604 DOI: 10.1103/PhysRevLett.123.200604
  • [8] Yichen Huang and Aram W. Harrow “Scrambling and thermalization in translation-invariant systems” arXiv:1907.13392
  • [9] Yichen Huang “Extensive entropy from unitary evolution” arXiv:2104.02053 In Preprints 2021, pp. 2021040254
  • [10] Yichen Huang “Long-time properties of generic Floquet systems are approximately periodic with the driving period” In New Journal of Physics 26.7 IOP Publishing, 2024, pp. 072001 DOI: 10.1088/1367-2630/ad5eb1
  • [11] Ainesh Bakshi, Allen Liu, Ankur Moitra and Ewin Tang “High-Temperature Gibbs States are Unentangled and Efficiently Preparable” arXiv:2403.16850
  • [12] Saul Pilatowsky-Cameo and Soonwon Choi “Quantum thermalization of translation-invariant systems at high temperature” arXiv:2409.07516
  • [13] Huzihiro Araki “Gibbs states of a one dimensional quantum lattice” In Communications in Mathematical Physics 14.2, 1969, pp. 120–157
  • [14] Andreas Bluhm, Ángela Capel and Antonio Pérez-Hernández “Exponential decay of mutual information for Gibbs states of local Hamiltonians” In Quantum 6 Verein zur Förderung des Open Access Publizierens in den Quantenwissenschaften, 2022, pp. 650 DOI: 10.22331/q-2022-02-10-650
  • [15] M. Kliesch et al. “Locality of Temperature” In Physical Review X 4.3 American Physical Society, 2014, pp. 031019 DOI: 10.1103/PhysRevX.4.031019