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

Deterministic Equations for Feedback Control of Open Quantum Systems

Alberto J. B. Rosal Department of Physics and Astronomy, University of Rochester, Rochester, New York 14627, USA University of Rochester Center for Coherence and Quantum Science, Rochester, New York 14627, USA    Patrick P. Potts Department of Physics and Swiss Nanoscience Institute, University of Basel, Klingelbergstrasse 82 CH-4056, Switzerland    Gabriel T. Landi Department of Physics and Astronomy, University of Rochester, Rochester, New York 14627, USA University of Rochester Center for Coherence and Quantum Science, Rochester, New York 14627, USA
(September 22, 2025)
Abstract

Feedback control in open quantum dynamics is crucial for the advancement of various coherent platforms. For a proper theoretical description, however, most feedback schemes rely on stochastic trajectories, which require significant statistical sampling and lack analytical insight. Currently, only a handful of deterministic feedback master equations exist in the literature. In this letter we derive a set of deterministic equations for describing feedback schemes based on generic causal signals. Our formulation is phrased in terms of sequentially applied quantum instruments, and is therefore extremely general, recovering various known results in the literature as particular cases. We then specialize this result to the case of quantum jumps and derive a new deterministic equation that allows for feedback based on the channel of the last jump, as well as the time since the last jump occurred. The strength of this formalism is illustrated with a detailed study of population inversion of a qubit using coherent drive and feedback, where our formalism allows all calculations to be performed analytically.

Introduction.— Feedback control finds numerous applications across practically all quantum coherent platforms, in tasks such as cooling protocols Hopkins et al. (2003); D’Urso et al. (2003); Steck et al. (2006); Bushev et al. (2006); Jacobs et al. (2015); Frimmer et al. (2016); Guo et al. (2019); Buffoni et al. (2019); Manikandan and Qvarfort (2023); De Sousa et al. (2025), quantum error correction Sarovar et al. (2004); Marcos et al. (2020), entanglement generation Ristè et al. (2013), thermodynamics Pekola (2015); Potts and Samuelsson (2018); Barker et al. (2022), transport van der Wiel et al. (2002), control Vijay et al. (2012); Ristè et al. (2012); Campagne-Ibarcq et al. (2013), and quantum batteries Mitchison et al. (2021). Notably, feedback has been instrumental in the experimental realization of Maxwell’s demon in the quantum regime Vidrighin et al. (2016); Naghiloo et al. (2018); Ribezzi-Crivellari and Ritort (2019), prompting the development of generalized formulations of the second law of thermodynamics to incorporate feedback-based measurement processes Sagawa and Ueda (2010, 2012); Funo et al. (2013). These developments highlight the fundamental importance of quantum feedback in both theory and experiment, motivating further research into its underlying principles and applications.

Broadly speaking, feedback refers to the process of dynamically controlling a system based on previous detection outcomes (see Fig. 1(a)). In continuously monitored open quantum systems this can be readily implemented using stochastic differential equations, both in the quantum jump and the quantum diffusion (e.g. homodyne/heterodyne) unravellings Belavkin (1983, 1992a, 1987, 1992b); Wiseman (1994a); Wiseman and Milburn (1993, 2010); Korotkov (2001); Jacobs (2014); Zhang et al. (2017). However, this approach typically requires computationally expensive statistical sampling and often obscures physical insights, as it does not yield analytical expressions for key system properties, such as the steady state (when it exists). For this reason, deterministic feedback equations have long been sought for Sarovar et al. (2004); Tilloy (2024); Annby-Andersson et al. (2022); Wiseman and Milburn (1993); Wiseman (1994b); Wiseman and Milburn (2010); Kewming et al. (2024).

Refer to caption
Figure 1: (a) Schematics of a feedback protocol. Based on the detection outcomes, the feedback can modify the system’s dynamics (e.g., the Hamiltonian HH), the parameters of the environment (e.g., the temperature of a thermal bath), or the measurement scheme itself. (b) Summary of results. Eq. (2) describes a feedback protocol applicable to general measurement scenarios with a causal signal yn=fn(xn,yn1)y_{n}=f_{n}(x_{n},y_{n-1}). Eqs. (8) and (9) describe a jump-time-dependent feedback. In Sup , we show how previously known results from Annby-Andersson et al. (2022); Wiseman and Milburn (1993); Wiseman (1994b); Kewming et al. (2024) can be recovered as special cases of Eq. (2).

Let x1:n=(x1,x2,,xn)x_{1:n}=(x_{1},x_{2},\dots,x_{n}) denote some stochastic data obtained from an experiment. Typically, feedback schemes do not make use of the entire dataset. Instead, they rely on a compressed representation of the data, which we henceforth refer to as the signal. This signal is defined as a function of the data, yn=yn(x1:n)y_{n}=y_{n}(x_{1:n}). Hence, to classify different types of feedback schemes, it is helpful to clarify (i) what signal is being used and (ii) what the corresponding feedback actions are. For a given choice of signal, one can then construct various types of feedback actions. An illustrative example is homodyne detection-based feedback Wiseman and Milburn (1993): the data xnx_{n} is the homodyne current; the signal is the instantaneous outcome (yn=xny_{n}=x_{n}); and the action is a drive in the system Hamiltonian H(yn)H(y_{n}) that depends on the magnitude of the signal. Ref. Annby-Andersson et al. (2022) recently extended this to low-pass (LP) filters of the form yn=i=1nδtγeγδt(ni)xiy_{n}=\sum_{i=1}^{n}\delta t\gamma e^{-\gamma\delta t(n-i)}x_{i}, for a bandwidth γ\gamma. The particular case in which all detection weights are equal to one corresponds to a signal that depends on the entire history (often called the “integrated charge”), yn=j=1nxjy_{n}=\sum_{j=1}^{n}x_{j}, and was studied in Kewming et al. (2024). In these cases, the authors derive deterministic equations that describe the feedback protocol based on the corresponding signals.

Refer to caption
Figure 2: Diagram of the conditional evolution under feedback. The system’s evolution depends on the detection outcomes, which condition the feedback actions. The instruments x\mathcal{M}_{x} encompass both the dynamical evolution and the measurement process at each step.

To further illustrate this point, Ref. Wiseman (1994a) derived a famous deterministic jump equation that implements the following protocol: if a quantum jump is observed, then we apply a quantum channel. The channel is now an abrupt action, that instantaneously modifies the system. Extending this to a smooth time-dependent action, such as turning on a Hamiltonian drive, is one of our present goals. Furthermore, for more general jump-based feedback — especially those involving time-dependent actions — no deterministic equations are typically available. For instance, in the case of quantum jumps, the output data xnx_{n} at each time step tn=nδtt_{n}=n\delta t is not a continuous, noisy current, but rather a discrete sequence of “clicks”, being either xn=0x_{n}=0 (no jump) or xn=k=1,2,x_{n}=k=1,2,\ldots for a jump in channel kk. A natural kind of feedback for quantum jumps would be: if a quantum jump is observed, modify the Hamiltonian for some fixed amount of time. This protocol is, in fact, routinely used in quantum dot experiments which monitor the jumps using a quantum point contact Hofmann et al. (2016); Ye et al. (2025). However, this cannot be described using only feedback based on the latest outcome (yn=xny_{n}=x_{n}) because most of the time we would have yn=0y_{n}=0. Consequently, feedback schemes that depend on jump timing generally lack a corresponding deterministic evolution equation.

In this letter we consider the most general type of continuously monitored open quantum system, formulated in terms of quantum instruments Milz and Modi (2021). We first show (Result 1) that it is possible to derive a deterministic feedback equation for any signal that can be written as yn=fn(xn,yn1)y_{n}=f_{n}(x_{n},y_{n-1}), where fnf_{n} is an update rule that depends only on the previous signal and the present outcome. To illustrate the generality and power of this result, in the Supplemental Material Sup we use it to derive various other deterministic feedback equations Annby-Andersson et al. (2022); Wiseman and Milburn (1993); Wiseman (1994b); Kewming et al. (2024) in the literature (see Fig. 1(b)). Next, we specialize this to the case of quantum jumps, and derive a deterministic equation (Result 2) that allows for feedback based on: (i) what was the channel of the last jump [Eq. (6)] and (ii) how long has it been since the last jump occurred [Eq. (7)]. The combination of these two features allow for non-trivial feedback strategies that are experimentally relevant Hofmann et al. (2016); Ye et al. (2025). For example, one can turn on coherent drives based on whether a tunneling event into or out of a quantum dot has occurred, or depending on whether a photon has been emitted or absorbed by an atom. More sophisticated strategies are also possible, such as turning on a drive only after a certain time has elapsed since the last detected jump.

Sequential measurements.- Continuously measured quantum systems can be described by the sequential application of super-operators {x}\{\mathcal{M}_{x}\}, corresponding to trace non-increasing maps such that xx\sum_{x}\mathcal{M}_{x} is a quantum channel Milz and Modi (2021). These maps are called instruments, and they describe the joint transition of the quantum state induced by the dynamical evolution and measurement process, whereas xx represents a possible measurement outcome.

For instance, one may assume that the measurements are described by the set {Vx}\{V_{x}\} of Kraus operators applied at regular intervals tj=jδtt_{j}=j\delta t, where j=0,1,2,j=0,1,2,\dots and δt>0\delta t>0. In between, the system may evolve from tt to t+δtt+\delta t with some dynamical map ρ(t+δt)=Λt+δt,tρ(t)\rho(t+\delta t)=\Lambda_{t+\delta t,t}\rho(t). Given an initial state ρ0\rho_{0}, after nn steps one obtains a sequence of outcomes x1:n=(x1,x2,,xn)x_{1:n}=(x_{1},x_{2},\dots,x_{n}), and the corresponding conditional state of the system is denoted by ρ1:n\rho_{1:n}. In this case, the instruments are defined as the maps xρ=Vx[Λt+δt,t(ρ)]Vx\mathcal{M}_{x}\rho=V_{x}[\Lambda_{t+\delta t,t}(\rho)]V_{x}^{\dagger}, and according to the measurement postulate, the probability of the next outcome xn+1x_{n+1} given the previous dataset x1:nx_{1:n} is given by P(xn+1|x1:n)=Tr[xn+1(ρx1:n)]P(x_{n+1}|x_{1:n})=\mathrm{Tr}[\mathcal{M}_{x_{n+1}}(\rho_{x_{1:n}})], and the post-detection state becomes ρx1:n+1=xn+1(ρx1:n)P(xn+1|x1:n)\rho_{x_{1:n+1}}=\frac{\mathcal{M}_{x_{n+1}}(\rho_{x_{1:n}})}{P(x_{n+1}|x_{1:n})}.

A feedback mechanism is a dynamic adjustment of the instruments x\mathcal{M}_{x} at each step, based on prior outcomes. The most general mechanism could depend on the entire detection history, and would therefore give rise to instruments of the form xn(x1:n1)\mathcal{M}_{x_{n}}(x_{1:n-1}), where the variable in the subscript defines the possible random outcomes and the variable in parenthesis defines the past information that can be used to alter the instrument. In practice, however, one seldom uses the entire data record x1:nx_{1:n}, but only some summary statistic. Here we implement this by defining a signal yny_{n} which evolves according to the update rule yn=fn(xn,yn1)y_{n}=f_{n}(x_{n},y_{n-1}), for some function fnf_{n}. This defines a general type of causal signal. But it is worth noting that this is still entirely general. For instance, if we choose an update rule of the form yn=append(yn1,xn)y_{n}=\text{append}(y_{n-1},x_{n}), then the signal becomes the entire dataset. But, of course, in this case the memory required grows with each step. In practice, however, one is often interested in signals with finite memory. Feedback is implemented by assuming that, in each step, the instruments can depend on the signal at that instant (see Fig. 2). The stochastic dynamics therefore reads

P(xn+1|x1:n)\displaystyle P(x_{n+1}|x_{1:n}) =Tr[xn+1(yn)ρx1:n],\displaystyle=\mathrm{Tr}[\mathcal{M}_{x_{n+1}}(y_{n})\rho_{x_{1:n}}], (1)
ρx1:n+1\displaystyle\rho_{x_{1:n+1}} =xn+1(yn)ρx1:nP(xn+1|x1:n).\displaystyle=\frac{\mathcal{M}_{x_{n+1}}(y_{n})\rho_{x_{1:n}}}{P(x_{n+1}|x_{1:n})}.

Our goal is to develop a deterministic equation for this general feedback protocol. To do that we define the signal-resolved state Kewming et al. (2024); Annby-Andersson et al. (2022) (also called Hybrid Quantum-Classical State Tilloy (2024); Layton et al. (2024); Diósi (2023); Oppenheim et al. (2023)) as ϱn(y)=E[ρ1:nδy,yn]\varrho_{n}(y)=E[\rho_{1:n}\delta_{y,y_{n}}], where E[]E[\cdot] denotes the average over trajectories, and δy,yn\delta_{y,y_{n}} is the Kronecker delta. The trace of this state, Tr[ϱn(y)]=P(yn=y)\mathrm{Tr}[\varrho_{n}(y)]=P(y_{n}=y), gives the distribution of the stochastic signal at step nn, while the marginalization over all possible signals yields the unconditional state of the system, yϱn(y)=ρ¯n\sum_{y}\varrho_{n}(y)=\bar{\rho}_{n}. Notice that ρx1:n\rho_{x_{1:n}} is a stochastic (conditional) state, but ϱn(y)\varrho_{n}(y) is deterministic. As our first result, we show in Sup that ϱn(y)\varrho_{n}(y) evolves according to a deterministic equation.

Result 1.

If an instrument {x(y)}\{\mathcal{M}_{x}(y)\} depends only on a causal signal yn=fn(yn,yn1)y_{n}=f_{n}(y_{n},y_{n-1}), the corresponding signal-resolved state ϱn(y)\varrho_{n}(y) evolves according to the deterministic equation

ϱn+1(y)=x,yδy,fn(x,y)x(y)ϱn(y),\varrho_{n+1}(y)=\sum_{x^{\prime},y^{\prime}}\delta_{y,f_{n}(x^{\prime},y^{\prime})}\mathcal{M}_{x^{\prime}}(y^{\prime})\varrho_{n}(y^{\prime})\leavevmode\nobreak\ , (2)

where the sum runs over all possible outcomes xx^{\prime} and all possible signal values yy^{\prime}.

