arXiv:1806.10871v1 [quant-ph] 28 Jun 2018 · arXiv:1806.10871v1 [quant-ph] 28 Jun 2018 Simulating...

13
arXiv:1806.10871v1 [quant-ph] 28 Jun 2018 Simulating dynamic quantum phase transitions in photonic quantum walks Kunkun Wang, 1, 2 Xingze Qiu, 3, 4 Lei Xiao, 1, 2 Xiang Zhan, 1, 2 Zhihao Bian, 1, 2 Wei Yi, 3, 4, and Peng Xue 1, 2, 5, 1 Beijing Computational Science Research Center, Beijing 100084, China 2 Department of Physics, Southeast University, Nanjing 211189, China 3 Key Laboratory of Quantum Information, University of Science and Technology of China, CAS, Hefei 230026, China 4 Synergetic Innovation Center in Quantum Information and Quantum Physics, University of Science and Technology of China, CAS, Hefei 230026, China 5 State Key Laboratory of Precision Spectroscopy, East China Normal University, Shanghai 200062, China Signaled by non-analyticities in the time evolution of physical observables, dynamic quantum phase transitions (DQPTs) emerge in quench dynamics of topological systems and possess an inter- esting geometric origin captured by dynamic topological order parameters (DTOPs). In this work, we report the experimental study of DQPTs using discrete-time quantum walks of single photons. We simulate quench dynamics between distinct Floquet topological phases using quantum-walk dy- namics, and experimentally characterize DQPTs and the underlying DTOPs through interference- based measurements. The versatile photonic quantum-walk platform further allows us to experi- mentally investigate DQPTs for mixed states and in parity-time-symmetric non-unitary dynamics for the first time. Our experiment directly confirms the relation between DQPTs and DTOPs in quench dynamics of a topological system, and opens up the avenue of simulating emergent topolog- ical phenomena using discrete-time quantum-walk dynamics. The study of phase transitions lies at the core of the description of equilibrium states of matter [1]. Besides conventional continuous phase transitions that are sig- naled by symmetry breaking, topological phase transi- tions, characterized by the change of topology in their ground-state wavefunctions, have attracted much atten- tion since the discovery of quantum Hall effects [2, 3]. Re- cent experimental progress has further led to the exciting possibility of creating novel quantum phases of matter in dynamical processes [4–18], and thus raised the challeng- ing question on the understanding of emergent phases and phase transitions in non-equilibrium dynamics. Proposed as temporal analogues to continuous phase transitions, dynamical quantum phase transitions (DQPTs) are associated with non-analyticities in the time evolution of physical observables [19–21], and have been experimentally observed in various systems [15–18]. DQPT occurs as a consequence of the emergence of dy- namic Fisher zeros [22, 23], where the Loschmidt ampli- tude G(t)= ψ(0)|ψ(t)vanishes at critical times and the corresponding rate function g(t)= 1/N ln |G(t)| 2 becomes non-analytical [21]. Here |ψ(t)is the time- evolved state, and N is the overall degrees of freedom of the system. Whereas it is still unclear to what ex- tent key concepts of continuous phase transitions can be extended to describe DQPTs, an intriguing discovery is the geometric origin of DQPTs, captured by dynamic topological order parameters (DTOPs), which suggests the intimate connection between DQPTs and emergent topological phenomena in dynamic processes [24–27]. A particularly important scenario is the quench dy- namics of topological systems, where the ground state |ψ i of the initial Hamiltonian H i is time-evolved un- der the final Hamiltonian H f . Here two different types of DQPTs can occur: topological DQPTs, whose occur- rence is intimately related to the topology of H i and H f ; and accidental DQPTs, which are to the contrary. Specif- ically, for quench dynamics of one-dimensional topolog- ical systems, topological DQPTs necessarily exist when ground states of H i and H f belong with distinct topolog- ical phases [28–30]. These topological DQPTs provide a crucial link between static topological phases and emer- gent topological phenomena in quench dynamics, and represent an exemplary case where the relation between topology and dynamics can be investigated. In two di- mensions, topological DQPTs recently observed in cold atomic gases are associated with dynamic vortices [8], which have been viewed the effective DTOP. In one di- mension, on the other hand, relations between DQPTs and DTOPs have yet to be experimentally investigated. In this work, we report the experimental simula- tion of topological DQPTs using discrete-time quan- tum walks (QWs) of single photons in one dimension. We map single-photon QW dynamics to quenches be- tween Floquet topological phases (FTPs) [31], and probe inner products of the initial and time-evolved states via inference-based measurements. We then investigate DQPTs by constructing quantities such as the rate func- tion and DTOPs from our measurements. An advantage of photonic QW dynamics lies in the relative ease of in- troducing decoherence and loss, which further allows us to experimentally investigate DQPTs for mixed states and in non-unitary quench processes. Consistent with theoretical predictions [25, 32], we find DQPTs persist in unitary quench dynamics of mixed states, as well as in parity-time (PT )-symmetric non-unitary quench dynam- ics in the PT -symmetry-unbroken regime. In both cases, however, the behaviors of DTOPs are quite different. Simulating quench dynamics between FTPs:— We study DQPTs in quench dynamics using discrete-time

Transcript of arXiv:1806.10871v1 [quant-ph] 28 Jun 2018 · arXiv:1806.10871v1 [quant-ph] 28 Jun 2018 Simulating...

arX

iv:1

806.

1087

1v1

[qu

ant-

ph]

28

Jun

2018

Simulating dynamic quantum phase transitions in photonic quantum walks

Kunkun Wang,1, 2 Xingze Qiu,3, 4 Lei Xiao,1, 2 Xiang Zhan,1, 2 Zhihao Bian,1, 2 Wei Yi,3, 4, ∗ and Peng Xue1, 2, 5, †

1Beijing Computational Science Research Center, Beijing 100084, China2Department of Physics, Southeast University, Nanjing 211189, China

3Key Laboratory of Quantum Information, University of Science and Technology of China, CAS, Hefei 230026, China4Synergetic Innovation Center in Quantum Information and Quantum Physics,

University of Science and Technology of China, CAS, Hefei 230026, China5State Key Laboratory of Precision Spectroscopy,

East China Normal University, Shanghai 200062, China

Signaled by non-analyticities in the time evolution of physical observables, dynamic quantumphase transitions (DQPTs) emerge in quench dynamics of topological systems and possess an inter-esting geometric origin captured by dynamic topological order parameters (DTOPs). In this work,we report the experimental study of DQPTs using discrete-time quantum walks of single photons.We simulate quench dynamics between distinct Floquet topological phases using quantum-walk dy-namics, and experimentally characterize DQPTs and the underlying DTOPs through interference-based measurements. The versatile photonic quantum-walk platform further allows us to experi-mentally investigate DQPTs for mixed states and in parity-time-symmetric non-unitary dynamicsfor the first time. Our experiment directly confirms the relation between DQPTs and DTOPs inquench dynamics of a topological system, and opens up the avenue of simulating emergent topolog-ical phenomena using discrete-time quantum-walk dynamics.

The study of phase transitions lies at the core of thedescription of equilibrium states of matter [1]. Besidesconventional continuous phase transitions that are sig-naled by symmetry breaking, topological phase transi-tions, characterized by the change of topology in theirground-state wavefunctions, have attracted much atten-tion since the discovery of quantum Hall effects [2, 3]. Re-cent experimental progress has further led to the excitingpossibility of creating novel quantum phases of matter indynamical processes [4–18], and thus raised the challeng-ing question on the understanding of emergent phasesand phase transitions in non-equilibrium dynamics.Proposed as temporal analogues to continuous

phase transitions, dynamical quantum phase transitions(DQPTs) are associated with non-analyticities in thetime evolution of physical observables [19–21], and havebeen experimentally observed in various systems [15–18].DQPT occurs as a consequence of the emergence of dy-namic Fisher zeros [22, 23], where the Loschmidt ampli-tude G(t) = 〈ψ(0)|ψ(t)〉 vanishes at critical times andthe corresponding rate function g(t) = −1/N ln |G(t)|2becomes non-analytical [21]. Here |ψ(t)〉 is the time-evolved state, and N is the overall degrees of freedomof the system. Whereas it is still unclear to what ex-tent key concepts of continuous phase transitions can beextended to describe DQPTs, an intriguing discovery isthe geometric origin of DQPTs, captured by dynamictopological order parameters (DTOPs), which suggeststhe intimate connection between DQPTs and emergenttopological phenomena in dynamic processes [24–27].A particularly important scenario is the quench dy-

namics of topological systems, where the ground state|ψi〉 of the initial Hamiltonian H i is time-evolved un-der the final Hamiltonian H f. Here two different typesof DQPTs can occur: topological DQPTs, whose occur-

rence is intimately related to the topology of H i and H f;and accidental DQPTs, which are to the contrary. Specif-ically, for quench dynamics of one-dimensional topolog-ical systems, topological DQPTs necessarily exist whenground states of H i and H f belong with distinct topolog-ical phases [28–30]. These topological DQPTs provide acrucial link between static topological phases and emer-gent topological phenomena in quench dynamics, andrepresent an exemplary case where the relation betweentopology and dynamics can be investigated. In two di-mensions, topological DQPTs recently observed in coldatomic gases are associated with dynamic vortices [8],which have been viewed the effective DTOP. In one di-mension, on the other hand, relations between DQPTsand DTOPs have yet to be experimentally investigated.In this work, we report the experimental simula-

tion of topological DQPTs using discrete-time quan-tum walks (QWs) of single photons in one dimension.We map single-photon QW dynamics to quenches be-tween Floquet topological phases (FTPs) [31], and probeinner products of the initial and time-evolved statesvia inference-based measurements. We then investigateDQPTs by constructing quantities such as the rate func-tion and DTOPs from our measurements. An advantageof photonic QW dynamics lies in the relative ease of in-troducing decoherence and loss, which further allows usto experimentally investigate DQPTs for mixed statesand in non-unitary quench processes. Consistent withtheoretical predictions [25, 32], we find DQPTs persist inunitary quench dynamics of mixed states, as well as inparity-time (PT )-symmetric non-unitary quench dynam-ics in the PT -symmetry-unbroken regime. In both cases,however, the behaviors of DTOPs are quite different.Simulating quench dynamics between FTPs:— We

study DQPTs in quench dynamics using discrete-time

2

State Preparation

2( /2)C

Evolution

2( /2)C

22.5° 22.5°

t=1 t=7

Evolution

Trigger

Compensated Crystal

BBO

PBS NPBS

HWP QWP

PPBS

BD

APD

Measurement

45°

Path 1

Path 2

D1

D0

D2

D3

D4

H0 H1Q1

H2 Q2

1( /2)C

1( /2)C

(a) (b)

! 0 !

1

!

0

!

2

0

0

2 2

2

2

0

0 0

2

2

0

FIG. 1. (a) Experimental setup for the simulation of DQPTs using QWs. Pairs of single photons are generated via type-Ispontaneous parametric down conversion (SPDC) using a non-linear β-Barium-Borate (BBO) crystal. One photon serves as atrigger and the other signal photon is prepared in an arbitrary linear polarization state using polarizing beam splitters (PBSs),wave plates (WPs) with certain setting angles and a non-polarizing beam splitter (NPBS). Coin rotations and conditionaltranslations are realized by two half-wave plates (HWPs) and a beam displacer (BD), respectively. For non-unitary QWs,a sandwich-type HWP-PPBS-HWP setup is inserted to introduce the partial measurement, where PPBS is an abbreviationfor partially polarizing beam splitters. Avalanche photodiodes (APDs) detect the signal and heralding photons. (b) Phase

diagram for QWs governed by Floquet operators U and U , labeled by the winding number ν as a function of coin parameters(θ1, θ2). Note that topological phase boundaries and winding numbers for unitary QW dynamics governed by U and non-unitary

dynamics governed by U are the same. Dashed red lines represent boundaries between PT -symmetry-unbroken and brokenregimes for U , with PT -symmetry-broken regimes lying inbetween the red lines near topological phase boundaries, which arerepresented by solid black lines. Black star represents coin parameters of the initial Floquet operator U i or U i, other symbolsindicate coin parameters of final Floquet operators in different cases.

QWs on a one-dimensional homogeneous lattice L (L ∈Z), where we use polarization states of single photons|H〉 , |V 〉 to represent coin states and spatial modes toencode walker states. As illustrated in Fig. 1(a), themain component of our setup is a cascaded interferomet-ric network. The resulting QW dynamics is governed bythe Floquet operator

U = C(θ1/2)SC(θ2)SC(θ1/2). (1)

Here, the coin operator C(θ) rotates the single-photonpolarization by θ about the y-axis. The shift operator Smoves the walker in |H〉 (|V 〉) to the left (right) by onelattice site.QWs governed by U support non-trivial FTPs, which

are characterized by winding numbers [9, 31, 33]. Topo-logical properties of U can be understood by consider-ing the effective Hamiltonian Heff defined through U =e−iHeff , where, for homogeneous QWs, Heff(k) = Ekh ·σin quasimomentum k space. Here σ is the Pauli vector,±Ek are the quasienergies, and h marks the direction ofthe spinor eigenvector at each quasi-momentum k. AsHeff satisfies chiral symmetry with ΓHeffΓ = −Heff andΓ = σx, winding numbers are defined as the number oftimes h winds around the x-axis as k varies through thefirst Brillouin zone (1BZ). As illustrated in Fig. 1(b), byvarying coin parameters (θ1, θ2), the system can changebetween FTPs with distinct winding numbers.For a typical QW process, the photon is initialized in

a local state |ψi〉 at x = 0, and is subject to repeatedoperations of U , such that at the t-th step, the photon isin the state |ψ(t)〉 = U t|ψi〉 = e−iHefft|ψi〉. Importantly,

if we choose |ψi〉 to be an eigenstate of U i = e−iHieff , the

resulting QW dynamics realize stroboscopic simulationof quenches between FTPs associated with H i

eff and Heff,respectively. For the discrete-time QW protocol consid-ered here, Floquet operators U i having localized eigen-states exist, whose coin parameters are on the horizontalblack dashed lines in Fig. 1(b). The corresponding H i

eff

are topologically trivial with νi = 0. For contrast, in thefollowing, we label the Floquet operator actually driv-ing the QW as U f, with νf the corresponding windingnumber.

Initialization and detection:— Experimentally, We ini-tialize the walker photon at x = 0, with its coin stategiven by the density matrix ρ0 = p |ψi

−〉 〈ψi−| + (1 −

p) |ψi+〉 〈ψi

+|, where |ψi±〉 = (|H〉 ∓ i |V 〉)/

√2. The ini-

tial state is therefore a pure state when p = 0, 1, anda mixed state otherwise. Importantly, |x = 0〉 ⊗ |ψi

±〉are eigenstates of U i with the coin parameters (θi1 =π/4, θi2 = −π/2). We then implement QWs governedby U f with coin parameters (θf1, θ

f2). To reduce experi-

mental error, we choose (θf1, θf2) on blue dash-dotted lines

in Fig. 1(b), as the spatial spread of the resulting QWdynamics is small.