The particular case yn=xny_{n}=x_{n} is treated separately in Sup , section S.V.B. The intuition behind (2) is that because ϱn(y)\varrho_{n}(y) is resolved in yy, the state at step n+1n+1 should be just a sum over all possible trajectories consistent with the constraint yn=fn(xn,yn1)y_{n}=f_{n}(x_{n},y_{n-1}). Solving (2) yields not only the unconditional state ρ¯n\bar{\rho}_{n}, but also the probability distribution P(yn=y)P(y_{n}=y) of the stochastic signal yny_{n}. Moreover, it can be used to study either transients or the steady-state, obtained by setting ϱn+1(y)=ϱn(y)=ϱss(y)\varrho_{n+1}(y)=\varrho_{n}(y)=\varrho_{\rm ss}(y). To illustrate the generality of this result, in Sup we use it to derive various other deterministic equations in the literature Annby-Andersson et al. (2022); Wiseman and Milburn (1993); Wiseman (1994b); Kewming et al. (2024).

Quantum jumps.- We now specialize Result 1 to the case of quantum jumps, for which deterministic feedback equations are much more scarce. We assume the system evolves according to a quantum master equation of the form

tρt=ρ=i[H,ρt]+kLkρtLk12{LkLk,ρt},\partial_{t}\rho_{t}=\mathcal{L}\rho=-i\big{[}H,\rho_{t}\big{]}+\sum_{k}L_{k}\rho_{t}L_{k}^{\dagger}-\frac{1}{2}\big{\{}L_{k}^{\dagger}L_{k},\rho_{t}\big{\}}, (3)

with Hamiltonian HH and jump operators LkL_{k}. Our goal is to augment this with the possibility that H(y)H(y) and {Lk(y)}\{L_{k}(y)\} both depend on the recorded signal yy.

Let Σ\Sigma denote the set of jumps that are assumed to be monitored. At the stochastic level, the outcome at each step δt\delta t is either xn=0x_{n}=0 (no jump) or xn=kΣx_{n}=k\in\Sigma when there is a jump of type kk. The corresponding instruments read Landi et al. (2024)

0(y)ρ\displaystyle\mathcal{M}_{0}(y)\rho =(1+0(y)δt)ρ,\displaystyle=(1+\mathcal{L}_{0}(y)\delta t)\rho, (4)
k(y)ρ\displaystyle\mathcal{M}_{k}(y)\rho =δt𝒥k(y)ρ,\displaystyle=\delta t\mathcal{J}_{k}(y)\rho, (5)

where 𝒥k(y)ρ=Lk(y)ρLk(y)\mathcal{J}_{k}(y)\rho=L_{k}(y)\rho L_{k}^{\dagger}(y) and 0(y)=(y)kΣ𝒥k(y)\mathcal{L}_{0}(y)=\mathcal{L}(y)-\sum_{k\in\Sigma}\mathcal{J}_{k}(y).

We consider two types of signals. First, one that records the channel kΣk\in\Sigma of the last jump. This can be constructed with a filter function fnf_{n} of the form

kn=xn+kn1δxn,0.k_{n}=x_{n}+k_{n-1}\delta_{x_{n},0}\leavevmode\nobreak\ . (6)

Second, a signal that counts how much time has elapsed since the last jump, which is built using

τn=δxn,0(τn1+δt),\tau_{n}=\delta_{x_{n},0}(\tau_{n-1}+\delta t)\leavevmode\nobreak\ , (7)

We refer to them as the jump signal and the counting signal, respectively. The total signal is yn=(kn,τn)y_{n}=(k_{n},\tau_{n}) and we consider feedback mechanisms that may depend on both. In this case, the instruments are defined by Eqs. (4) and (5), and the feedback action corresponds to allowing both the system’s Hamiltonian HH and the jump operators LkL_{k} at step nn to depend on the last jump kn1k_{n-1} and the elapsed time τn1\tau_{n-1} since its detection. Starting from Result 1, we show in Sup the following result:

Result 2.

Let ϱt(k,τ)\varrho_{t}(k,\tau) denote the state of the system resolved in both the jump and counting signals (6), (7). In the limit δt0\delta t\to 0 this evolves according to the deterministic equations

ϱt(k,0)\displaystyle\varrho_{t}(k,0) =\displaystyle= kΣ0t𝑑τ𝒥k(k,τ)ϱt(k,τ),\displaystyle\sum_{k^{\prime}\in\Sigma}\int\limits_{0}^{t}\leavevmode\nobreak\ d\tau^{\prime}\mathcal{J}_{k}(k^{\prime},\tau^{\prime})\varrho_{t}(k^{\prime},\tau^{\prime}), (8)
ϱt(k,τ)\displaystyle\varrho_{t}(k,\tau) =\displaystyle= G(k,τ)ϱtτ(k,0),\displaystyle G(k,\tau)\varrho_{t-\tau}(k,0), (9)

where G(k,τ)𝒯[e0τ𝑑s0(k,s)]G(k,\tau)\equiv\mathcal{T}\left[e^{\int_{0}^{\tau}ds\mathcal{L}_{0}(k,s)}\right] is the propagator of the no-jump evolution and 𝒯[]\mathcal{T}[\cdot] is the time-ordering operator.

In this limit δt0\delta t\to 0 the signals become continuous-time stochastic processes ktk_{t} and τt\tau_{t}. They specify that, at time tt, the last jump that happened was in channel ktk_{t}, and occurred at time tτtt-\tau_{t}. The quantity Tr[ϱt(k,τ)]\mathrm{Tr}[\varrho_{t}(k,\tau)] is the joint distribution of (kt,τt)(k_{t},\tau_{t}), and from (7) it follows that τt[0,t]\tau_{t}\in[0,t]. The resolved state is normalized as kΣ0tTr[ϱt(k,τ)]𝑑τ=1\sum_{k\in\Sigma}\int_{0}^{t}\mathrm{Tr}[\varrho_{t}(k,\tau)]d\tau=1, and the unconditional state is recovered as

ρ¯t=kΣ0t𝑑τϱt(k,τ).\bar{\rho}_{t}=\sum_{k\in\Sigma}\int_{0}^{t}d\tau\varrho_{t}(k,\tau)\leavevmode\nobreak\ . (10)

Result 2 cannot, in general, be written as a differential equation. Notwithstanding, it is still amenable to analytical treatment — as detailed in Sup — especially if one is interested in the steady-state. In contrast, let us now consider the special case where the feedback is independent of the counting signal (7), but might still depend on the jump signal (6). We show in Sup that the marginal state ϱt(k)=0t𝑑τϱt(k,τ)\varrho_{t}(k)=\int_{0}^{t}d\tau\varrho_{t}(k,\tau) will evolve as

tϱt(k)\displaystyle\partial_{t}\varrho_{t}(k) =\displaystyle= i[H(k),ϱt(k)]12kΣ{Lk(k)Lk(k),ϱt(k)}\displaystyle-i\big{[}H(k),\varrho_{t}(k)\big{]}-\frac{1}{2}\sum_{k^{\prime}\in\Sigma}\big{\{}L_{k^{\prime}}^{\dagger}(k)L_{k^{\prime}}(k),\varrho_{t}(k)\big{\}} (11)
+kΣLk(k)ϱt(k)Lk(k).\displaystyle+\sum_{k^{\prime}\in\Sigma}\ L_{k}(k^{\prime})\varrho_{t}(k^{\prime})L_{k}^{\dagger}(k^{\prime}).

This is now simply a system of coupled master equations, one for each possible jump channel kΣk\in\Sigma. A similar equation was previously derived in Blanchard and Jadczyk (1995) in a different context.

Refer to caption
Figure 3: (a–d) Inversion of the population of a two-level system under time-dependent feedback control based on quantum jump detections. (a) Diagram of the feedback action: upon detection of a decay event, an external drive is applied to return the system to the excited state. This feedback includes a possible time delay τ0\tau_{0} before the drive is turned on, and τ1\tau_{1} is the drive duration. (b) Population of the excited state in the ideal case without any feedback delay (τ0=0\tau_{0}=0), and considering the optimal drive duration τ1opt\tau_{1}^{\text{opt}} (Eq. (15)). The no-feedback line represents the qubit coupled with a thermal bath without feedback. (c) Effect of feedback delay on the excited state population in the low-temperature limit (N¯0\bar{N}\rightarrow 0), for different values of the delay τ0\tau_{0}. (d) Maximum allowable feedback delay in the low-temperature limit and the optimal drive duration for different feedback regimes.

Returning to the general Result 2, the situation simplifies when the feedback is applied only on the no-jump part of the dynamics (e.g. in the Hamiltonian). In this case 𝒥k(k,τ)=𝒥k\mathcal{J}_{k}(k^{\prime},\tau^{\prime})=\mathcal{J}_{k} and Eqs. (8)- (10) imply that

ϱt(k,τ)=G(k,τ)𝒥kρ¯tτ.\varrho_{t}(k,\tau)=G(k,\tau)\mathcal{J}_{k}\bar{\rho}_{t-\tau}. (12)

The signal-resolved state can hence be written in terms of the unconditional state ρ¯t\bar{\rho}_{t} which, in turn, satisfies the closed integral equation

ρ¯t=kΣ0t𝑑τG(k,τ)𝒥kρ¯tτ.\bar{\rho}_{t}=\sum_{k\in\Sigma}\int_{0}^{t}d\tau G(k,\tau)\mathcal{J}_{k}\bar{\rho}_{t-\tau}\leavevmode\nobreak\ . (13)

The steady-state, if it exists, is obtained by setting tt\to\infty and will therefore be the solution of the algebraic equation

ρ¯ss=Ωρ¯ss,Ω=kΣ0𝑑τG(k,τ)𝒥k.\bar{\rho}_{\rm ss}=\Omega\bar{\rho}_{\rm ss},\qquad\Omega=\sum_{k\in\Sigma}\int_{0}^{\infty}d\tau G(k,\tau)\mathcal{J}_{k}. (14)

Combining this with Eq. (12) yields the steady-state signal-resolved state. In turn, this can be used to calculate various time-related properties, such as waiting-time distributions, the average time between jumps, etc.

Example.– To illustrate our results, consider a coherently driven qubit coupled to a thermal reservoir with temperature TT and energy ω\omega under the action of feedback. Let |e\ket{e} (excited) and |g\ket{g} (ground) denote the computational basis. We assume a master equation of the form (3) with jump operators L=γ(N¯+1)|ge|L_{-}=\sqrt{\gamma(\bar{N}+1)}\leavevmode\nobreak\ |g\rangle\langle e| and L+=γN¯|eg|L_{+}=\sqrt{\gamma\bar{N}}\leavevmode\nobreak\ |e\rangle\langle g| describing, respectively, the emission and absorption of a quanta. Here N¯=(exp(ω/T)1)1\bar{N}=(\text{exp}(\omega/T)-1)^{-1} is the Bose-Einstein distribution and γ>0\gamma>0 is the coupling strength. The 3 possible outcomes, at each dtdt, are therefore xn=1,+1,0x_{n}=-1,+1,0 (emission, absorption, no-jump). We consider a drive with intensity λ\lambda and frequency ωd\omega_{d}, working in a frame rotating at ωd\omega_{d} and under the rotating wave approximation. The Hamiltonian when the driving is turned off/on are, respectively, Hoff=Δ2σzH_{\text{off}}=-\frac{\Delta}{2}\sigma_{z} and Hon=Δ2σz+λσxH_{\text{on}}=-\frac{\Delta}{2}\sigma_{z}+\lambda\sigma_{x}, where Δ=ωωd\Delta=\omega-\omega_{d}. Our goal is to invert the populations of |e\ket{e} and |g\ket{g} (achieving Pe>PgP_{e}>P_{g}) using the quantum jump detection record and feedback.

The protocol is defined as follows: If xn=+1x_{n}=+1 (thermal absorption) the system’s Hamiltonian is set to H=HoffH=H_{\text{off}}. If xn=1x_{n}=-1 (thermal emission), and a time τ0\tau_{0} elapses without any subsequent absorption, we turn the external drive (H=HonH=H_{\text{on}}) for a duration τ1\tau_{1}. The idea is that, upon detecting an emission and observing no absorption for a time τ0\tau_{0}, the external drive acts on the system for a time τ1\tau_{1}, inducing Rabi oscillations that transfer the population from |g\ket{g} to |e\ket{e} (see Fig. 3(a)). Whenever the drive is active and an absorption event is detected, the drive is subsequently turned off.

This feedback protocol affects only the system’s Hamiltonian; hence, the steady state can be obtained from Eq. (14). The analytical formula for the population of the excited state, Pee|ρss|eP_{e}\equiv\left<e|\rho_{ss}|e\right>, is provided Sup for the case Δ=0\Delta=0, as a function of the bath parameters N¯\bar{N} and γ\gamma, as well as the feedback parameters λ\lambda, τ0\tau_{0}, and τ1\tau_{1}. By maximizing PeP_{e} with respect to τ1\tau_{1}, one obtains the optimal duration τ1opt\tau_{1}^{\text{opt}} for which the external drive should remain on. We find that

τ1opt=2[2π+arctan((p16p28p2))πθ(8p)]λ16p2,\tau_{1}^{\text{opt}}=\frac{2\left[2\pi+\arctan{\left(\frac{p\sqrt{16-p^{2}}}{8-p^{2}}\right)}-\pi\theta(\sqrt{8}-p)\right]}{\lambda\sqrt{16-p^{2}}}\leavevmode\nobreak\ , (15)

where p=γ/λp=\gamma/\lambda and θ(x)\theta(x) is the Heaviside function. This is only finite for p<4p<4, and diverges otherwise. It is noteworthy that τ1opt\tau_{1}^{\text{opt}} is independent of both the feedback delay τ0\tau_{0} and the bath occupation N¯\bar{N}; it depends only the ratio p=γ/λp=\gamma/\lambda between the coupling rate to the bath and the external drive. Fig. 3(b) shows PeP_{e} vs. N¯\bar{N} for τ0=0\tau_{0}=0. We find that for γ/λ1.145\gamma/\lambda\leq 1.145 (determined numerically), there exists a critical value N¯c0\bar{N}_{\text{c}}\geq 0 such that Pe>1/2P_{e}>1/2 for all N¯<N¯c\bar{N}<\bar{N}_{\text{c}}. This therefore establishes the population inversion threshold. For γ/λ>1.145\gamma/\lambda>1.145 the drive is so weak that a pi pulse takes a time that is longer than the time-scale over which the bath exchanges energy with the qubit.