Due to the lattice-translational symmetry, time evo-lutions in different quasimomentum k-sectors are decou-pled and governed by U f

k, the Fourier component of U f.We construct the Loschmidt amplitude G(k, t) in eachquasimomentum k-sector according to

G(k, t) := Tr[

ρ0(

U fk

)t]

=∑

x

e−ikxP (x, t), (2)

3

0 2 4 6 8t

(t)

m

(a) (b)

0 2 4 6 8t

1,3 theor1,3 expt2,4 theor2,4 expt

(t)

0

2

-2

0.0

0.5

1.0

1.5

0

2

-2

0.0

0.5

1.0

1.5expt

theorg

g

g

FIG. 2. Rate function (upper layer) and νm(t) (lower layer)of seven-step unitary QWs as functions of time steps. Theinitial state of the walker-coin system is |x = 0〉⊗ |ψi

−〉. QWsare governed by U f with (θf1 = −π/2, θf2 = 3π/8) (a) and U f

with (θf1 = −π/2, θf2 = π/4) (b), respectively. Error bars arederived from simulations where we consider all the systematicinaccuracies of the experiment.

where P (p, x, t) = p〈ψi−|ψ−(x, t)〉+(1− p)〈ψi

+|ψ+(x, t)〉,and |ψ±(x, t)〉 =

k eikx

(

U fk

)t |ψi±〉. Experimentally,

P (p, x, t) is measured by performing interference-basedmeasurements at the t-th step [33, 34]. More specifically,after the photons pass through the QW interferometricnetwork, we project the polarization state of the pho-tons at each position x onto the initial polarization stateof the photons at x = 0, and perform coincidence mea-surements on the walker photons and the trigger photonssuccessively up to t by single-photon avalanche photodi-odes (APDs).We then construct the rate function according to

g(t) = −∑

k∈1BZ ln |G(k, t)|2, where we have usedG(t) =∏

k∈1BZ G(k, t). Hence by construction, g(t) correspondsto the rate function of a quench starting from a half-filledFloquet band, where the initial state is a direct productof single-particle density matrices ρ0 in different k-sectorsof 1BZ. This is in contrast to the case of single-photonQW dynamics, where the initial state is a superpositionof coin states in different k-sectors. We therefore empha-size that whereas DQPTs do not actually occur in QWsof single photons, we can simulate DQPTs in quench dy-namics of topological systems using the setup.From the measured G(k, t), we further calculate

DTOPs characterizing DQPTs. In one dimension,DTOPs are defined as [24]

νm(t) =1

∫ km+1

km

∂φGk (t)

∂kdk, (3)

where the Pancharatnam geometric phase (PGP)

φGk (t) = φk(t) − φdynk (t). Here φk(t) is defined through

G(k, t) = |G(k, t)|eiφk(t), and φdynk (t) is the dynamicphase. km (m = 1, 2, ...) are fixed points of the dynamics,where the corresponding density matrices do not evolvein time and PGP vanishes at all times. νm(t) thereforecharacterizes the S1 → S1 mapping from the momen-

0 2 4 6 8t

0

-1

0

1

-1

(t)

m

1

(a) (b)

0 2 4 6 8t

expt

theor

(t)

0

2

1

3

1,3 theor1,3 expt2,4 theor2,4 expt

0

2

1

3

g

g

g

FIG. 3. Rate function (upper layer) and νm(t) (lower layer)of a seven-step unitary QW. The walker starts at x = 0 andthe QW is governed by the final Floquet operator U f with(θf1 = −π/2, θf2 = 3π/8). The initial coin state is a mixedstate with p = 0.9 (a) and p = 0.7 (b), respectively. Errorbars are derived from simulations where we consider all thesystematic inaccuracies of the experiment.

0.0

0.2

-0.2

(t)

m

0 2 4 6 8t

(b)

0.0

0.6

0.9

0 2 4 6 8t

(a)

0.3

(t)

expt

theor1,3 theor1,3 expt2,4 theor2,4 expt

g

g

g

FIG. 4. (a) Rate function and (b) νm(t) of a seven-step uni-tary QW. The QW is governed by the final Floquet operatorU f with (θf1 = −π/16, θf2 = −3π/16), which is in the sameFTP as the initial state. Error bars are derived from simula-tions where we consider all the systematic inaccuracies of theexperiment.

tum submanifold between km and km+1 to eiφGk (t). The

DTOP is quantized and can only change value at DQPTs,where G(kc, tc) = 0 and φGkc

(tc) becomes ill-defined atcritical kc and tc. While fixed points necessarily existwhen U i and U f have different winding numbers, kc ex-ists between adjacent fixed points and leads to DQPTsat tc = (2n− 1)t0 (n ∈ N), where the critical time scalet0 = π/(2Ef

kc) and ±Ef

k is the quasienergy of U fk. As a

result, νm(t) exhibits abrupt jumps at the DQPTs asso-ciated with kc ∈ (km, km+1).

DQPT in unitary dynamics:— We first study DQPTsfor pure states in unitary dynamics. We initialize photonsin the coin state |ψi

−〉 at x = 0. The photons are thensubject to unitary time evolutions governed by the Flo-quet operator U f with (θf1 = −π/2, θf2 = 3π/8). This cor-responds to a quench between FTPs with νi = 0 and νf =−2. Here, the fixed points k1,2,3,4 = −π,−π/2, 0, π/2,and kc = ±π/4,±3π/4. Note U f

k has a discrete sym-metry U f

k = U fk+π in addition to the time-reversal sym-

metry. Under these symmetries, Efkc

are degenerate andthere is only one critical time scale t0 = 4. In Fig. 2(a),we show the rate function, which becomes non-analytic

4

0

2

-

0 2 4 6 8t

0.0

0.8

0 2 4 6 8t

(a)

0.4

0 2 4 6 8t

0.0

((

(t)

m

1.2

(t)

(t)

expt

theor1,3 theor1,3 expt2,4 theor2,4 expt

g g

g

g

FIG. 5. (a) Rate function and νm(t) of a seven-step non-unitary QW with a loss parameter l = 0.36. The initial state of walker-

coin system is |0〉 ⊗ |ψ−〉. The QW is governed by the non-unitary Floquet operator U f with (θf1 = −π/3, θf2 = π/5). (b) Rate

function of the QW governed by U f with[

θf1 = −π/2, θf2 = (π − ξ)/2]

, where ξ = arccos(1/α) and α = (1+√1− l)/(2 4

√1− l).

The two critical time scales are t0 = 1.7183, 2.1482, which give rise to non-analyticities in the rate function as indicated by ver-tical dashed lines in (a). Theoretically calculated fixed points are located at k1,2,3,4 = −1.0094π,−0.4470π,−0.0094π, 0.5530π,and the critical momenta kc = −0.7888π,−0.1534π, 0.2112π, 0.8466π. Experimental errors are due to photon-counting statis-tics.

at the first critical time tc = t0. Whereas it is difficult todirectly identify non-analyticities of g(t) in discrete-timedynamics, DQPTs are unambiguously revealed by jumpsin the quantized DTOP across tc. The abrupt jump isconfirmed by particularly large error bars in the mea-sured νm(t) at the critical time [33]. Further, due to thesymmetry of U f

k, we have ν1,3(t) = −ν2,4(t), where ν4(t)is integrated in the range (π/2, π).

We then fix the initial coin state and change the finalFloquet operator to U f with (θf1 = −π/2, θf2 = π/4).As shown in Fig. 2(b), the critical time scale changes tot0 = 2, and the rate function becomes non-analytic atodd multiples of t0. The quantized DTOPs also featureabrupt jumps at critical times. As locations of km and kcare the same as those of the previous case, there is onlyone critical time scale as well.

In the second case study, we initialize photons at x = 0and in a mixed coin state characterized by ρ0 with p = 0.7and p = 0.9, respectively. The QW is governed by U f

with (θf1 = −π/2, θf2 = 3π/8). Whereas the resultingQW dynamics still correspond to quenches between FTPswith νi = 0 and νf = −2, coin states of time-evolvedstates remain mixed. We show the resulting rate func-tions and DTOPs in Fig. 3. Whereas the occurrence ofDQPTs are still signaled by non-analyticities in the ratefunctions, DTOPs are no longer quantized. This is be-

cause PGPs do not vanish at km, such that eiφGk no longer

forms a closed S1 manifold between km and km+1. Con-sequently, νm(t) is no longer the winding number char-acterizing such a map. These results are consistent withprevious theoretical studies [25]. Note locations of kmand kc are the same as in the previous cases.

For comparison, we also study the case where thequench dynamics is between FTPs with νi = 0 andνf = 0. For this purpose, we choose U f with (θf1 =−π/16, θf2 = −3π/16). As shown in Fig. 4, the rate func-tion is smooth in time and νm(t) remains zero, indicatingthe absence of DQPTs. Here km = 0,±π/2, π.DQPT in PT -symmetric non-unitary dynamics:—

The ease of introducing loss in photonics further allows

us to explore DQPTs in non-unitary dynamics [32, 35].We enforce non-unitary dynamics by performing a par-tial measurement Me = 1w ⊗

√l |−〉 〈−| in the basis

|±〉 at each time step, with 1w =∑

x |x〉 〈x|, |±〉 =

(|H〉 ± |V 〉)/√2, and l the loss parameter, which is fixed

at l = 0.36 in our experiment. The non-unitary QW isthen governed by

U = γC(θ1/2)SC(θ2/2)MC(θ2/2)SC(θ1/2), (4)

where M = 1w ⊗(

|+〉 〈+|+√1− l |−〉 〈−|

)

, and γ =

(1− l)−1/4.Topological properties of U are characterized by wind-

ing numbers defined through the global Berry phase [36–38]. The resulting topological phase diagram is the sameas that of U [33]. Crucially, U also possess PT sym-metry, therefore its quasienergy spectra can be entirelyreal in the PT -symmetry-unbroken regime [39–41]. Inthe regime with spontaneously broken PT symmetry, onthe other hand, quasienergies of U can be complex. Theboundary between regimes with unbroken and brokenPT symmetry is plotted in Fig. 1(b) as red dashed lines,with PT -symmetry-broken regimes surrounding topolog-ical phase boundaries. It can be shown that DQPTsnecessarily occur for quench processes between distinctFTPs in the PT -symmetry-unbroken regime [32, 33].Similar to the unitary case, we initialize photons in

the state |x = 0〉 ⊗ |ψi−〉, with the corresponding U i in

the PT -symmetry-unbroken regime with νi = 0. Thewalker is evolved under the final non-unitary Floquetoperator U f with (θf1 = −π/3, θf2 = π/5), which is inthe PT -symmetry-unbroken regime with νf = −2. TheLoschmidt amplitude, the rate function g(t), and theDTOP νm(t) can be constructed similar to the unitarycase [33], albeit QW dynamics is now non-unitary. Asillustrated in Fig. 5(a), non-analyticities in the rate func-tion have two distinct time scales, which correspond totwo different DTOPs [see Fig. 5(b)], both quantized anddemonstrating abrupt jumps at odd multiples of the cor-responding critical time scale.The emergence of two critical time scales is intimately

5

connected with the breaking of time-reversal symmetryof the non-unitary dynamics [32]. In this case, whereasfixed points still exist when U i and U f are in the PT -symmetry-unbroken regime and feature distinct FTPs,they are no longer located at high-symmetry points andhave to be solved numerically. As a consequence, ν1,3(t)and ν2,4(t) feature jumps at different critical times, giv-ing rise to two critical time scales.Finally, we study the case when the final non-unitary

Floquet operator is in the PT -symmetry-broken regime.The resulting rate function is shown in Fig. 5(b), whereno DQPTs can be identified. As fixed points are alsoabsent in the dynamics, DTOPs cannot be defined inthis case [32].Conclusions:— We have demonstrated photonic QWs

as an ideal platform for the simulation of DQPTs inquench dynamics of topological systems. By construct-ing key quantities such as the rate function and DTOPsfrom interference measurements, we experimentally in-vestigate the relation between DQPTs and DTOPs inboth unitary and non-unitary quench dynamics. Our ex-periment opens up the avenue of investigating dynamictopological phenomena using QW dynamics, and pavesthe way for a more systematic study of DQPTs in novelsituations such as engineered non-unitary dynamics or inhigher dimensions.Acknowledgement:– This work has been supported

by the Natural Science Foundation of China (GrantNos. 11474049, 11674056, and 11522545) and the Nat-ural Science Foundation of Jiangsu Province (GrantNo. BK20160024). WY acknowledges supportfrom the National Key R&D Program (Grant Nos.2016YFA0301700,2017YFA0304100). KW and XQ con-tributed equally to this work.

[email protected][email protected]

[1] L. D. Landau, On the theory of phase transitions, Phys.Z. Sowjetunion 11, 26 (1937).

[2] M. Z. Hasan, and C. L. Kane, Colloquium: Topologicalinsulators, Rev. Mod. Phys. 82, 3045-3067 (2010).

[3] X. L. Qi, and S. C. Zhang, Topological insulators andsuperconductors, Rev. Mod. Phys. 83, 1057-1110 (2011).

[4] G. Jotzu, M. Messer, R. Desbuquois, M. Lebrat, T.Uehlinger, D. Greif, and T. Esslinger, Experimental real-ization of the topological Haldane model with ultracoldfermions, Nature 515, 237 (2014).

[5] N. Flaschner, B. S. Rem, M. Tarnowski, D. Vogel, D.-S.Luhmann, K. Sengstock, and C. Weitenberg, Experimen-tal reconstruction of the Berry curvature in a FloquetBloch band, Science 352, 1091 (2016).

[6] B. Song, L. Zhang, C. He, T. F. J. Poon, E. Hai-iyev, S. Zhang, X.-J. Liu, and G.-B. Jo, Observationof symmetry-protected topological band with ultracoldfermions, Sci. Adv. 4, eaao4748 (2018).

[7] M. Tarnowski, F. Nur-Unal, N. Flaschner, B. S. Rem, A.

Eckard, K. Sengstock, and C. Weitenberg, Characteriz-ing topology by dynamics: Chern number from linkingnumber, arXiv:1709.01046.

[8] N. Flaschner, D. Vogel, M. Tarnowski, B. S. Rem, D.-S.Luhmann, M. Heyl, J. C. Budich, L. Mathey, K. Seng-stock, and C. Weitenberg, Observation of dynamical vor-tices after quenches in a system with topology, Nat. Phys.14, 265 (2018)

[9] T. Kitagawa, M. A. Broome, A. Fedrizzi, M. S. Rudner,E. Berg, I. Kassal, A. Aspuru-Guzik, E. Demler, and A.G. White, Observation of topologically protected boundstates in photonic quantum walks, Nat. Commun. 3, 882(2012).