Next we consider a more realistic scenario where we have a time delay τ0\tau_{0} between the last detected jump xn=1x_{n}=-1 and the moment the external drive is turned on. Even in this more general case, the optimal drive duration that maximizes the population of the excited state remains given by Eq. (15Sup . In the low-temperature limit N¯0\bar{N}\rightarrow 0, the expression for PeP_{e}, evaluated at the optimal drive duration, simplifies to

limN¯0Pe=12eγτ1opt/2+14(γλ)2+γτ0.\displaystyle\lim_{\bar{N}\rightarrow 0}P_{e}=\frac{1}{2-e^{-\gamma\tau_{1}^{\text{opt}}/2}+\frac{1}{4}\left(\frac{\gamma}{\lambda}\right)^{2}+\gamma\tau_{0}}\leavevmode\nobreak\ . (16)

The time delay appear as an additive term γτ0\gamma\leavevmode\nobreak\ \tau_{0} in the denominator. The largest delay that still provides a population inversion (Pe=1/2P_{e}=1/2) in the low-temperature limit (N¯0\bar{N}\rightarrow 0) is

τ0max=4eγτ1opt/2(γ/λ)24γ.\displaystyle\tau_{0}^{\text{max}}=\frac{4e^{-\gamma\tau_{1}^{\text{opt}}/2}-(\gamma/\lambda)^{2}}{4\gamma}\leavevmode\nobreak\ . (17)

In Fig. 3(c), we illustrate how PeP_{e} is affected by varying the delay τ0\tau_{0}. We consider N¯=0\bar{N}=0, for which Pe=1/2P_{e}=1/2 occurs precisely at γ/λ=1.145\gamma/\lambda=1.145. The plots show how increasing τ0\tau_{0} deteriorates the population inversion. Fig. 3(d) plots Eqs. (15) and (17) vs. γ/λ\gamma/\lambda, again for N¯=0\bar{N}=0. In the strong-drive regime (γλ\gamma\ll\lambda), we find that τ0max1/γ\tau_{0}^{\text{max}}\sim 1/\gamma, indicating that the maximum delay is on the order of the system’s natural timescale γ1\gamma^{-1}. Conversely, τ0max=0\tau_{0}^{\text{max}}=0 for γ/λ>1.145\gamma/\lambda>1.145, reflecting that population inversion is no longer possible in this regime, regardless of the feedback delay. Additionally, we observe that γτ1opt\gamma\tau_{1}^{\text{opt}} increases with γ/λ\gamma/\lambda, implying that as the drive strength λ\lambda decreases relative to γ\gamma, the drive must remain on for a longer duration to optimize PeP_{e}.

As a final comment, it is worth mentioning that this protocol can be adapted to increase the population of the ground state: by turning on the drive when an absorption is detected, the drive will move the system toward the ground state through Rabi oscillations. In this case, the protocol effectively operates as a cooling protocol based on quantum jump detections.

Conclusion.—We derived a deterministic equation that describes a general feedback protocol, encompassing both discrete and continuous feedback. It encompass previous results as particular cases, and allows for the derivation of novel feedback schemes, as we have shown here in the context of quantum jumps. The method is based on a signal-resolved state, which provides access not only to the quantum dynamics, but also to the full statistics of the detection signal. Furthermore, it is a deterministic approach, which introduces new analytical tools that offer deeper physical insight into the interplay between feedback and quantum dynamics. Its applications span a wide range of fields, with particular relevance to quantum control, thermodynamics, and quantum information theory. We illustrated this with a detailed account of the use quantum jump detections and feedback for population inversion of a two-level system.

Acknowledgments.— The authors acknowledge fruitful discussions with Mark Mitchison, Guilherme Fiusa, Abhaya Hegde.

References

Appendix A Supplementary Material

We begin by introducing the language of instruments in Section B, defining the concepts that were used throughout this work. Next, we provide a detailed proof of the central results presented in the main text – namely, Results 1 and 2 – in Section C. Furthermore, we present a formulation in terms of matrix equations for the main results in Section D, where we also include a discussion about the steady state and generator of the feedback dynamics. In Section E, we present the details of the example used in the main text, namely the inversion protocol, and finally in Section F we show how some previous results can be recovered starting from Eq. (18).

ϱn+1(y)=x,yδy,fn+1(x,y)x(y)ϱn(y)\boxed{\varrho_{n+1}(y)=\sum_{x^{\prime},y^{\prime}}\delta_{y,f_{n+1}(x^{\prime},y^{\prime})}\mathcal{M}_{x^{\prime}}(y^{\prime})\varrho_{n}(y^{\prime})} (18)
Protocol Signal Instrument Equation
Combined Jump time dependent feedback yn=xn+yn1δxn,0y_{n}=x_{n}+y_{n-1}\delta_{x_{n},0} xρ=VxρVx\mathcal{M}_{x}\rho=V_{x}\rho V_{x}^{\dagger} Result 2
τn=δxn,0(τn1+δt)\tau_{n}=\delta_{x_{n},0}(\tau_{n-1}+\delta t)
Quantum Fokker-Planck Master Equation yn=j=1nγδteγ(nj)δtxjy_{n}=\sum_{j=1}^{n}\gamma\delta te^{-\gamma(n-j)\delta t}x_{j} zρ=Kz[eδt(ρ)]Kz\mathcal{M}_{z}\rho=K_{z}[e^{\delta t\mathcal{L}}(\rho)]K_{z}^{\dagger} Eq. (86)
Single Jump feedback yn=xny_{n}=x_{n} xρ=(x)[Vx(ρ)Vx]\mathcal{M}_{x}\rho=\mathcal{F}(x)[V_{x}(\rho)V_{x}^{\dagger}] Eq. (95)
Diffusion feedback yn=xny_{n}=x_{n} zρ=(z){Kz[eδt(ρ)]Kz}\mathcal{M}_{z}\rho=\mathcal{F}(z)\{K_{z}[e^{\delta t\mathcal{L}}(\rho)]K_{z}^{\dagger}\} Eq. (105)
Charge-based feedback yn=i=1nxiy_{n}=\sum_{i=1}^{n}x_{i} xρ=VxρVx\mathcal{M}_{x}\rho=V_{x}\rho V_{x}^{\dagger} Eq. (116)
Table 1: This table shows how previous results can be recovered starting from Eq. (18). Here, the operators VxV_{x} describe quantum jump detection, and xρ=KzρKz\mathcal{M}_{x}\rho=K_{z}\rho K_{z}^{\dagger} represents a Gaussian measurement, where KzK_{z} is defined in Eq. (79). The super-operator (x)\mathcal{F}(x), introduced below, is the well-known feedback super-operator, and eδte^{\delta t\mathcal{L}} denotes unitary evolution, where ρ=i[H,ρ]\mathcal{L}\rho=-i[H,\rho] for some Hamiltonian HH.

Appendix B Instruments and Quantum Measurements

Let us consider a collection {x}xΣ\{\mathcal{M}_{x}\}_{x\in\Sigma} of trace non-increasing maps – Tr[xρ]Tr[ρ]\mathrm{Tr}[\mathcal{M}_{x}\rho]\leq\mathrm{Tr}[\rho] for any density operator ρ\rho – where Σ\Sigma is an arbitrary set. If the sum xΣx\sum_{x\in\Sigma}\mathcal{M}_{x} is a completely positive trace-preserving (CPTP) map, then the maps x\mathcal{M}_{x} are called instruments. Instruments can describe both transitions induced by measurement processes and those arising from dynamical evolutions. First, let us show that any measurement can be written in terms of instruments, and subsequently consider a dynamical evolution combined with detection.

Given a observable A^\hat{A} that can assume any value aa in the set ΣA\Sigma_{A}, a measurement of A^\hat{A} is defined by a set {Va}aΣA\{V_{a}\}_{a\in\Sigma_{A}} such that

aΣAVaVa=𝟙,\sum_{a\in\Sigma_{A}}V_{a}^{\dagger}V_{a}=\mathbb{1}, (19)

where the set {Va}aΣ\{V_{a}\}_{a\in\Sigma} corresponds to the Kraus operators associated with the measurement of A^\hat{A}. Also, the probability distribution of the outcomes aa in a state ρ\rho is given by

p(a|ρ)=Tr[VaρVa],p(a|\rho)=\mathrm{Tr}[V_{a}\rho V_{a}^{\dagger}]\leavevmode\nobreak\ , (20)

and the post-detection state when we measure result aa is

ρa=VaρVaTr[VaρVa].\rho_{a}=\frac{V_{a}\rho V_{a}^{\dagger}}{\mathrm{Tr}[V_{a}\rho V_{a}^{\dagger}]}\leavevmode\nobreak\ . (21)

For instance, if one considers a projective measurement of AA, then Va=|ψaψa|V_{a}=\left|\psi_{a}\right>\left<\psi_{a}\right|, where |ψa\left|\psi_{a}\right> is the eigenvector of AA with eigenvalue aa.

Now, we can define the new maps ~VaρVa\widetilde{\mathcal{M}}\equiv V_{a}\rho V_{a}^{\dagger} for any aΣAa\in\Sigma_{A}, hence the probability distribution p(a|ρ)p(a|\rho) and the post-detection state ρa\rho_{a} can be written as

p(a|ρ)\displaystyle p(a|\rho) =\displaystyle= Tr[~aρ],\displaystyle\mathrm{Tr}[\widetilde{\mathcal{M}}_{a}\rho]\leavevmode\nobreak\ , (22)
ρa\displaystyle\rho_{a} =\displaystyle= ~aρTr[~aρ].\displaystyle\frac{\widetilde{\mathcal{M}}_{a}\rho}{\mathrm{Tr}[\widetilde{\mathcal{M}}_{a}\rho]}\leavevmode\nobreak\ . (23)

Let us remember that a set {Vk}k\{V_{k}\}_{k} satisfying the completeness relation kVkVk=𝟙\sum_{k}V_{k}^{\dagger}V_{k}=\mathbb{1} defines a quantum channel (CPTP map) Λ\Lambda by

Λ(ρ)=kVkρVk,\Lambda(\rho)=\sum_{k}V_{k}\rho V_{k}^{\dagger}\leavevmode\nobreak\ , (24)

where Eq. (24) is called the Kraus decomposition of the channel Λ\Lambda. Thus, we can conclude that aΣ~a\sum_{a\in\Sigma}\widetilde{\mathcal{M}}_{a} is a CPTP map. Furthermore, since aΣAVaVa=𝟙\sum_{a\in\Sigma_{A}}V_{a}^{\dagger}V_{a}=\mathbb{1} it follows that VaVa𝟙V_{a}^{\dagger}V_{a}\leq\mathbb{1}, hence Tr[~aρ]Tr[ρ]\mathrm{Tr}[\widetilde{\mathcal{M}}_{a}\rho]\leq\mathrm{Tr}[\rho] for any quantum state ρ\rho. Therefore, ~a\widetilde{\mathcal{M}}_{a} is an instrument that describes the detection of the outcome aΣAa\in\Sigma_{A}. We can generalize the measurement postulates using the concept of instruments, where a quantum measurement is described by a collection {x}xΣ\{\mathcal{M}_{x}\}_{x\in\Sigma} of trace non-increasing super-operators x\mathcal{M}_{x} such that the sum xΣx\sum_{x\in\Sigma}\mathcal{M}_{x} corresponds to a CPTP map, where Σ\Sigma is the set of possible outcomes xx, the probability distribution of xx is given by px=Tr[xρ]p_{x}=\mathrm{Tr}[\mathcal{M}_{x}\rho], and the system state updates to ρxρpx\rho\rightarrow\frac{\mathcal{M}_{x}\rho}{p_{x}} after a measurement of the result xx.

With the language of instruments, we can also address the dynamical evolution between two detections. For example, let us consider a sequential measurement, where a quantum measurement described by a set {Vk}k\{V_{k}\}_{k} of Kraus operators is applied at any time tj=jΔtt_{j}=j\Delta t, and let us suppose that the system evolves dynamically from tjt_{j} to tj+1t_{j+1} according to the a master equation tρt=ρt\partial_{t}\rho_{t}=\mathcal{L}\rho_{t}, where \mathcal{L} is (for simplicity) time independent. In this case, the instrument that describes joint measurement-dynamics transitions between two detection is given by

x=~xeΔt,\mathcal{M}_{x}=\widetilde{\mathcal{M}}_{x}e^{\Delta t\mathcal{L}}\leavevmode\nobreak\ , (25)

where ~xρ=VxρVx\widetilde{\mathcal{M}}_{x}\rho=V_{x}\rho V_{x}^{\dagger} describes the measurement, and eΔte^{\Delta t\mathcal{L}} describes the dynamical evolution of the system. Therefore, when we are dealing with sequential measurements, the instruments {x}\{\mathcal{M}_{x}\} are also including the possibility of dynamical evolution between two detections.

Appendix C Proof of the main results

C.1 Proof of the Result 1

Suppose a sequential quantum detection described by the instruments {x(y)}xΣ\{\mathcal{M}_{x}(y)\}_{x\in\Sigma}, as defined in the main text, where yy denotes a stochastic signal. Let us consider a causal signal governed by the recursive relation

yn+1=fn+1(xn+1,yn),y_{n+1}=f_{n+1}(x_{n+1},y_{n}), (26)

where fn+1(x,y)f_{n+1}(x,y) is a given function. This signal is causal in the sense that it depends on the previous signal yny_{n} and the measurement outcome xn+1x_{n+1} at time tn+1=(n+1)δtt_{n+1}=(n+1)\delta t. After nn measurements, we have the data set x1:n(x1,,xn)x_{1:n}\equiv(x_{1},\cdots,x_{n}), and the conditional state is denoted by ρn(x1:n)\rho_{n}(x_{1:n}). The signal-resolved state is defined as

ϱn(y)E1:n[ρn(x1,,xn)δy,yn],\varrho_{n}(y)\equiv E_{1:n}[\rho_{n}(x_{1},\dots,x_{n})\delta_{y,y_{n}}]\leavevmode\nobreak\ , (27)

where E1:nE_{1:n} represents the ensemble average over all possible trajectories with nn outcomes in the data set. By using the conditional update rule for the post-measurement state, ρn+1=xn+1(yn)ρn(x1:n)P(xn+1|x1:n)\rho_{n+1}=\frac{\mathcal{M}_{x_{n+1}}(y_{n})\rho_{n}(x_{1:n})}{P(x_{n+1}|x_{1:n})}, we obtain

ϱn+1(y)\displaystyle\varrho_{n+1}(y) =E1:n+1[ρn+1(x1:n+1)δy,yn+1]\displaystyle=E_{1:n+1}[\rho_{n+1}(x_{1:n+1})\delta_{y,y_{n+1}}]
=E1:n+1[xn+1(yn)ρn(x1:n)P(xn+1|x1:n)δy,fn+1(xn+1,yn)].\displaystyle=E_{1:n+1}\left[\frac{\mathcal{M}_{x_{n+1}}(y_{n})\rho_{n}(x_{1:n})}{P(x_{n+1}|x_{1:n})}\delta_{y,f_{n+1}(x_{n+1},y_{n})}\right]\leavevmode\nobreak\ .