[10] F. Cardano, M. Maffei, F. Massa, B. Piccirillo, C. de Li-sio, G. D. Filippis, V. Cataudella, E. Santamato, and L.Marrucci, Statistical moments of quantum-walk dynam-ics reveal topological quantum transitions, Nat. Com-mun. 7, 11439 (2016).

[11] F. Cardano, A. D’Errico, A. Dauphin, M. Maffei, B. Pic-cirillo, C. de Lisio, G. D. Filippis, V. Cataudella, E. San-tamato, L. Marrucci, M. Lewenstein, and P. Massignan,Detection of Zak phases and topological invariants in achiral quantum walk of twisted photons, Nat. Commun.8, 15516 (2017).

[12] L. Xiao, X. Zhan, Z. H. Bian, K. K. Wang, X. Zhang, X.P. Wang, J. Li, K. Mochizuki, D. Kim, N. Kawakami, W.Yi, H. Obuse, B. C. Sanders, and P. Xue, Observation oftopological edge states in paritytime-symmetric quantumwalks, Nat. Phys. 13, 1117 (2017).

[13] J. M. Zeuner, M. C. Rechtsman, Y. Plotnik, Y. Lumer,S. Nolte, M. S. Rudner, M. Segev, and A. Szameit, Ob-servation of a topological transition in the bulk of a non-Hermitian System, Phys. Rev. Lett. 115, 040402 (2015).

[14] X. Zhan, L. Xiao, Z. Bian, K. Wang, X. Qiu, B. C.Sanders, W. Yi, and P. Xue, Detecting topological invari-ants in nonunitary discrete-time quantum walks, Phys.Rev. Lett. 119, 130501 (2017).

[15] P. Jurcevic, H. Shen, P. Hauke, C. Maier, T. Brydges,C. Hempel, B. P. Lanyon, M. Heyl, R. Blatt, and C. F.Roos, Direct observation of dynamical quantum phasetransitions in an interacting many-body system, Phys.Rev. Lett. 119, 080501 (2017).

[16] J. Zhang, G. Pagano, P. W. Hess, A. Kyprianidis, P.Becker, H. Kaplan, A. V. Gorshkov, Z.-X. Gong, and C.Monroe, Observation of a many-body dynamical phasetransition with a 53-qubit quantum simulator, Nature551, 601 (2017).

[17] H. Bernie, S. Schwartz, A. Keesling, H. Levine, A. Om-ran, H. Pichler, S. Choi, A. Zibrov, M. Endres, M.Greiner, V. Vuletic, and M. D. Lukin, Probing many-body dynamics on a 51-atom quantum simulator, Nature551, 579 (2017).

[18] X.-Y. Guo, C. Yang, Y. Zeng, Y. Peng, H.-K. Li, H.Deng, Y.-R. Jin, S. Chen, D. Zheng, and H. Fan, Ob-servation of dynamical quantum phase transition by asuperconducting qubit simulation, arXiv:1806.09269.

[19] M. Heyl, A. Polkovnikov, and S. Kehrein, Dynamicalquantum phase transitions in the transverse-field Isingmode, Phys. Rev. Lett. 110, 135704 (2013).

[20] M. Heyl, Scaling and universality at dynamical quantumphase transitions, Phys. Rev. Lett. 115, 140602 (2015).

[21] M. Heyl, Dynamical quantum phase transitions: a re-view, Rep. Prog. Phys. 81, 054001 (2018).

[22] M. E. Fisher, The Nature of Critical Points, Lectures in

6

Theoretical Physics Vol. 7 (University of Colorado Press,Boulder, CO, 1965).

[23] K. Brandner, V. F. Maisi, J. P. Pekola, J. P. Garrahan,and C. Flindt, Experimental determination of dynamicalLee-Yang zeros, Phys. Rev. Lett. 118, 180601 (2017).

[24] J. C. Budich, and M. Heyl, Dynamical topological or-der parameters far from equilibrium, Phys. Rev. B 93,085416 (2016).

[25] M. Heyl, and J. C. Budich, Dynamical topological quan-tum phase transitions for mixed states, Phys. Rev. B 96,180304 (2017).

[26] U. Bhattacharya, and A. Dutta, Emergent topology anddynamical quantum phase transitions in two-dimensionalclosed quantum systems, Phys. Rev. B 96, 014302 (2017).

[27] X. Qiu, T.-S. Deng, G.-C. Guo, and W. Yi, Dynamicaltopological invariants and reduced rate functions for dy-namical quantum phase transitions in two dimensions,arXiv:1804.09032.

[28] S. Vajna, and B. Dora, Topological classification of dy-namical phase transitions, Phys. Rev. B 91, 155127(2015).

[29] Z. Huang, and A. V. Balatsky, Dynamical quantum phasetransitions: role of topological nodes in wave functionoverlaps, Phys. Rev. Lett. 117, 086802 (2016).

[30] U. Bhattacharya, and A. Dutta, Interconnections be-tween equilibrium topology and dynamical quantumphase transitions in a linearly ramped Haldane model,Phys. Rev. B 95, 184307 (2017).

[31] T. Kitagawa, M. S. Rudner, E. Berg, and E. Demler,Exploring topological phases with quantum walks, Phys.Rev. A 82, 033429 (2010).

[32] X. Qiu, T.-S. Deng, Y. Hu, P. Xue, and W. Yi, Fixedpoints and emergent topological phenomena in a parity-time-symmetric quantum quench, arXiv:1806.10268.

[33] See Supplemental Materials for details.[34] N. Y. Halpern, Jarzynski-like equality for the out-of-

time-ordered correlator, Phys. Rev. A 95, 012120 (2017).[35] L. Zhou, Q. Wang, H. Wang, and J. Gong, Dynami-

cal quantum phase transitions in non-Hermitian lattices,arXiv:1711.10741.

[36] J. Garrison, and E. Wright, Complex geometrical phasesfor dissipative systems, Phys. Lett. A 128, 177 (1988).

[37] S.-D. Liang, and G.-Y. Huang, Topological invarianceand global Berry phase in non-Hermitian systems, Phys.Rev. A 87, 012118 (2013).

[38] S. Lieu, Topological phases in the non-Hermitian Su-Schieffer-Heeger model, Phys. Rev. B 97, 045106 (2018).

[39] C. M. Bender, and S. Boettcher, Real spectra in non-Hermitian Hamiltonians having PT symmetry, Phys.Rev. Lett. 80, 5243-5246 (1998).

[40] C. M. Bender, D. C. Brody, and H. F. Jones, Complexextension of quantum mechanics, Phys. Rev. Lett. 89,270401 (2002).

[41] C. M. Bender, Making sense of non-Hermitian Hamilto-nians, Rep. Prog. Phys. 70, 947-1018 (2007).

7

Supplemental Materials

Experimental implementation

The main component of our setup is the cascaded interferometric network as shown in Fig. 1(a) of the maintext. For the single-photon source, we use 401.8nm continuous-wave diode laser to pump an optically nonlinear β-Barium-Borate (BBO) crystal. The polarization-degenerate photon pairs are generated by the non-collinear type-Ispontaneous parametric down conversion (SPDC). The trigger photon heralds the presence of a signal photon.

We implement the coin operator C(θ) =∑

x |x〉 〈x| ⊗ e−iθσy via two half-wave plates (HWPs). The angle of thefirst HWP is set to 0 and that of the second is θ. The shift operator S via a beam displacer (BD) whose opticalaxis is cut so that the photons in |V 〉 are directly transmitted and those in |H〉 undergo a lateral displacement into aneighboring mode [14].

To realize a non-unitary quantum walk (QW), the non-unitary dynamics is enforced by performing a partial mea-surementMe = 1w⊗

√l |−〉 〈−| in the basis |±〉 = (|H〉± |V 〉)/

√2 at each step, with 1w =

x |x〉 〈x| and l the lossparameter and is fixed to 0.36 in our experiment. The partial measurement operatorMe is realized by a sandwich-typesetup involving two HWPs at 22.5 and a partially polarizing beam splitter (PPBS). At each step, after applyingthe partial measurement, photons in the state |−〉 are reflected by the PPBS with probability l and the rest photonscontinue propagating in the quantum-walk dynamics. By choosing proper coin parameters (θf1, θ

f2), the non-unitary

QW can possess parity-time (PT ) symmetry. By changing coin parameters, the non-unitary QW can be either in thePT -symmetry-unbroken or -broken regimes.

In our experiment, the initial coin state is chosen as either a pure state such as |ψi−〉 or a mixed state such as

ρ0 = p |ψi−〉 〈ψi

−| + (1 − p) |ψi+〉 〈ψi

+|, where |ψi±〉 = (|H〉 ∓ i |V 〉)/

√2. As |ψi

−〉 is a special case of ρ0 with p = 1,we only show how to prepare the signal photon in a mixed state. After photons passing the first polarizing beamsplitter (PBS), only the horizontally polarized photons are injected into the remaining optical setup. The first HWP(H0) is set to θ = arccos(

√p)/2. Hence, we are able to prepare an arbitrary mixed state by controlling the setting

angle of H0. For the pure initial state |ψi−〉, the setting angle of H0 is taken to be 0. The second PBS splits the

photons polarization-wise into two spatial paths, where photons are prepared in either |ψi+〉 or |ψi

−〉 with HWPs (H1

or H2) and quarter-wave plates (QWPs, Q1 or Q2). The optical-path difference between two paths is longer than thecoherence length of photons. Finally, a 50 : 50 non-polarizing beam splitter (NPBS) combines photons from two pathsand then re-splits them into two paths. Photons in one path are injected into the cascaded interferometric networkwhich implements QWs. Photons in the other path are in the initial state, and are used for measurement after thephases caused by the optical difference between the two paths are compensated.

Interference-based measurements of inner products

In quasimomentum space,

G(k, t) = Tr[

ρ0(

U fk

)t]

= p〈ψi−|ψ−(t)〉+ (1 − p)〈ψi

+|ψ+(t)〉

=∑

x

e−ikxP (x, t),

where P (p, x, t) = p〈ψi−|ψ−(x, t)〉+(1−p)〈ψi

+|ψ+(x, t)〉, |ψ±(t)〉 =(

U fk

)t |ψi±〉, and |ψ±(x, t)〉 =

k eikx|ψ±(t)〉 is the

coin state on site x at time t when the initial state is |x = 0〉⊗ |ψi±〉. The trace is taken with respect to the coin state

in the k-sector.

In this section, we discuss how to experimentally measure P (p, x, t) through interference-based measurements.Without loss of generality, we write

|ψi−〉 = (α, β)T, |ψi

+〉 = (a, b)T, (S1)

|ψ−(x, t)〉 = (α′, β′)T, |ψ+(x, t)〉 = (a′, b′)T.

The polarization state of the photons after the t-th steps at x is given by the density matrix

ρt(x) = p |ψ−(x, t)〉 〈ψ−(x, t)|+ (1− p) |ψ+(x, t)〉 〈ψ+(x, t)| . (S2)

8

At the measurement stage, the first PBS combines the photons in the states ρ0 and ρt(x) and re-splits them intopaths 1 and 2 depending on their polarizations. The polarization states of the photons in paths 1 and 2 are

ρP1 = p

(

α′2 α′α∗

αα′∗ α2

)

+ (1− p)

(

a′2 a′a∗

aa′∗ a2

)

, (S3)

ρP2 = p

(

β2 ββ′∗

β′β∗ β′2

)

+ (1 − p)

(

b2 bb′∗

b′b∗ b′2

)

,

respectively.Using a QWP, a HWP and a PBS, we apply the projector |ψi

+〉 〈ψi+| on photons in path 1. The probability

P11 =1

2

[

pα′2 + (1 − p)a′2 + pα2 + (1 − p)a2]

+ Im [pα′α∗ + (1 − p)a′a∗] (S4)

can be read out from the coincidence between the detectors D1 and D0 (to record the triggering photons). Similarly,the probability P ′

11 is read out from the coincidence between D2 and D0. Thus, the probability of photons measuredin path 1 is

P1 = P11 + P ′11 = pα′2 + (1− p)a′2 + pα2 + (1− p)a2. (S5)

By changing the setting angles of the QWP and HWP, we apply the other projector |+〉 〈+| on photons in path 1.The resulting probability is

P12 =1

2

[

pα′2 + (1 − p)a′2 + pα2 + (1 − p)a2]

+Re [pα′α∗ + (1 − p)a′a∗] , (S6)

which can be read out from the coincidence between the detectors D1 and D0.Similarly, we obtain the probability of photons in path 2 by measuring in the basis state |ψi

+〉 and |+〉, i.e., P21

and P22 respectively. These can be read out from the coincidence between D4 and D0. With the probability P ′21 read

out from the coincidence between D3 and D0, we have P2 = P21 + P ′21. Hence, we can calculate P (p, x, t) from the

directly measured probabilities as

P (p, x, t) = i

(

P11 −P1

2− P21 +

P2

2

)

+

(

P12 −P1

2+ P22 −

P2

2

)

. (S7)

From P (p, x, t), we then construct the rate function and DTOPs as outlined in the main text.

Error analysis

We have four sources of systematic errors in our experimental setup: imperfections of optical elements, differencein photon loss in different optical paths, statistical noise, and decoherence in the interference-based measurements.First, the accuracy of angles of WPs is about ±0.1. Second, losses in different optical paths split by the NPBS are

estimated in an independent measurement with an accuracy of ±2%. Third, errors due to photon-counting statisticsare scaled by the square root of the number of click events. Here, total coincidence counts are about 40, 000 over acollection time of 20s. For unitary QWs, errors due to Poissonian statistics are a minor contribution compared toother systematic deviations. However, for non-unitary QWs, as non-unitarity is introduced via loss of photons, totalcoincidence counts decrease, and errors due to photon-counting statistics become significant compared to other errors.Thus in Fig. 5, error bars are derived only from photon-counting statistics. Fourth, in our experiment, the majorsource of decoherence is dephasing, caused by the optical-path difference between paths split by NPBS. We assumethat the dephasing rate η = 0.97 is a constant to simplify estimation. Under dephasing, the time-evolved coin stateis not pure anymore. We estimate the density matrix of the coin state as ε(ρc) = ηρc + (1 − η)σzρcσz , where σz isthe standard Pauli operator.Taking into account all possible combinations of systematic errors listed above, we use experimentally measured

values and estimated error ranges as input for Monte Carlo simulations. We then use analytic results as reference, andtake the largest positive (negative) deviation of numerical results from the reference as the negative (positive) errorbar. Due to decoherence, the error bars are typically asymmetric. We also note that at critical times where νm(t)undergo abrupt jumps, the error bars also become large and asymmetric, as the system dynamics become sensitive tosmall perturbations at these times.