Let us denote En+1|1:n[]E_{n+1|1:n}[\cdot] as the conditional average over xn+1x_{n+1} given the previous outcomes x1:nx_{1:n}. Using the identity E1:n+1[]=E1:n[En+1|1:n[]]E_{1:n+1}[\cdot]=E_{1:n}[E_{n+1|1:n}[\cdot]], we find

ϱn+1(y)\displaystyle\varrho_{n+1}(y) =E1:n{xδy,fn+1(x,yn)x(yn)ρn(x1:n)}\displaystyle=E_{1:n}\left\{\sum_{x^{\prime}}\delta_{y,f_{n+1}(x^{\prime},y_{n})}\mathcal{M}_{x^{\prime}}(y_{n})\rho_{n}(x_{1:n})\right\}
=xE1:n[δy,fn+1(x,yn)x(yn)ρn(x1:n)].\displaystyle=\sum_{x^{\prime}}E_{1:n}\left[\delta_{y,f_{n+1}(x^{\prime},y_{n})}\mathcal{M}_{x^{\prime}}(y_{n})\rho_{n}(x_{1:n})\right]\leavevmode\nobreak\ .

To isolate the definition of ϱn(y)\varrho_{n}(y), we introduce a dummy summation over yy^{\prime} and a Kronecker delta as follows

ϱn+1(y)\displaystyle\varrho_{n+1}(y) =xyE1:n{δy,fn+1(x,y)δy,ynx(yn)ρn(x1:n)},\displaystyle=\sum_{x^{\prime}}\sum_{y^{\prime}}E_{1:n}\left\{\delta_{y,f_{n+1}(x^{\prime},y^{\prime})}\delta_{y^{\prime},y_{n}}\mathcal{M}_{x^{\prime}}(y_{n})\rho_{n}(x_{1:n})\right\}\leavevmode\nobreak\ ,

where we have inserted a Kronecker delta along with a sum over yy^{\prime}, then replaced yny_{n} in the expression with yy^{\prime} via δy,yn\delta_{y^{\prime},y_{n}}. Notably, this also allows us to substitute yny_{n} inside x(yn)\mathcal{M}_{x^{\prime}}(y_{n}) with the deterministic value yy^{\prime}, giving

ϱn+1(y)\displaystyle\varrho_{n+1}(y) =xyE1:n{δy,fn+1(x,y)δy,ynx(y)ρn(x1:n)}\displaystyle=\sum_{x^{\prime}}\sum_{y^{\prime}}E_{1:n}\left\{\delta_{y,f_{n+1}(x^{\prime},y^{\prime})}\delta_{y^{\prime},y_{n}}\mathcal{M}_{x^{\prime}}(y^{\prime})\rho_{n}(x_{1:n})\right\}
=x,yδy,fn+1(x,y)x(y)E1:n{ρn(x1:n)δy,yn}.\displaystyle=\sum_{x^{\prime},y^{\prime}}\delta_{y,f_{n+1}(x^{\prime},y^{\prime})}\mathcal{M}_{x^{\prime}}(y^{\prime})E_{1:n}\left\{\rho_{n}(x_{1:n})\delta_{y^{\prime},y_{n}}\right\}\leavevmode\nobreak\ .

Finally, we identify the definition of the signal-resolved state as ϱn(y)=E1:n{ρn(x1:n)δy,yn}\varrho_{n}(y^{\prime})=E_{1:n}\left\{\rho_{n}(x_{1:n})\delta_{y^{\prime},y_{n}}\right\}, and we arrive at

ϱn+1(y)=x,yδy,fn+1(x,y)x(y)ϱn(y),\varrho_{n+1}(y)=\sum_{x^{\prime},y^{\prime}}\delta_{y,f_{n+1}(x^{\prime},y^{\prime})}\mathcal{M}_{x^{\prime}}(y^{\prime})\varrho_{n}(y^{\prime})\leavevmode\nobreak\ , (28)

as described in Result 1.

C.2 Proof of the Result 2

C.2.1 Jump time dependent feedback

Suppose that we are applying a sequential measurement of quantum jumps described by the Kraus operators Vk=δtLkV_{k}=\sqrt{\delta t}L_{k} for kΣk\in\Sigma, and V0=𝟙iδtHeffV_{0}=\mathbb{1}-i\delta tH_{\text{eff}}, where Heff=Hi2kΣLkLkH_{\text{eff}}=H-\frac{i}{2}\sum_{k\in\Sigma}L_{k}^{\dagger}L_{k} is the effective Hamiltonian, and Σ\Sigma is the set of all possible jumps. Hence, the correspondent instrument is defined by xρ=VxρVx\mathcal{M}_{x}\rho=V_{x}\rho V_{x}^{\dagger}. Consider a feedback scheme conditioned on the last detected quantum jump and the time interval since its occurrence. The stochastic signal that records the last jump is given by

yn=xn+yn1δxn,0,y_{n}=x_{n}+y_{n-1}\delta_{x_{n},0}, (29)

where δxn,0\delta_{x_{n},0} is the Kronecker delta, and xnx_{n} represents the outcome of the quantum jump detection at time tn=nδtt_{n}=n\delta t. In this context, xn=0x_{n}=0 corresponds to a no-jump detection, while xn=kx_{n}=k indicates that a quantum jump occurred in channel kk. Furthermore, the signal that tracks the time elapsed since the last detected jump is given by

τn=δxn,0(τn1+δt),\tau_{n}=\delta_{x_{n},0}(\tau_{n-1}+\delta t)\leavevmode\nobreak\ , (30)

We are considering generic initial condition for both signals, where y0=k¯Σy_{0}=\bar{k}\in\Sigma is fixed and τ0=0\tau_{0}=0. It is equivalent to say that the system is prepared such that we have a jump k¯\bar{k} at time t=0t=0. Hence, the feedback action is implemented in the quantum jump detections at time tn=nδtt_{n}=n\delta t by considering the Kraus operators Vk(yn1,τn1)V_{k}(y_{n-1},\tau_{n-1}).

By construction, both signals defined in Eq. (29) and Eq. (30) satisfy the causality condition yn=f(xn,yn1)y_{n}=f(x_{n},y_{n-1}), as required in Result 1. Hence, applying the Result 1 (Eq. (18)) to those signals, we obtain

ϱn+1(y,τ)\displaystyle\varrho_{n+1}(y,\tau) =\displaystyle= x,y,τδy,x+yδx,0δτ,δx,0(τ+δt)x(y,τ)ϱn(y,τ)\displaystyle\sum_{x^{\prime},y^{\prime},\tau^{\prime}}\delta_{y,x^{\prime}+y^{\prime}\delta_{x^{\prime},0}}\delta_{\tau,\delta_{x^{\prime},0}(\tau^{\prime}+\delta t)}\mathcal{M}_{x^{\prime}}(y^{\prime},\tau^{\prime})\varrho_{n}(y^{\prime},\tau^{\prime}) (31)
=\displaystyle= y,τδy,yδτ,(τ+δt)0(y,τ)ϱn(y,τ)+xΣy,τδy,xδτ,0x(y,τ)ϱn(y,τ)\displaystyle\sum_{y^{\prime},\tau^{\prime}}\delta_{y,y^{\prime}}\delta_{\tau,(\tau^{\prime}+\delta t)}\mathcal{M}_{0}(y^{\prime},\tau^{\prime})\varrho_{n}(y^{\prime},\tau^{\prime})+\sum_{x^{\prime}\in\Sigma}\sum_{y^{\prime},\tau^{\prime}}\delta_{y,x^{\prime}}\delta_{\tau,0}\mathcal{M}_{x^{\prime}}(y^{\prime},\tau^{\prime})\varrho_{n}(y^{\prime},\tau^{\prime}) (32)
=\displaystyle= h=0nδτ,(hδt+δt)0(y,hδt)ϱn(y,hδt)+δτ,0yΣh=0ny(y,hδt)ϱn(y,hδt).\displaystyle\sum_{h=0}^{n}\delta_{\tau,(h\delta t+\delta t)}\mathcal{M}_{0}(y,h\delta t)\varrho_{n}(y,h\delta t)+\delta_{\tau,0}\sum_{y^{\prime}\in\Sigma}\sum_{h=0}^{n}\mathcal{M}_{y}(y^{\prime},h\delta t)\varrho_{n}(y^{\prime},h\delta t)\leavevmode\nobreak\ . (33)

The first equality follows directly from Result 1 [Eq. (18)]. In the second equality, we decompose the sum over xx^{\prime} into the two possible cases: x=0x^{\prime}=0 (no-jump detection) and xΣx^{\prime}\in\Sigma (jump detection). The transition from the second to the third line is obtained by using the Kronecker delta to evaluate the first sum over yy^{\prime} and the last sum over xx^{\prime}. Finally, in Eq. (33), we make the summation ranges explicit: yΣy^{\prime}\in\Sigma and τ=hδt\tau^{\prime}=h\delta t with h=0,1,,nh=0,1,\dots,n.

Note that Eq. (33) can be separated into two cases: τ=0\tau=0 and τ>0\tau>0. When τ=0\tau=0, only the second term survives, yielding

ϱn+1(y,0)=yΣh=0nδt𝒥y(y,hδt)ϱn(y,hδt),\varrho_{n+1}(y,0)=\sum_{y^{\prime}\in\Sigma}\sum_{h=0}^{n}\delta t\mathcal{J}_{y}(y^{\prime},h\delta t)\varrho_{n}(y^{\prime},h\delta t)\leavevmode\nobreak\ , (34)

where 𝒥yϱ=LyϱLy\mathcal{J}_{y}\varrho=L_{y}\varrho L_{y}^{\dagger} represents the channel yΣy\in\Sigma. On the other hand, for τ>0\tau>0, namely τ=pδt\tau=p\delta t where p=1,,n+1p=1,\cdots,n+1, only the first term of Eq. (33) survives, and we have

ϱn+1(y,pδt)=h=0nδpδt,(hδt+δt)0(y,hδt)ϱn(y,hδt)=0(y,pδtδt)ϱn(y,pδtδt)=eδt0(y,(p1)δt)ϱn(y,(p1)δt),\varrho_{n+1}(y,p\delta t)=\sum_{h=0}^{n}\delta_{p\delta t,(h\delta t+\delta t)}\mathcal{M}_{0}(y,h\delta t)\varrho_{n}(y,h\delta t)=\mathcal{M}_{0}(y,p\delta t-\delta t)\varrho_{n}(y,p\delta t-\delta t)=e^{\delta t\mathcal{L}_{0}(y,(p-1)\delta t)}\varrho_{n}(y,(p-1)\delta t)\leavevmode\nobreak\ , (35)

where we used the Kronecker delta to perform the sum over hh, and 0ϱ=V0ϱV0=eδt0ϱ\mathcal{M}_{0}\varrho=V_{0}\varrho V_{0}^{\dagger}=e^{\delta t\mathcal{L}_{0}}\varrho for quantum jumps. Note that p=1,,n+1p=1,\cdots,n+1 ensures (p1)δt0(p-1)\delta t\geq 0. Then, by applying the last equation recursively, we obtain

ϱn+1(y,pδt)=eδt0(y,(p1)δt)eδt0(y,(p2)δt)eδt0(y,0)ϱn+1p(y,0).\varrho_{n+1}(y,p\delta t)=e^{\delta t\mathcal{L}_{0}(y,(p-1)\delta t)}e^{\delta t\mathcal{L}_{0}(y,(p-2)\delta t)}\cdots e^{\delta t\mathcal{L}_{0}(y,0)}\varrho_{n+1-p}(y,0)\leavevmode\nobreak\ . (36)

By taking the continuous measurement limit δt0\delta t\rightarrow 0 and pp\rightarrow\infty such that τ=pδt\tau=p\delta t remains fixed (and similarly t=n,δtt=n,\delta t), and that ϱn(y,p,δt)ϱt(y,τ)\varrho_{n}(y,p,\delta t)\rightarrow\varrho_{t}(y,\tau), Eq. (36) becomes

ϱt(y,τ)=𝒯[e0τ𝑑s0(y,s)]ϱtτ(y,0).\varrho_{t}(y,\tau)=\mathcal{T}\left[e^{\int_{0}^{\tau}ds\mathcal{L}_{0}(y,s)}\right]\varrho_{t-\tau}(y,0)\leavevmode\nobreak\ . (37)

Finally, by applying this same limit in Eq. (34), we have

ϱt(y,0)=yΣ0t𝑑τ𝒥y(y,τ)ϱt(y,τ),\varrho_{t}(y,0)=\sum_{y\in\Sigma}\int_{0}^{t}d\tau^{\prime}\mathcal{J}_{y}(y^{\prime},\tau^{\prime})\varrho_{t}(y^{\prime},\tau^{\prime})\leavevmode\nobreak\ , (38)

and by using Eq. (36) we finally have

ϱt(y,0)=yΣ0t𝑑τ𝒥y(y,τ)𝒯[e0τ𝑑s0(y,s)]ϱtτ(y,0).\varrho_{t}(y,0)=\sum_{y\in\Sigma}\int_{0}^{t}d\tau^{\prime}\mathcal{J}_{y}(y^{\prime},\tau^{\prime})\mathcal{T}\left[e^{\int_{0}^{\tau^{\prime}}ds\mathcal{L}_{0}(y^{\prime},s)}\right]\varrho_{t-\tau}(y^{\prime},0)\leavevmode\nobreak\ . (39)

Note that Eq. (39) represents a sum over all possible previous trajectories: a jump of type yy can occur at time tt if it was preceded by a jump of type yy^{\prime} at time tτt-\tau^{\prime}, followed by no jumps in the interval (tτ,t](t-\tau^{\prime},t]. The sum over yy^{\prime} and the integral over τ\tau^{\prime} account for all such possibilities. This completes the proof of Eq. (37) and Eq. (39) of Result 2.

In the next section, let us analyses the particularly interesting case where the feedback depends only on the quantum jumps part, disregarding the time elapsed since the last detected jump.