9

Topological invariants for QW dynamics

In this section, we discuss the calculation of winding numbers for U and U defined in the main text. First, weconsider the unitary QW dynamics governed by U . Taking the Fourier transform, we derive the Floquet operator Uk

in each k-sector

Uk = n0σ0 − n1σ1 − in2σ2 − in3σ3,

n0 = cos(2k) cos θ1 cos θ2 − sin θ1 sin θ2,

n1 = 0,

n2 = cos(2k) cos θ2 sin θ1 + cos θ1 sin θ2,

n3 = − sin(2k) cos θ2.

(S8)

Topological properties of U are characterized by the winding number, defined as

ν = − 1

∫ π

−π

(

n× dn

dk

)

1

. (S9)

Winding number calculated in this way is equivalent to that from the Bloch vector of the effective Hamiltonian.In non-unitary QW dynamics governed by U , topologically inequivalent phases can be distinguished by the winding

number ν = ϕB/2π. Here, the global Berry phase ϕB is

ϕB = −i∑

µ=±

∫ π

−π

dk〈χµ| ∂

∂k |ψµ〉〈χµ|ψµ〉

, (S10)

where the right (left) eigenvector is defined as Uk|ψµ〉 = λµ|ψµ〉 (U †k |χµ〉 = λ∗µ|χµ〉).

Topological invariants defined through the global Berry phase ϕB have the advantage that they invoke the formal-ism of biorthonormal basis and characterize topological properties in both the PT -symmetry-preserving and brokenregimes.

Parity-time symmetry

U = F (γM)G, where γ = (1 − l)−1/4, has PT symmetry, where PT U(PT )−1 = U−1 with the symmetry operatorPT =

x | − x〉〈x| ⊗ σzK and K the complex conjugation. In momentum space, U can be resolved into the followingexpression

U = d0σ0 − id1σ1 − id2σ2 − id3σ3,

d0 = α(cos(2k) cos θ1 cos θ2 − sin θ1 sin θ2),

d1 = iβ,

d2 = α(cos(2k) cos θ2 sin θ1 + cos θ1 sin θ2),

d3 = − α sin(2k) cos θ2,

(S11)

where α = γ2 (1 +

√1− l), β = γ

2 (1 −√1− l). The eigenvalues of U is λ± = d0 ∓ i

1− d20, and the quasienergy

i ln(λ±). When d20 < 1 for all k, the quasienergy spectra is real, and the system is in the PT -symmetry-unbrokenregime. If d20 ≥ 1 for some k, the PT symmetry is spontaneously broken and the quasienergy can be complex.

DQPT for pure states

In this section, we discuss the simulation of topological DQPTs for pure states using QW dynamics. Due to thelattice translational symmetry, time evolutions in different k-sectors are decoupled. In each k-sector, we consider theQW dynamics initialized in the coin state |ψi

−〉, and governed by the Floquet operator U fk. We require that |ψi

−〉should be an eigenstate of the Floquet operator U i

k, and write Uk|ψf±〉 = e∓iEf

k |ψf±〉, where |ψf

±〉 are eigenstates of U fk

and ±Efk are the corresponding quasienergy. With these, the time-evolved state is

|ψ−(t)〉 = (U fk)

t|ψi−(k)〉 = eiE

fktc−|ψf

−〉+ e−iEfktc+|ψf

+〉, (S12)

10

where c− = 〈ψf−|ψi

−〉 and c+ = 〈ψf+|ψi

−〉.The Loschmidt amplitude is then

G(k, t) = 〈ψi−|ψ−(t)〉 = eiE

fkt|c−|2 + e−iEf

kt|c+|2 := |G(k, t)|eiφ(k,t). (S13)

The PGP is defined through φG(k, t) = φ(k, t) − φdyn(k, t), where φdyn(k, t) = Efkt(|c−|2 − |c+|2) is the dynamical

phase. It is straightforward to show that when c+(km) = 0 or c−(km) = 0, PGP vanishes at all times. Further, onecan show that at these km, the corresponding density matrices do not evolve in time. We therefore identify thesemomenta as fixed points of dynamics.DQPTs are caused by dynamic Fisher zeros, where the Loschmidt amplitude G(kc, tc) vanishes at the time tc. Here

kc is the critical momentum, determined by |c−(kc)|2 = |c+(kc)|2. It follows that DQPTs occur at tc = (2n − 1)t0,where t0 = π/2Ef

kcis the critical time scale. Notably, as we will show later in the Supplemental Material, when the

system is quenched between different topological phases, fixed points with c+ = 0 and those with c− = 0 alwaysemerge in pairs. As |c+| − |c−| are continuous functions of k, there must be at least one critical momentum satisfying|c+(kc)| = |c−(kc)| inbetween two fixed points of different kinds. DQPTs necessarily exist in this case.We now consider the more concrete case discussed in the main text, where the initial and final Floquet operators are

parameterized by (θi1 = π/4, θi2 = −π/2) and (θf1 = −π/2, θf2 ∈ (0, π/2)), respectively. In this case, eigenstates of U i

are spatially localized on the lattice, and |ψi−〉 = (|H〉+ i|V 〉)/

√2 for any k. It follows that |c−|2 = 1

2 (1 + cos 2k) and|c+|2 = 1

2 (1−cos 2k). There are four fixed points at k1,2,3,4 = −π,−π/2, 0, π/2, and there are four critical momentaat kc = ±π/4,±3π/4. Importantly, due to symmetries of the unitary Floquet operator U f

k, Efkc

are degenerate at

different kc. Hence there is only one critical time scale t0 = π/(2Efkc).

We define the rate function

g(t) = − 1

π

∫ π

−π

dk ln |G(k, t)|, (S14)

where N is the system size. As illustrated in Fig. S1, DQPTs occur when g(t) becomes non-analytical at criticaltimes.DQPTs are characterized by DTOPs, defined in terms of the PGP as

νm(t) =1

∫ km+1

km

dk∂φG(k, t)

∂k. (S15)

As the PGP φG(k, t) vanishes km, νm(t) characterizes the S1 → S1 mapping from the momentum submanifold

between km and km+1 to eiφGk (t). νm(t) is therefore quantized, and can only change values at critical times, when

G(kc, tc) = 0 and φG(kc, tc) becomes ill-defined (see Fig. S1).Finally, we emphasize that the rate function constructed in Eq. (S14) corresponds to that of a system initialized

in the state∏

k∈1BZ |ψi−〉. We are therefore simulating quench dynamics of a many-body system using single-photon

QW dynamics.

DQPT for mixed states

We now discuss the simulation of DQPTs for mixed states, where the Loschmidt amplitude in each k-sector isdefined as [25]

G(k, t) := Tr[

ρ0(

U fk

)t]

. (S16)

For an initial density matrix ρ0 = p |ψi−〉 〈ψi

−|+ (1 − p) |ψi+〉 〈ψi

+|, we have

G(k, t) = p(eiEfkt|c−|2 + e−iEf

kt|c+|2) + (1− p)(e−iEfkt|c−|2 + eiE

fkt|c+|2) := |G(k, t)|eiφ(k,t). (S17)

It is apparent that the critical momenta and critical times satisfy the same relations as in the pure-state case, with|c−(kc)| = |c+(kc) and DQPTs occurring at tc = (2n − 1)t0. Also similar to the pure-state case, there is only onecritical-time scale t0 = π/(2Ef

kc).