C.2.2 Jump feedback

Let us consider that the feedback depends solely on the last detected jump, hence the instruments can be replaced as x(y,τ)x(y)\mathcal{M}_{x}(y,\tau)\rightarrow\mathcal{M}_{x}(y). The signal-resolved for this feedback dependence is related with the previous one by ϱt(y)=0t𝑑τϱt(y,τ)\varrho_{t}(y)=\int_{0}^{t}d\tau\varrho_{t}(y,\tau), that corresponds to a marginalization over the time entry of ϱt(y,τ)\varrho_{t}(y,\tau). Therefore, by applying this marginalization on both sides of Eq. (33), we have

ϱn+1(y)\displaystyle\varrho_{n+1}(y) =\displaystyle= 0(y)ϱn(y)+yΣy(y)ϱn(y)\displaystyle\mathcal{M}_{0}(y)\varrho_{n}(y)+\sum_{y^{\prime}\in\Sigma}\mathcal{M}_{y}(y^{\prime})\varrho_{n}(y^{\prime}) (40)
=\displaystyle= V0(y)ϱnV0(y)+δtyΣLy(y)ϱn(y)Ly(y),\displaystyle V_{0}(y)\varrho_{n}V_{0}^{\dagger}(y)+\delta t\sum_{y^{\prime}\in\Sigma}L_{y}(y^{\prime})\varrho_{n}(y^{\prime})L_{y}^{\dagger}(y^{\prime})\leavevmode\nobreak\ , (41)

where in the second equality we have used the instruments related with quantum jump detections. For the first term, we have

V0(y)ϱnV0(y)=(𝟙+δt0(y))ϱn(y)=ϱn(y)iδt[H(y),ϱn(y)]δt2kΣ{Lk(y)Lk(y),ϱn(y)},V_{0}(y)\varrho_{n}V_{0}^{\dagger}(y)=(\mathbb{1}+\delta t\mathcal{L}_{0}(y))\varrho_{n}(y)=\varrho_{n}(y)-i\delta t[H(y),\varrho_{n}(y)]-\frac{\delta t}{2}\sum_{k\in\Sigma}\{L_{k}^{\dagger}(y)L_{k}(y),\varrho_{n}(y)\}\leavevmode\nobreak\ , (42)

then we have

ϱn+1(y)=ϱn(y)iδt[H(y),ϱn(y)]δt2kΣ{Lk(y)Lk(y),ϱn(y)}+δtyΣLy(y)ϱn(y)Ly(y).\varrho_{n+1}(y)=\varrho_{n}(y)-i\delta t[H(y),\varrho_{n}(y)]-\frac{\delta t}{2}\sum_{k\in\Sigma}\{L_{k}^{\dagger}(y)L_{k}(y),\varrho_{n}(y)\}+\delta t\sum_{y^{\prime}\in\Sigma}L_{y}(y^{\prime})\varrho_{n}(y^{\prime})L_{y}^{\dagger}(y^{\prime})\leavevmode\nobreak\ . (43)

By taking the continuous measurement limit δt0\delta t\rightarrow 0, where ϱn(y)ϱt(y)\varrho_{n}(y)\rightarrow\varrho_{t}(y) and (ϱn+1(y)ϱn(y))/δttϱt(y)(\varrho_{n+1}(y)-\varrho_{n}(y))/\delta t\rightarrow\partial_{t}\varrho_{t}(y), we finally obtain

tϱt(y)\displaystyle\partial_{t}\varrho_{t}(y) =\displaystyle= i[H(y),ϱt(y)]+kΣ(Ly(k)ϱt(k)Ly(k)12{Lk(y)Lk(y),ϱt(y)}).\displaystyle-i\big{[}H(y),\varrho_{t}(y)\big{]}+\sum_{k\in\Sigma}\left(\ L_{y}(k)\varrho_{t}(k)L_{y}^{\dagger}(k)-\frac{1}{2}\{L_{k}^{\dagger}(y)L_{k}(y),\varrho_{t}(y)\}\right). (44)

Therefore, the master equation given by Eq. (44) provides the feedback dynamics when only the last jump is considered. We can recover the unconditional state of the system by marginalizing the signal-resolved state over yy, ρ¯t=yΣϱt(y)\bar{\rho}_{t}=\sum_{y\in\Sigma}\varrho_{t}(y), and the probability of detecting a jump yy at time tt is given by Tr[ϱt(y)]\mathrm{Tr}[\varrho_{t}(y)].

Appendix D Matrix equations: generators and steady state of feedback dynamics

We can rewrite our results in terms of matrix equations as follows. Let us define the propagator Λy(n)(y)xδy,fn(x,y)x(y)\Lambda_{y}^{(n)}(y^{\prime})\equiv\sum_{x^{\prime}}\delta_{y,f_{n}(x^{\prime},y^{\prime})}\mathcal{M}_{x^{\prime}}(y^{\prime}), so that Eq. (18) takes the form ϱn+1(y)=yΛy(n)(y)ϱn(y)\varrho_{n+1}(y)=\sum_{y^{\prime}}\Lambda_{y}^{(n)}(y^{\prime})\varrho_{n}(y^{\prime}), where yΛy(n)(y)=xx(y)\sum_{y}\Lambda_{y}^{(n)}(y^{\prime})=\sum_{x^{\prime}}\mathcal{M}_{x^{\prime}}(y^{\prime}) is completely positive and trace-preserving (CPTP) for any yy^{\prime}, ensuring consistency with the general evolution of a hybrid quantum state Oppenheim (2023). This new representation of Eq. (18) resembles a matrix product. To make this structure explicit, let us define the column vector ϱn\vec{\varrho}_{n}, whose components correspond to the signal-resolved states ϱn(y)\varrho_{n}(y), with yy ranging over the possible values of the stochastic signal yny_{n} at time tn=nδtt_{n}=n\delta t. We also introduce the double-indexed object 𝚲𝐧[Λy(n)(y)]y,y\mathbf{\Lambda_{n}}\equiv[\Lambda_{y}^{(n)}(y^{\prime})]{y,y^{\prime}}, which represents a matrix whose components are the superoperators Λy(n)(y)\Lambda_{y}^{(n)}(y^{\prime}). With this notation, and using the standard matrix product, Eq. (18) takes the compact form

ϱn+1=𝚲𝐧ϱn.\vec{\varrho}_{n+1}=\mathbf{\Lambda_{n}}\cdot\vec{\varrho}_{n}\leavevmode\nobreak\ . (45)

Moreover, within this matrix representation, we can define the generator of the feedback dynamics. Note that ϱn+1ϱn=(𝚲𝐧Id)ϱn\vec{\varrho}_{n+1}-\vec{\varrho}_{n}=(\mathbf{\Lambda_{n}}-I_{d})\vec{\varrho}_{n}, where IdI_{d} represents the identity. Hence, if limΔt0(𝚲𝐧Id)/Δt\lim_{\Delta t\rightarrow 0}(\mathbf{\Lambda_{n}}-I_{d})/\Delta t is well defined, we can introduce the generator fb=limΔt0(𝚲𝐧Id)/Δt\mathcal{L}_{\text{fb}}=\lim_{\Delta t\rightarrow 0}(\mathbf{\Lambda_{n}}-I_{d})/\Delta t, then tϱt=fbϱt\partial_{t}\vec{\varrho}_{t}=\mathcal{L}_{\text{fb}}\cdot\vec{\varrho}_{t}. Finally, the signal-resolved stead state is defined as ϱsslimtϱt\vec{\varrho}_{ss}\equiv\lim_{t\rightarrow\infty}\vec{\varrho}_{t}, or equivalently ϱss(y)=limtϱt(y)\varrho_{ss}(y)=\lim_{t\rightarrow\infty}\varrho_{t}(y), and the unconditional steady state is recovered by marginalizing over yy, ρ¯ss=yϱss(y)\bar{\rho}_{ss}=\sum_{y}\varrho_{ss}(y). By considering Eq. (45), we have

ϱss=𝚲ϱss,\vec{\varrho}_{ss}=\mathbf{\Lambda_{\infty}}\cdot\vec{\varrho}_{ss}\leavevmode\nobreak\ , (46)

where 𝚲=limn𝚲𝐧\mathbf{\Lambda_{\infty}}=\lim_{n\rightarrow\infty}\mathbf{\Lambda_{n}}. Hence, one find that the signal-resolved steady state can be seen as the eigenvector of 𝚲\mathbf{\Lambda_{\infty}} with unit eigenvalue.

One can find the steady state of the system by considering the limit tt\rightarrow\infty in Result 2. Since 0(y,τ)\mathcal{L}_{0}(y,\tau) is a no-jump Liouvillian, it has negative eigenvalues Landi et al. (2024). Hence, G(y,τ)=𝒯[e0τ𝑑s0(y,s)]G(y^{\prime},\tau^{\prime})=\mathcal{T}\left[e^{\int_{0}^{\tau^{\prime}}ds\mathcal{L}_{0}(y^{\prime},s)}\right] decays to zero for large τ\tau^{\prime}. Thus, in the limit tt\rightarrow\infty, we can take ϱtτ(y,0)\varrho_{t-\tau^{\prime}}(y^{\prime},0) out of the integral over τ\tau^{\prime} in Eq. (39), yielding

ϱss(y,0)=yΣ(0𝑑τ𝒥y(y,τ)G(y,τ))ϱss(y,0).\varrho_{ss}(y,0)=\sum_{y^{\prime}\in\Sigma}\left(\int_{0}^{\infty}\leavevmode\nobreak\ d\tau^{\prime}\mathcal{J}_{y}(y^{\prime},\tau^{\prime})G(y^{\prime},\tau^{\prime})\right)\varrho_{ss}(y^{\prime},0). (47)

Following the same way as in Eq. (45), we can rewrite Eq. (47) in a matrix form as follows. Let us define

𝚲[0𝑑τ𝒥y(y,τ)G(y,τ)]y,y,\mathbf{\Lambda_{\infty}}\equiv\left[\int_{0}^{\infty}\leavevmode\nobreak\ d\tau^{\prime}\mathcal{J}_{y}(y^{\prime},\tau^{\prime})G(y^{\prime},\tau^{\prime})\right]_{y,y^{\prime}}\leavevmode\nobreak\ , (48)

and ϱss(0)\vec{\varrho}_{ss}(0) represents a column vector whose the components are ϱss(y,0)\varrho_{ss}(y,0) for yΣy\in\Sigma. Then, Eq. (47) becomes ϱss(0)=𝚲ϱss(0)\vec{\varrho}_{ss}(0)=\mathbf{\Lambda_{\infty}}\cdot\vec{\varrho}_{ss}(0), where the steady state is the eigenvector of 𝚲\mathbf{\Lambda_{\infty}} corresponding to the eigenvalue 1, as discussed in Eq. (46). Therefore, the eigenvector of 𝚲\mathbf{\Lambda}_{\infty} provides us ϱss(y,0)\varrho_{ss}(y,0). On the other hand, Eq. (37) gives ϱt(y,τ)\varrho_{t}(y,\tau), where we have

ϱss(y,τ)=𝒯[e0τ𝑑s0(y,s)]ϱss(y,0).\varrho_{ss}(y,\tau)=\mathcal{T}\left[e^{\int_{0}^{\tau}ds\mathcal{L}_{0}(y,s)}\right]\varrho_{ss}(y,0)\leavevmode\nobreak\ . (49)

Now, let us find the steady state for the jump feedback, namely when the feedback is based only on the most recent jump. In this case, the feedback dynamics is described by Eq. (44). Note that we can rewrite this equation as

ϱt(y)t=Φ(y)ϱt(y)+kΣLy(k)ϱt(k)Ly(k),\frac{\partial\varrho_{t}(y)}{\partial t}=\Phi(y)\varrho_{t}(y)+\sum_{k\in\Sigma}L_{y}(k)\varrho_{t}(k)L_{y}^{\dagger}(k)\leavevmode\nobreak\ , (50)

where

Φ(y)ϱt(y)i[H(y),ϱt(y)]12kΣ{Lk(y)Lk(y),ϱt(y)}.\Phi(y)\varrho_{t}(y)\equiv-i[H(y),\varrho_{t}(y)]-\frac{1}{2}\sum_{k\in\Sigma}\left\{L_{k}^{\dagger}(y)L_{k}(y),\varrho_{t}(y)\right\}. (51)

The term kΣLy(k)ϱt(k)Ly(k)\sum_{k\in\Sigma}L_{y}(k)\varrho_{t}(k)L_{y}^{\dagger}(k) couples different values of kΩk\in\Omega in the differential equation for ϱt(y)\varrho_{t}(y) and describes the contribution of various jump channels due to the feedback action.

Let us define a column vector ϱt\vec{\varrho}_{t} whose components are the elements ϱt(y)\varrho_{t}(y) for each jump yΣy\in\Sigma. Hence, the previous equation can be compactly expressed as

ϱtt=Ψϱt+Ξϱt,\frac{\partial\vec{\varrho}_{t}}{\partial t}=\Psi\vec{\varrho}_{t}+\Xi\vec{\varrho}_{t}\leavevmode\nobreak\ , (52)

where Ψ\Psi is a diagonal matrix whose diagonal elements are given by the super-operators Φ\Phi, i.e., [Ψ]mn=Φ(m)δmn[\Psi]_{mn}=\Phi(m)\delta_{mn} with m,nΣm,n\in\Sigma, and Ξ\Xi is a matrix whose entries are the super-operators 𝒥m(n)\mathcal{J}_{m}(n), i.e., [Ξ]mn=𝒥m(n)[\Xi]_{mn}=\mathcal{J}_{m}(n). Defining ΠΨ+Ξ\Pi\equiv\Psi+\Xi, we finally obtain

ϱtt=Πϱt.\frac{\partial\vec{\varrho}_{t}}{\partial t}=\Pi\vec{\varrho}_{t}\leavevmode\nobreak\ . (53)

Therefore, Eq. (53) provides a compact, vectorial formulation of Result 2 for the signal-resolved state when the feedback is based solely on the last detected jump. In the steady state, where tϱt=0\partial_{t}\vec{\varrho}_{t}=0, the solution satisfies the eigenvalue equation

Πϱss=0,\Pi\vec{\varrho}_{ss}=0\leavevmode\nobreak\ , (54)

providing the signal-resolved stead state ϱt(y)\varrho_{t}(y) for each jump yΣy\in\Sigma. Consequently, the unconditional steady state is given by

ρss=yΣϱss(y).\rho_{ss}=\sum_{y\in\Sigma}\varrho_{ss}(y)\leavevmode\nobreak\ . (55)

Appendix E Example and details

E.1 Rotating frame and Rotating Wave Approximation

In the example presented in the main text, we used a rotating frame together with the rotating wave approximation. In this section, we briefly review these concepts. Consider a two-level system with transition frequency ω\omega, driven by an external field with frequency ωd\omega_{d} and coupling strength λ\lambda. The Hamiltonian in the Schrodinger picture is given by (=1\hbar=1)

H=ω2σz+Ωcos((tωd))σx,H=-\frac{\omega}{2}\sigma_{z}+\Omega\cos{(t\omega_{d})}\sigma_{x}\leavevmode\nobreak\ , (56)

where σz,x\sigma_{z,x} are the Pauli matrices. In this convention, the ground state |g\left|g\right> has energy ω/2-\omega/2. The Hamiltonian in the rotating frame is defined as

Heitωd2σzHeitωd2σz+ωd2σz=Δ2σz+Ω2(σ++σ)+Ω2(e2iωdtσ++e2iωdtσ),H^{\prime}\equiv e^{\frac{-it\omega_{d}}{2}\sigma_{z}}He^{\frac{it\omega_{d}}{2}\sigma_{z}}+\frac{\omega_{d}}{2}\sigma_{z}=-\frac{\Delta}{2}\sigma_{z}+\frac{\Omega}{2}(\sigma_{+}+\sigma_{-})+\frac{\Omega}{2}(e^{-2i\omega_{d}t}\sigma_{+}+e^{2i\omega_{d}t}\sigma_{-})\leavevmode\nobreak\ , (57)

where we used that eitωdσz/2σ±eitωdσz/2=eitωdσ±e^{-it\omega_{d}\sigma_{z}/2}\sigma_{\pm}e^{it\omega_{d}\sigma_{z}/2}=e^{\mp it\omega_{d}}\sigma_{\pm}, σ+=|eg|\sigma_{+}=\left|e\right>\left<g\right|, σ=|ge|\sigma_{-}=\left|g\right>\left<e\right|, and Δ=ωωd\Delta=\omega-\omega_{d}.