DTOPs in this case are defined as νm(t) = 12π

∫ km+1

kmdk ∂φG(k,t)

∂k . However, as φG(km, t) 6= φG(kn, t) for m 6= n,

φG(k, t) is no longer periodic in the interval k ∈ (km, km+1), and νm(t) in the mixed-state case are not quantized.In Fig. S2, we show time evolutions of the rate function and DTOP for a typical unitary dynamics of mixed state.Whereas non-analyticities still occur in g(t) periodically and νm(t) still demonstrate abrupt jumps at critical times,νm(t) are not quantized.

11

0

0.8

1.6

0 4 8 12 16 20

-3

0

3

FIG. S1. Rate function (a) and DTOPs (b) for unitary dynamics of pure states. The coin parameters are the same as those inFig. 2(a) of the main text.

0

0.8

1.6

0 4 8 12 16 20

-2

0

2

0 4 8 12 16 20

-2

0

2

0

1.5

3

FIG. S2. Rate function (upper row) and DTOPs (lower row) for unitary dynamics of mixed states. The coin parameters arethe same as those in Fig. 3 of the main text.

DQPTs for non-unitary dynamics

We now discuss the simulation of DQPTs in PT -symmetric non-unitary dynamics of pure states. As we will provelater, DQPTs occur when the system is quenched between different topological phases in the PT -unbroken regime.

We consider the Loschmidt amplitude in each k-sector

G(k, t) := 〈ψi−(k)|ψ−(t)〉 = b+c+e

−iEfkt + b−c−e

iEfkt, (S18)

where b± = 〈ψi−|ψf

±〉, c±(k) = 〈χf±|ψi

−〉. For U f in the PT -unbroken regime, Efk is real for all k. The time-evolved

12

-3

0

3

0 4 8 12

-1.4

-0.7

0

0

0.4

0.8

FIG. S3. Rate function (a) and DTOPs (b) for non-unitary dynamics in the PT -symmetry-unbroken regime. (c) Rate functionfor non-unitary dynamics in the PT -symmetry-broken regime. The coin parameters are the same as those in Fig. 5 of the maintext.

state is then

|ψ−(t)〉 = (U fk)

t |ψi−〉 = c+e

−iEfkt |ψf

+〉+ c−eiEf

kt |ψf−〉 , (S19)

According to the expression of G(k, t), DQPTs occur periodically at t0 = (n + 12 )π/E

fkc

(n ∈ N) when Efkc

isreal. Here kc satisfies |b+(kc)c+(kc)| = |b−(kc)c−(kc)|. When the system is quenched between different topologicalphases in the PT -symmetry-unbroken regime, fixed points with c+ = 0 and those with c− = 0 always emerge inpairs. As |b+c+| − |b−c−| are continuous functions of k, there must be at least one critical momentum kc satisfy-ing |b+(kc)c+(kc)| = |b−(kc)c−(kc)| inbetween two fixed points of different kinds. We thus conclude that DQPTsnecessarily occur in this case. At these critical times, the rate function, defined as

g(t) = − 1

π

∫ π

−π

dk ln |G(k, t)| (S20)

becomes non-analytic. A typical behavior of g(t) is shown in Fig. S3(a).

Similar to the unitary case, we define DTOPs between two adjacent fixed points as

νm(t) =1

∫ km+1

km

∂φGk (t)

∂kdk. (S21)

Here, φGk (t) = φk(t) − φdynk (t), where φk(t) is defined through G(k, t) = |G(k, t)|eiφk(t) and the dynamic phase

φdynk (t) = (b−c− − b+c+)Efkt. At critical points, G(k, t) vanishes, which leads to abrupt jumps in νm(t). Further, as

φGk (t) vanishes at fixed points, νm characterizes the S1 → S1 mapping from the momentum submanifold between km

and km+1 to eiφGk (t), and is therefore quantized.

Importantly, in non-unitary dynamics, due to the breaking of time-reversal symmetry, fixed points km are no longerlocated at high-symmetry points. As a consequence, Ef

kccan take two different values for kc inbetween different fixed

points. This gives rise to two different critical time scales [see Fig. S3(b)].

In contrast, when the final U f is in the PT -symmetry-broken regime, DQPTs are typically absent. As illustratedin Fig. S3(c), the rate function is smooth in time, and DTOPs are not well defined due to the absence of fixed points.

13

Fixed points and topology

In this section, we show that the existence and number of fixed points are intimately connected with windingnumbers νi,f of the Floquet operators U i,f respectively, when both Floquet operators belong to the PT -symmetry-unbroken regime with completely real eigenspectra. We focus on the non-unitary case, as derivations in the unitarycase with U are essentially the same when taking the loss parameter l = 0.Starting from Eq. (S11), and focus on the PT -symmetry-unbroken regime, where d20 < 1 for ∀k, we derive the right

and left eigenvectors of Uk

|ψ±〉 = V † 1√2 cos 2Ω

(±e±iΩ, e+iϑe∓iΩ)T , (S22)

〈χ±| =1√

2 cos 2Ω(±e±iΩ, e−iϑe∓iΩ)V. (S23)

Here V = eiπ/22

σy , ϑ and Ω are respectively defined through −d3 + id2 = deiϑ and sin 2Ω = −id1/d, withd2 = d22 + d23. As d1/d ∈ (0, 1), we set 2Ω ∈ (0, π/2) and cos 2Ω > 0. Notice that the orthonormal conditions(

〈χ±|ψ±〉 = 1, 〈χ±|ψ∓〉 = 0)

are always satisfied in this region.

We then have

d

dk|ψ±〉 = V † 1

(2 cos 2Ω)3/2(i2e∓iΩΩ′, ieiϑe∓iΩ(2 cos 2Ω)ϑ′ ∓ i2eiϑe±iΩΩ′)T , (S24)

where Ω′ = dΩ/dk and ϑ′ = dϑ/dk. It is then straightforward to derive the Berry connection A± =−i 〈χ±| d

dk |ψ±〉 /〈χ±|ψ±〉 = 12ϑ

′ ± i2ϑ

′ tan 2Ω, and hence A+ +A− = ϑ′.

The winding number of U is defined as ν = ϕB/2π, where the global Berry phase ϕB = ϕZ+ + ϕZ−, with

ϕZ± =

dkA± = −i∮

dk〈χ±| d

dk |ψ±〉〈χ±|ψ±〉

. (S25)

We then have ν =∮

dkϑ′. Since ϑ = arg(−d3 + id2) = arg(−n3 + in2) = arg(i(n2 + in3)), U has the same windingnumber as U . Here the vector n is defined in Eq. (S8).It follows that

c± = 〈χf±|ψi

−〉 =∓e−i(Ωi∓Ωf) + eiϑ

0

ei(Ωi∓Ωf)

2√cos 2Ωi cos 2Ωf

, (S26)

where ϑ0 = ϑi − ϑf. Consider a unit vector n0 on the y-z plane whose polar angle is given by ϑ0. As k ∈ 1BZ, thenumber of times n

0 winds around the x-axis is therefore∮

dk∂φ0/∂k = νf − νi. From Eq. (S26), the condition forc± = 0 is φ0 = 2(Ωf − Ωi) (mod 2π) and φ0 = π − 2(Ωf + Ωi) (mod 2π), respectively. Therefore, the number of fixedpoints with c+ = 0 or c− = 0 should be at least

∣νf − νi∣

∣ each.

Finally, we note that the existence of fixed points are not guaranteed when either U i or U f is in the PT -symmetry-broken regime. In the absence of fixed points, topological DQPTs are absent and DTOPs cannot be defined.