The rotating wave approximation consists of neglecting the rapidly oscillating terms in the Hamiltonian, which average out over time and have negligible effect on the system’s dynamics. Hence, taking Ω2(e2iωdtσ++e+2iωdtσ)0\frac{\Omega}{2}(e^{-2i\omega_{d}t}\sigma_{+}+e^{+2i\omega_{d}t}\sigma_{-})\approx 0, the Hamiltonian becomes

H=Δ2σz+λσx,H^{\prime}=-\frac{\Delta}{2}\sigma_{z}+\lambda\sigma_{x}, (58)

where λΩ/2\lambda\equiv\Omega/2, and σx=σ++σ\sigma_{x}=\sigma_{+}+\sigma_{-}. To return to the Schrodinger picture, we need to apply the inverse transformation, where

ρ=eitωd2σzρeitωd2σz.\rho=e^{\frac{it\omega_{d}}{2}\sigma_{z}}\rho^{\prime}e^{\frac{-it\omega_{d}}{2}\sigma_{z}}. (59)

E.2 Inversion protocol

In the main text, we considered a two-level system of energy ω\omega coupled to a thermal bath under the action of feedback. By employing a time-dependent, jump-based feedback protocol described by Result 2, we demonstrated that it is possible not only to increase the excited state population compared to the thermal case without feedback, but also to achieve population inversion, with Pe>PgP_{e}>P_{g}. In this section, we provide a step-by-step derivation of the equations governing the populations under feedback. We consider continuous monitoring of quantum jumps, described by the jump operators

L=γ(N¯+1)|ge|,L+=γN¯|eg|,L_{-}=\sqrt{\gamma(\bar{N}+1)}\leavevmode\nobreak\ |g\rangle\langle e|,\quad L_{+}=\sqrt{\gamma\bar{N}}\leavevmode\nobreak\ |e\rangle\langle g|, (60)

which correspond to the emission (xn=1x_{n}=-1) and absorption (xn=+1x_{n}=+1) of a quantum, respectively. The outcome xn=0x_{n}=0 represents a no-jump detection.

The feedback action is defined as follows: if an emission is detected, we turn on an external drive with frequency ωd\omega_{d} for a total duration τ1\tau_{1}, after a time delay τ0\tau_{0} without absorption. During this feedback window, and working in a rotating frame under the rotating wave approximation, the Hamiltonian becomes

Hon=Δ2σz+λσx,H_{\text{on}}=-\frac{\Delta}{2}\sigma_{z}+\lambda\sigma_{x}\leavevmode\nobreak\ , (61)

where Δ=ωωd\Delta=\omega-\omega_{d}, otherwise we remove the drive (λ0\lambda\rightarrow 0), and we obtain the thermal qubit with Hamiltonian

Hoff=Δ2σz.H_{\text{off}}=-\frac{\Delta}{2}\sigma_{z}\leavevmode\nobreak\ . (62)

Since this feedback protocol affects only the system’s Hamiltonian, then the unconditional state evolves according to

ρ¯t=yΣ0t𝑑τ𝒯[e0τ𝑑s0(y,s)]𝒥yρ¯tτ,\bar{\rho}_{t}=\sum_{y\in\Sigma}\int_{0}^{t}d\tau\mathcal{T}\left[e^{\int_{0}^{\tau}ds\mathcal{L}_{0}(y,s)}\right]\mathcal{J}_{y}\bar{\rho}_{t-\tau}\leavevmode\nobreak\ , (63)

and the steady-state, if it exists, is the solution of the algebraic equation

ρ¯ss=Ωρ¯ss,Ω=y{+1,1}0𝑑τ𝒯[e0τ𝑑s0(y,s)]𝒥y.\bar{\rho}_{\rm ss}=\Omega\bar{\rho}_{\rm ss},\qquad\Omega=\sum_{y\in\{+1,-1\}}\int_{0}^{\infty}d\tau\mathcal{T}\left[e^{\int_{0}^{\tau}ds\mathcal{L}_{0}(y,s)}\right]\mathcal{J}_{y}. (64)

In order to verify whether a steady state exists, it is sufficient to show that the no-jump Liouvillian has only eigenvalues with negative real parts. The reasoning is as follows: if 0(y,τ)\mathcal{L}_{0}(y,\tau) has only negative eigenvalues, then the term 𝒯[e0τ𝑑s0(y,s)]\mathcal{T}\left[e^{\int_{0}^{\tau}ds\,\mathcal{L}_{0}(y,s)}\right] vanishes as τ\tau\rightarrow\infty. Therefore, in the steady-state limit of Eq. (63), when tt\rightarrow\infty, only contributions with finite τ\tau survive in 𝒯[e0τ𝑑s0(y,s)]\mathcal{T}\left[e^{\int_{0}^{\tau}ds\,\mathcal{L}_{0}(y,s)}\right], and we can replace ρtτρss\rho_{t-\tau}\rightarrow\rho_{ss} inside the integral. For this feedback protocol, the no-jump Liouvillian are given by

0(y,s)={0onif y=1 and s[τ0,τ0+τ1],0offotherwise ,\mathcal{L}_{0}(y,s)=\begin{cases}\mathcal{L}_{0}^{\text{on}}&\text{if }y=1\text{ and }s\in[\tau_{0},\tau_{0}+\tau_{1}],\\ \mathcal{L}_{0}^{\text{off}}&\text{otherwise },\end{cases} (65)

where 0off\mathcal{L}_{0}^{\text{off}} and 0on\mathcal{L}_{0}^{\text{on}} are the no-jump Liouvillian related with HoffH_{\text{off}} and HonH_{\text{on}}, respectively. By applying the vectorization technic Landi et al. (2024), the 2×22\times 2 density states ρ\rho become a column vector of 44 entries, and the Liouvillians become a 4×44\times 4 matrix that act on these vectors. By considering a resonant external drive (Δ=0\Delta=0), the no-jumps Liouvillians are given by

0off\displaystyle\mathcal{L}_{0}^{\text{off}} =(N¯γ000012(1+2N¯)γ000012(1+2N¯)γ0000(1+N¯)γ),\displaystyle=\begin{pmatrix}-\bar{N}\gamma&0&0&0\\ 0&-\frac{1}{2}(1+2\bar{N})\gamma&0&0\\ 0&0&-\frac{1}{2}(1+2\bar{N})\gamma&0\\ 0&0&0&-(1+\bar{N})\gamma\end{pmatrix}, (66)
0on\displaystyle\mathcal{L}_{0}^{\text{on}} =(N¯γiλiλ0iλ12(1+2N¯)γ0iλiλ012(1+2N¯)γiλ0iλiλ(1+N¯)γ),\displaystyle=\begin{pmatrix}-\bar{N}\gamma&-i\lambda&i\lambda&0\\ -i\lambda&-\frac{1}{2}(1+2\bar{N})\gamma&0&i\lambda\\ i\lambda&0&-\frac{1}{2}(1+2\bar{N})\gamma&-i\lambda\\ 0&i\lambda&-i\lambda&-(1+\bar{N})\gamma\end{pmatrix}\leavevmode\nobreak\ , (67)

and we can verify that both of them have eigenvalues with negative real part. Hence, the steady state exists and it is given by Eq. (64). We can compute Ω\Omega directly from the definition by using Eq. (65), and one finds

Ω=[0off]1𝒥1+{[0off]1+([0off]1[0on]1)eτ00off+([0on]1[0off]1)eτ10oneτ00off}𝒥1.\Omega=-[\mathcal{L}_{0}^{\text{off}}]^{-1}\mathcal{J}_{1}+\left\{-\left[\mathcal{L}_{0}^{\text{off}}\right]^{-1}+\left(\left[\mathcal{L}_{0}^{\text{off}}\right]^{-1}-\left[\mathcal{L}_{0}^{\text{on}}\right]^{-1}\right)e^{\tau_{0}\mathcal{L}_{0}^{\text{off}}}+\left(\left[\mathcal{L}_{0}^{\text{on}}\right]^{-1}-\left[\mathcal{L}_{0}^{\text{off}}\right]^{-1}\right)e^{\tau_{1}\mathcal{L}_{0}^{\text{on}}}e^{\tau_{0}\mathcal{L}_{0}^{\text{off}}}\right\}\mathcal{J}_{-1}\leavevmode\nobreak\ . (68)

Finally, since we have the operator Ω\Omega, the steady state of the system is its eigenvector with eigenvalue 1, and the population of the ground and excited state are the diagonal elements of ρss\rho_{ss}. Furthermore, the steady state found as a eigenvector of Ω\Omega belongs to the rotating frame picture. To return to the Schrodinger picture, we can use Eq. (59).

The population of the excited state in the Schrodinger picture is then given by Pe=11+ξP_{e}=\frac{1}{1+\xi}, and one finds that

ξ=(N¯+1)N¯uωr2δ(uωr2δ4(1+N¯)λ2ωr2eN¯τ0γψ),\xi=\frac{(\bar{N}+1)}{\bar{N}\,u\,\omega_{r}^{2}\,\delta}\Bigg{(}u\,\omega_{r}^{2}\,\delta-4(1+\bar{N})\lambda^{2}\omega_{r}^{2}e^{-\bar{N}\tau_{0}\gamma}-\psi\Bigg{)}\leavevmode\nobreak\ , (69)

where

ψ=2λ2e(N¯τ0+12uτ1)γ[4δ+8uλ2(eτ1ωr2+eτ1ωr2)(N¯+1)γu(ωr(eτ1ωr2eτ1ωr2)+γ(eτ1ωr2+eτ1ωr2))],\psi=2\lambda^{2}e^{-(\bar{N}\tau_{0}+\tfrac{1}{2}u\tau_{1})\gamma}\left[4\delta+8u\lambda^{2}\left(e^{\frac{\tau_{1}\omega_{r}}{2}}+e^{-\frac{\tau_{1}\omega_{r}}{2}}\right)-(\bar{N}+1)\gamma u\left(\omega_{r}\left(e^{\frac{\tau_{1}\omega_{r}}{2}}-e^{-\frac{\tau_{1}\omega_{r}}{2}}\right)+\gamma\left(e^{\frac{\tau_{1}\omega_{r}}{2}}+e^{-\frac{\tau_{1}\omega_{r}}{2}}\right)\right)\right]\leavevmode\nobreak\ , (70)

and

u\displaystyle u =\displaystyle= 2N¯+1,\displaystyle 2\bar{N}+1\leavevmode\nobreak\ , (71)
δ\displaystyle\delta =\displaystyle= N¯(N¯+1)γ2+4λ2,\displaystyle\bar{N}(\bar{N}+1)\gamma^{2}+4\lambda^{2}\leavevmode\nobreak\ , (72)
ωr\displaystyle\omega_{r} =\displaystyle= γ216λ2.\displaystyle\sqrt{\gamma^{2}-16\lambda^{2}}\leavevmode\nobreak\ . (73)

Note that when γ4λ\gamma\geq 4\lambda, then ωr\omega_{r}\in\mathbb{R}, and the exponential terms become hyperbolic functions. On the other hand, when γ<4λ\gamma<4\lambda, then ωr=i16λ2γ2\omega_{r}=i\sqrt{16\lambda^{2}-\gamma^{2}} is a pure imaginary number, and these terms become a sine and cosine.

Now, let us outline the derivation of the main results for the example presented in the main text. First, we compute the optimal time τ1opt\tau_{1}^{\text{opt}} during which the external drive should remain on to maximize PeP_{e}. Maximizing PeP_{e} is equivalent to minimizing ξ\xi. Note that the only term in ξ\xi that depends on τ1\tau_{1} is proportional to ψ\psi. Therefore, by solving the equation τ1ψ=0\partial_{\tau_{1}}\psi=0, we obtain the optimal duration τ1opt\tau_{1}^{\text{opt}} as a function of N¯\bar{N}, γ\gamma, λ\lambda, and τ0\tau_{0}. Once τ1opt\tau_{1}^{\text{opt}} is determined, we set τ1=τ1opt\tau_{1}=\tau_{1}^{\text{opt}} in all subsequent steps. Surprisingly, we found that the optimal duration of the drive depends only on the ratio γ/λ\gamma/\lambda, and remains the same regardless of the temperature or the feedback time delay.

We can also determine the population inversion points by solving the equation ξ=1\xi=1, which corresponds to the condition Pe=Pg=1/2P_{e}=P_{g}=1/2. The solution of this equation yields the critical value N¯c\bar{N}_{c} (or, equivalently, the critical temperature) for a given ratio γ/λ\gamma/\lambda, such that Pe>PgP_{e}>P_{g} for any N¯<N¯c\bar{N}<\bar{N}_{c}. We show numerically that in the ideal case with zero delay (τ0=0\tau_{0}=0), the equation ξ=1\xi=1 admits a solution with N¯c0\bar{N}_{c}\geq 0 only when γ/λ<1.145\gamma/\lambda<1.145.

Finally, to analyze the effect of feedback time delay on PeP_{e}, we consider the low-temperature limit N¯0\bar{N}\rightarrow 0, effectively removing thermal effects. We then compute the maximum time delay τ0max\tau_{0}^{\text{max}} by solving the equation limN¯0ξ=1\lim_{\bar{N}\rightarrow 0}\xi=1 for τ0\tau_{0}, given γ\gamma, λ\lambda, and τ1opt\tau_{1}^{\text{opt}}. This yields the maximum delay τ0max\tau_{0}^{\text{max}} for which Pe=PgP_{e}=P_{g}. Consequently, for any τ0<τ0max\tau_{0}<\tau_{0}^{\text{max}}, we have Pe>PgP_{e}>P_{g} provided γ/λ<1.145\gamma/\lambda<1.145.

Appendix F Deriving previous results as particular cases of the Result 1

In this section, we will show how the previous results Annby-Andersson et al. (2022); Wiseman and Milburn (1993); Wiseman (1994b); Kewming et al. (2024) can be seen as particular cases of the Result 1, given by the deterministic equation

ϱn+1(y)=x,yδy,fn+1(x,y)x(y)ϱn(y),\varrho_{n+1}(y)=\sum_{x^{\prime},y^{\prime}}\delta_{y,f_{n+1}(x^{\prime},y^{\prime})}\mathcal{M}_{x^{\prime}}(y^{\prime})\varrho_{n}(y^{\prime})\leavevmode\nobreak\ , (74)

for a given stochastic signal yny_{n} that satisfies the causality equation yn+1=fn(xn+1,yn)y_{n+1}=f_{n}(x_{n+1},y_{n}) for some function fn(x,y)f_{n}(x,y), where xnx_{n} is the outcome of the detection at time tn=nδtt_{n}=n\delta t. In each case, the feedback-measurement protocol is fixed by defining the instruments x\mathcal{M}_{x} and the adequate stochastic signal, as described in Table 1.

F.1 Low-pass signal and weak Gaussian measurements: Fokker-Planck master equation

Let us consider a sequential measurement described by the instruments x\mathcal{M}_{x}, as described in the main text. We will introduce the low-pass signal yny_{n} given by

yn=j=1nαpnjxj,y_{n}=\sum_{j=1}^{n}\alpha p^{n-j}x_{j}\leavevmode\nobreak\ , (75)

where α\alpha and pp are arbitrary parameters, and xnx_{n} is the outcome of the measurement at time tn=nδtt_{n}=n\delta t. From this definition, we can see that

yn=αzn+pyn1,y_{n}=\alpha z_{n}+py_{n-1}, (76)

then the low-pass signal satisfies the causality condition yn=f(xn,yn1)y_{n}=f(x_{n},y_{n-1}), where f(x,y)=αx+pyf(x,y)=\alpha x+py. From the Result 1, the corresponding deterministic equation of the signal-resolved state is given by

ϱn+1(y)=x,yδy,αx+pyx(y)ρn(y).\displaystyle\varrho_{n+1}(y)=\sum_{x^{\prime},y^{\prime}}\delta_{y,\alpha x^{\prime}+py^{\prime}}\mathcal{M}_{x^{\prime}}(y^{\prime})\rho_{n}(y^{\prime})\leavevmode\nobreak\ . (77)

For p0p\neq 0, we can solve the sum over yy^{\prime} by using the Kronecker delta, and the deterministic equation becomes

ϱn(y)=xx((yαx)/p)ϱn1((yαx)/p).\varrho_{n}(y)=\sum_{x}\mathcal{M}_{x}((y-\alpha x)/p)\varrho_{n-1}\big{(}(y-\alpha x)/p\big{)}. (78)

It is worth to mention that the case p=0p=0, where the signal corresponds to the actual detection, will be considered soon. Therefore, Eq. (78) describes a general feedback-measurement protocol where the stochastic signal is given by Eq. (75).

From now on, let us consider a sequential measurement described by the set of Kraus operators {Kx}x\{K_{x}\}_{x\in\mathbb{R}}, where

Kx=(2λδtπ)14eλδt(xA)2,K_{x}=\left(\frac{2\lambda\delta t}{\pi}\right)^{\frac{1}{4}}e^{-\lambda\delta t(x-A)^{2}}\leavevmode\nobreak\ , (79)

AA is a given observable of the system, and λ\lambda is the measurement strength Jacobs and Steck (2006); Bednorz et al. (2012). These operators KxK_{x} describe a weak Gaussian measurement of AA, and the corresponding instrument is defined as ~zρKzρKz\widetilde{\mathcal{M}}_{z}\rho\equiv K_{z}\rho K_{z}^{\dagger}. Now, let us consider that the system will evolve dynamically according to a Liouvillian \mathcal{L} between two detections. Hence, the total instrument z\mathcal{M}_{z} that describes this scenario is defined as

z(yn)~zeδt(yn),\mathcal{M}_{z}(y_{n})\equiv\widetilde{\mathcal{M}}_{z}e^{\delta t\mathcal{L}(y_{n})}\leavevmode\nobreak\ , (80)

where the feedback encoded in the stochastic signal yny_{n} can affect the dynamic of the system through (yn)\mathcal{L}(y_{n}). Therefore, from Eq. (78), we have

ϱn(y)=+𝑑x~xe(yαxp)δtϱn1(yαxp),\varrho_{n}(y)=\int_{-\infty}^{+\infty}dx\leavevmode\nobreak\ \widetilde{\mathcal{M}}_{x}e^{\mathcal{L}\left(\frac{y-\alpha x}{p}\right)\delta t}\varrho_{n-1}\left(\frac{y-\alpha x}{p}\right)\leavevmode\nobreak\ , (81)

where the sum is transformed into an integral because the outcomes xx are real numbers.

In the next step we will take the limit δt0\delta t\rightarrow 0 in the Eq. (81). Let us consider the change of variables defined as ξ(yαx)/p\xi\equiv(y-\alpha x)/p, and then x=(ypξ)/αx=(y-p\xi)/\alpha. Also, we will define p=exp((γδt))p=\exp{(-\gamma\delta t)} and α=γδt\alpha=\gamma\delta t. Then dξ=dx/(γδt)d\xi=-dx/(\gamma\delta t). Therefore, the deterministic evolution becomes

ϱn(y)=+𝑑ξ1γδt~(yeγδtξγδt)e(ξ)δtϱn1(ξ).\varrho_{n}(y)=\int_{-\infty}^{+\infty}d\xi\leavevmode\nobreak\ \frac{1}{\gamma\delta t}\leavevmode\nobreak\ \widetilde{\mathcal{M}}\left(\frac{y-e^{-\gamma\delta t}\xi}{\gamma\delta t}\right)e^{\mathcal{L}\left(\xi\right)\delta t}\varrho_{n-1}\left(\xi\right)\leavevmode\nobreak\ . (82)

Now, we can define

Ω(y|ξ)1γδt~(yeγδtξγδt),\Omega(y|\xi)\equiv\frac{1}{\gamma\delta t}\leavevmode\nobreak\ \widetilde{\mathcal{M}}\left(\frac{y-e^{-\gamma\delta t}\xi}{\gamma\delta t}\right)\leavevmode\nobreak\ , (83)

and then

ϱn(y)=+𝑑ξΩ(y|ξ)e(ξ)δtϱn1(ξ).\varrho_{n}(y)=\int_{-\infty}^{+\infty}d\xi\leavevmode\nobreak\ \Omega(y|\xi)\leavevmode\nobreak\ e^{\mathcal{L}\left(\xi\right)\delta t}\varrho_{n-1}\left(\xi\right)\leavevmode\nobreak\ . (84)

As shown in Annby-Andersson et al. (2022) (see supplemental material, Eq. (S7)), we can expand the propagator Ω(y|ξ)e(ξ)δt\Omega(y|\xi)\leavevmode\nobreak\ e^{\mathcal{L}\left(\xi\right)\delta t} in first order of δt\delta t, resulting in the following expression

Ω(y|ξ)ρ=δ(yξ)+δt[λδ(yξ)𝒟[A]γ𝒜(ξ)ρδ(yξ)+γ28λδ′′(yξ)ρ]+𝒪(δt2).\Omega(y|\xi)\rho=\delta(y-\xi)+\delta t\left[\lambda\delta(y-\xi)\mathcal{D}[A]-\gamma\mathcal{A}(\xi)\rho\delta^{\prime}(y-\xi)+\frac{\gamma^{2}}{8\lambda}\delta^{\prime\prime}(y-\xi)\rho\right]\\ +\leavevmode\nobreak\ \mathcal{O}(\delta t^{2})\leavevmode\nobreak\ . (85)

where 𝒟[A]ρ=AρA1/2{AA,ρ}\mathcal{D}[A]\rho=A\rho A^{\dagger}-1/2\{A^{\dagger}A,\rho\} is the Lindblad dissipator, and 𝒜(ξ)ρ:=12{Aξ,ρ}\mathcal{A}(\xi)\rho:=\frac{1}{2}\{A-\xi,\rho\}. Furthermore, δ(x)\delta^{\prime}(x) and δ′′(x)\delta^{\prime\prime}(x) denote the first and second derivatives of the Dirac distribution δ(x)\delta(x), respectively. Therefore, applying this first order expansion of the propagator Ω(y|ξ)\Omega(y|\xi) in Eq. (84) and taking the limit δt0\delta t\rightarrow 0, we finally have

tϱt(y)=(y)ϱt(y)+λ𝒟[A]ϱt(t)γξ{𝒜(ξ)ϱt(ξ)}|ξ=y+γ28λξ2ϱt(ξ)|ξ=y,\partial_{t}\varrho_{t}(y)=\mathcal{L}(y)\varrho_{t}(y)+\lambda\mathcal{D}[A]\varrho_{t}(t)-\gamma\partial_{\xi}\{\mathcal{A}(\xi)\varrho_{t}(\xi)\}|_{\xi=y}+\frac{\gamma^{2}}{8\lambda}\partial_{\xi}^{2}\varrho_{t}(\xi)|_{\xi=y}\leavevmode\nobreak\ , (86)

corresponding to the Fokker-Plank master equation of the signal-resolved state Annby-Andersson et al. (2022). Note that, in the limit δt0\delta t\rightarrow 0, Eq. (75) becomes

yt=0t𝑑sγeγ(ts)xt,y_{t}=\int_{0}^{t}ds\gamma e^{-\gamma(t-s)}x_{t}\leavevmode\nobreak\ , (87)

where xtx_{t} is the outcome of the weak Gaussian measurement at time tt. Therefore, we showed that Eq. (86) can be derived from Result 1 considering a sequential weak Gaussian measurement and a feedback-measurement protocol based on the low-pass signal.

F.2 Feedback based on the most recent detection

In the context of sequential detection, suppose the feedback depends solely on the most recent detection. This is equivalent to assuming that the stochastic signal is given by the current detection, yn=xny_{n}=x_{n}. It may be interpreted as a particular case of the low-pass signal with p=0p=0. Since this feedback protocol does not rely on the prior history of the experiment, it can be regarded as a memory-less feedback scheme. In this scenario, we have yn=f(xn,yn1)y_{n}=f(x_{n},y_{n-1}) with f(x,y)=xf(x,y)=x. According to Result 1, the deterministic equation governing the signal-resolved state is then given by

ϱn+1(y)=x,yδy,xx(y)ϱn(y),\varrho_{n+1}(y)=\sum_{x^{\prime},y^{\prime}}\delta_{y,x^{\prime}}\mathcal{M}_{x^{\prime}}(y^{\prime})\varrho_{n}(y^{\prime})\leavevmode\nobreak\ , (88)

and by performing the sum over xx^{\prime}, we obtain

ϱn+1(y)=yy(y)ϱn(y),\varrho_{n+1}(y)=\sum_{y^{\prime}}\mathcal{M}_{y}(y^{\prime})\varrho_{n}(y^{\prime})\leavevmode\nobreak\ , (89)

where yy denotes a possible detection outcome at time tn+1t_{n+1}, and the sum over yy^{\prime} accounts for all possible outcomes at time tnt_{n}. Equation (89) therefore describes a general feedback-measurement protocol based solely on the most recent detection. In what follows, we will explore how this equation applies to specific measurement schemes, thereby recovering earlier results.

F.2.1 Quantum Jumps

Let us consider the application of a quantum channel to the system’s state conditioned on quantum jump detection. The conditional stroboscopic evolution in this feedback-measurement scenario is described by

ρn=(x)[Vxρn1VxTr[Vxρn1Vx]],\rho_{n}=\mathcal{F}(x)\left[\frac{V_{x}\rho_{n-1}V_{x}^{\dagger}}{\mathrm{Tr}[V_{x}\rho_{n-1}V_{x}^{\dagger}]}\right]\leavevmode\nobreak\ , (90)

where the Kraus operators {Vx}x\{V_{x}\}_{x} describe the quantum jump detection process, and the quantum channel (x)\mathcal{F}(x), referred to as the feedback super-operator, defines the feedback protocol based on the most recent quantum jump Wiseman (1994a). The corresponding instrument for this sequential feedback-measurement protocol is given by xρ=(x)[Vx(ρ)Vx]\mathcal{M}_{x}\rho=\mathcal{F}(x)[V_{x}(\rho)V_{x}^{\dagger}], where x=0x=0 represents a no-jump detection, and xΣx\in\Sigma corresponds to a quantum jump. Then, using Eq. (89), we obtain the deterministic equation for the signal-resolved state:

ϱn(y)=x(y)Vyϱn1(x)Vy=(y)Vy[xϱn1(x)]Vy=(y)Vyρ¯nVy,\varrho_{n}(y)=\sum_{x}\mathcal{F}(y)V_{y}\varrho_{n-1}(x)V_{y}^{\dagger}=\mathcal{F}(y)V_{y}\left[\sum_{x}\varrho_{n-1}(x)\right]V_{y}^{\dagger}=\mathcal{F}(y)V_{y}\bar{\rho}_{n}V_{y}^{\dagger}\leavevmode\nobreak\ , (91)

where we have used the identity ρ¯n=xϱn(x)\bar{\rho}_{n}=\sum_{x}\varrho_{n}(x). Summing this equation over yy, we finally obtain

ρ¯n=y(y)Vyρ¯n1Vy.\bar{\rho}_{n}=\sum_{y}\mathcal{F}(y)V_{y}\bar{\rho}_{n-1}V_{y}^{\dagger}. (92)

Now, let us consider the following feedback super-operator

(k)={𝟙if k=0 (no-jump),e𝒦(k)if kΣ (jump k),\mathcal{F}(k)=\begin{cases}\mathbb{1}&\text{if }k=0\text{ (no-jump)},\\ e^{\mathcal{K}(k)}&\text{if }k\in\Sigma\text{ (jump $k$)},\end{cases} (93)

where 𝒦\mathcal{K} is an arbitrary Liouvillian super-operator. With this definition, we can write the evolution of the uncoditional state as

ρ¯n=V0ρ¯n1V0+yΣ(y)Vyρ¯n1Vy.\bar{\rho}_{n}=V_{0}\bar{\rho}_{n-1}V_{0}^{\dagger}+\sum_{y\in\Sigma}\mathcal{F}(y)V_{y}\bar{\rho}_{n-1}V_{y}^{\dagger}. (94)

Finally, using the definition of the Kraus operators VkV_{k} from the unraveling of the quantum master equation, and taking the limit δt0\delta t\rightarrow 0, we obtain the following master equation

tρ¯t=i[H,ρ¯t]+αΣ(e𝒦(α)[Lkρ¯tLk]12{LkLk,ρ¯t}),\displaystyle\partial_{t}\bar{\rho}_{t}=-i[H,\bar{\rho}_{t}]+\sum_{\alpha\in\Sigma}\left(e^{\mathcal{K}(\alpha)}\left[L_{k}\bar{\rho}_{t}L_{k}^{\dagger}\right]-\frac{1}{2}\{L_{k}^{\dagger}L_{k},\bar{\rho}_{t}\}\right)\leavevmode\nobreak\ , (95)

which corresponds to the well-known master equation for feedback based on the most recent jump, as developed in Wiseman (1994a).

F.2.2 Diffusion

Let us consider another scenario where the feedback is based solely on the most recent measurement outcome, but now with diffusion. Considering a weak continuous Gaussian measurement of an hermitian observable AA described by the Kraus operators defined in Eq. (79), the outcome of the measurements follows a stochastic equation given by

z(t)=12λδtdW(t)+Ac,z(t)=\frac{1}{2\sqrt{\lambda}\delta t}dW(t)+\left<A\right>_{c}\leavevmode\nobreak\ , (96)

where Ac=Tr[Aρc(t)]\left<A\right>_{c}=\mathrm{Tr}[A\rho_{c}(t)], dWdW is a Wiener increment, and ρc(t)\rho_{c}(t) is the conditional state at time tt. The stochastic master equation for ρc\rho_{c} is given by Jacobs and Steck (2006)

dρc(t)=δtρc(t)+λδt𝒟[A]ρc(t)+λdW{AAc,ρc(t)}.d\rho_{c}(t)=\delta t\mathcal{L}\rho_{c}(t)+\lambda\delta t\mathcal{D}[A]\rho_{c}(t)+\sqrt{\lambda}dW\{A-\left<A\right>_{c},\rho_{c}(t)\}. (97)

In this case, the feedback protocol is defined as follows: the system evolves dynamically from tn1=(n1)δtt_{n-1}=(n-1)\delta t to tn=nδtt_{n}=n\delta t, and we apply a weak Gaussian measurement of AA at this point, resulting in the outcome ZnZ_{n}. Hence, the feedback will be based on the most recent outcome of the detection through the feedback super-operator (z)\mathcal{F}(z). For simplicity, we will consider a unitary evolution.

The total instrument that describes this discrete evolution is given by (z)~zeδt\mathcal{F}(z)\widetilde{\mathcal{M}}_{z}e^{\delta t\mathcal{L}}, where ρ=i[H,ρ]\mathcal{L}\rho=-i[H,\rho] represents the unitary evolution between tn1t_{n-1} and tnt_{n}, ~zρKzρKz\widetilde{\mathcal{M}}_{z}\rho\equiv K_{z}\rho K_{z}^{\dagger} represents the measurement process, and (z)\mathcal{F}(z) is the feedback super-operator based on the measurement outcome. Following a similar approach as we developed for quantum jumps in Eq. (92), the deterministic equation for the unconditional state in this case is given by

ρ¯n=𝑑z(z)~zeδtρ¯n1,\bar{\rho}_{n}=\int_{-\infty}^{\infty}dz\mathcal{F}(z)\widetilde{\mathcal{M}}_{z}e^{\delta t\mathcal{L}}\bar{\rho}_{n-1}, (98)

where the sum was replaced by a integral since the outcomes of the Gaussian measurements are continuous real numbers.

Now, let us consider a particular feedback super-operator. Following the reference Wiseman (1994a), let us define

(z)=e2λzδt𝒦,\mathcal{F}(z)=e^{2\sqrt{\lambda}z\delta t\mathcal{K}}\leavevmode\nobreak\ , (99)

where 𝒦ρi[F,ρ]\mathcal{K}\rho\equiv-i[F,\rho] is a super-operator that represents a reversible transformation for a given hermitian operator FF. Hence, the deterministic equation becomes

ρ¯n=+𝑑ze2λzδt𝒦~zeδtρ¯n1.\bar{\rho}_{n}=\int_{-\infty}^{+\infty}dze^{2\sqrt{\lambda}z\delta t\mathcal{K}}\widetilde{\mathcal{M}}_{z}e^{\delta t\mathcal{L}}\bar{\rho}_{n-1}\leavevmode\nobreak\ . (100)

The next step corresponds to expand the deterministic equation on first order of δt\delta t. For the feedback super-operator, we have Wiseman (1994a)

(z)=1+2λz𝒦δt+𝒦22δt,\mathcal{F}(z)=1+2\sqrt{\lambda}z\mathcal{K}\delta t+\frac{\mathcal{K}^{2}}{2}\delta t\leavevmode\nobreak\ , (101)

and the unitary evolutions yields

eδtρ(𝟙+δt)ρ=ρiδt[H,ρ].e^{\delta t\mathcal{L}}\rho\approx(\mathbb{1}+\delta t\mathcal{L})\rho=\rho-i\delta t[H,\rho]\leavevmode\nobreak\ . (102)

Also, we can show that

~zρ=KzρKz=+dω2πeω28λδteiωz[eiωA2ρeiωA2λδt2(A2eiωA2ρeiωA2+eiωA2ρA2eiωA2AeiωA2ρAeiωA2)],\widetilde{\mathcal{M}}_{z}\rho=K_{z}\rho K_{z}^{\dagger}=\int_{-\infty}^{+\infty}\frac{d\omega}{2\pi}e^{\frac{-\omega^{2}}{8\lambda\delta t}}e^{-i\omega z}\left[e^{\frac{i\omega A}{2}}\rho e^{\frac{i\omega A}{2}}-\frac{\lambda\delta t}{2}\left(A^{2}e^{\frac{i\omega A}{2}}\rho e^{\frac{i\omega A}{2}}+e^{\frac{i\omega A}{2}}\rho A^{2}e^{\frac{i\omega A}{2}}-Ae^{\frac{i\omega A}{2}}\rho Ae^{\frac{i\omega A}{2}}\right)\right]\leavevmode\nobreak\ , (103)

and keeping only terms of order δt\delta t in the instrument e2λzδt𝒦~zeδte^{2\sqrt{\lambda}z\delta t\mathcal{K}}\widetilde{\mathcal{M}}_{z}e^{\delta t\mathcal{L}}, the deterministic equation becomes

ρ¯n=ρ¯n1+λδt𝒟[A]ρ¯n1iδt[H,ρ¯n1]+δt𝒟[F]ρ¯n1λδti[F,Aρ¯n1+ρ¯n1A].\bar{\rho}_{n}=\bar{\rho}_{n-1}+\lambda\delta t\mathcal{D}[A]\bar{\rho}_{n-1}-i\delta t[H,\bar{\rho}_{n-1}]+\delta t\mathcal{D}[F]\bar{\rho}_{n-1}-\sqrt{\lambda}\delta ti[F,A\bar{\rho}_{n-1}+\bar{\rho}_{n-1}A]\leavevmode\nobreak\ . (104)

Finally, taking the limit δt0\delta t\rightarrow 0 and writing A=AA=A^{\dagger}, one finds

tρ¯t=i[H,ρ¯t]+𝒟[λA]ρ¯t+𝒟[F]ρ¯ti[F,λAρ¯t+ρ¯tλA].\partial_{t}\bar{\rho}_{t}=-i[H,\bar{\rho}_{t}]+\mathcal{D}\left[\sqrt{\lambda}A\right]\bar{\rho}_{t}+\mathcal{D}[F]\bar{\rho}_{t}-i[F,\sqrt{\lambda}A\bar{\rho}_{t}+\bar{\rho}_{t}\sqrt{\lambda}A^{\dagger}]\leavevmode\nobreak\ . (105)

Eq. (105) coincides with the feedback master equation found in Wiseman and Milburn (1993) for ideal homodyne detection. In that case, one must make the identification λAL\sqrt{\lambda}A\rightarrow L, where LL is a (generally non-Hermitian) jump operator. In homodyne detection, the diffusion arises from continuous measurements of the quadrature operator x=L+Lx=L+L^{\dagger}. Therefore, our result recovers the feedback master equation of Wiseman and Milburn (1993), corresponding to diffusion induced by weak Gaussian measurements of a Hermitian operator AA.

F.3 Charge-based feedback

Let us consider a continuous monitoring of quantum jumps, as defined in the main text. In this case, xnx_{n} represents the outcome of the quantum jump detection at time tn=nδtt_{n}=n\delta t, where xn=0x_{n}=0 is a no-jump detection, and xn=kΣx_{n}=k\in\Sigma is a jump in the channel kk. Let us introduce the following random variable

dNk,nδxn,k,dN_{k,n}\equiv\delta_{x_{n},k}\leavevmode\nobreak\ , (106)

where δa,b\delta_{a,b} is the Kronecker delta. It implies that dNk,n=1dN_{k,n}=1 if we have a jump in the channel kk at time tnt_{n}, and dNk,n=0dN_{k,n}=0 otherwise. Furthermore, the total stochastic number of jumps in the channel kk until the time tnt_{n} is

Nk,nj=1ndNk,j=j=1nδxj,k.N_{k,n}\equiv\sum_{j=1}^{n}dN_{k,j}=\sum_{j=1}^{n}\delta_{x_{j},k}\leavevmode\nobreak\ . (107)

Hence, we can define the stochastic charge as

NnkΣνkNk,n=kΣj=1nνkδxj,k,N_{n}\equiv\sum_{k\in\Sigma}\nu_{k}N_{k,n}=\sum_{k\in\Sigma}\sum_{j=1}^{n}\nu_{k}\delta_{x_{j},k}\leavevmode\nobreak\ , (108)

corresponding to a linear combination of the total number of jumps in each channel kk. In the limit δt0\delta t\rightarrow 0, we have

N(t)=kΣνk0t𝑑tδ(x(t)k),N(t)=\sum_{k\in\Sigma}\nu_{k}\int_{0}^{t}dt^{\prime}\delta(x(t^{\prime})-k)\leavevmode\nobreak\ , (109)

where δ(x)\delta(x) denotes the Dirac delta. In this way, we have defined the stochastic charge in terms of the measurement outcomes xnx_{n} of the continuous monitoring of the quantum jumps. Note that this stochastic signal is causal, and it satisfies

Nn=kΣνkδxn,k+Nn1.N_{n}=\sum_{k\in\Sigma}\nu_{k}\delta_{x_{n},k}+N_{n-1}. (110)

Now, let us consider a feedback protocol where the detections are based on the continuous monitoring of the quantum jumps, and the stochastic signal is the stochastic charge (108). In this case, the instruments are given by Landi et al. (2024)

0(N)ρ\displaystyle\mathcal{M}_{0}(N)\rho =(1+0(N)δt)ρ,\displaystyle=(1+\mathcal{L}_{0}(N)\delta t)\rho, (111)
k(N)ρ\displaystyle\mathcal{M}_{k}(N)\rho =δt𝒥k(N)ρ,\displaystyle=\delta t\mathcal{J}_{k}(N)\rho, (112)

where 𝒥kρ=Lk(y)ρLk(y)\mathcal{J}_{k}\rho=L_{k}(y)\rho L_{k}^{\dagger}(y) and 0(y)=(y)kΣ𝒥k(y)\mathcal{L}_{0}(y)=\mathcal{L}(y)-\sum_{k\in\Sigma}\mathcal{J}_{k}(y). By applying Eq. (18) and (110), one finds

ϱn+1(N)\displaystyle\varrho_{n+1}(N) =\displaystyle= xNδN,kΣδx,k+Nx(N)ϱn(N)\displaystyle\sum_{x^{\prime}}\sum_{N^{\prime}}\delta_{N,\sum_{k\in\Sigma}\delta_{x^{\prime},k}+N^{\prime}}\mathcal{M}_{x^{\prime}}(N^{\prime})\varrho_{n}(N^{\prime}) (113)
=\displaystyle= NδN,N0(N)ϱn(N)+xΣNδN,νx+Nx(N)ϱn(N)\displaystyle\sum_{N^{\prime}}\delta_{N,N^{\prime}}\mathcal{M}_{0}(N^{\prime})\varrho_{n}(N^{\prime})+\sum_{x^{\prime}\in\Sigma}\sum_{N^{\prime}}\delta_{N,\nu_{x^{\prime}}+N^{\prime}}\mathcal{M}_{x^{\prime}}(N^{\prime})\varrho_{n}(N^{\prime}) (114)
=\displaystyle= 0(N)ϱn(N)+xΣx(Nνx)ϱn(Nνx).\displaystyle\mathcal{M}_{0}(N)\varrho_{n}(N)+\sum_{x^{\prime}\in\Sigma}\mathcal{M}_{x^{\prime}}(N-\nu_{x^{\prime}})\varrho_{n}(N-\nu_{x^{\prime}})\leavevmode\nobreak\ . (115)

Finally, using the definition of 0\mathcal{M}_{0} and k\mathcal{M}_{k}, and taking the limit of δt0\delta t\rightarrow 0, we have

tϱt(N)=0(N)ϱt(N)+kΣ𝒥k(Nνk)ϱt(Nνk).\partial_{t}\varrho_{t}(N)=\mathcal{L}_{0}(N)\varrho_{t}(N)+\sum_{k\in\Sigma}\mathcal{J}_{k}(N-\nu_{k})\varrho_{t}(N-\nu_{k})\leavevmode\nobreak\ . (116)

Eq. (116) was previously derived in Kewming et al. (2024), encompassing any feedback protocol based on charge detection, and we showed that this result follows directly from Eq. (18).