User:Tyler Thurtell/Quantum Trajectories

Source: Wikipedia, the free encyclopedia.

Quantum trajectories is a method for studying the dynamics of open quantum systems. In particular, it predicts the possibility of particular time evolutions in the continuous measurement regime. If the measurement record is not kept and many trajectories averaged together master equation dynamics are recovered. In this case quantum trajectories leads to an stochastic numerical algorithm that has some memory usage advantages. In the case that the measurement record is kept, individual trajectories correspond to possible observed time evolutions.

The quantum trajectory method was first discussed by Jean Dalibard, Yvan Catsinm and Klaus Mølmer et al.[1][2] The theory was developed around the same time by Howard Carmichael who presented the material in a series of lectures at the Université Libre de Bruxelles that were later published as a book[3]. Early discussions can also be found in[4][5]

Theory

The dynamics of open quantum systems are governed by the Lindblad master equation

where is the density matrix representing the quantum state and is the Hamiltonian of the system. The Lindbladian takes the form

The operators are called jump operators and depend on the particular physical situation being considered. We can rewrite this equation as

Where the effective Hamiltonian is given by

Over a very short time interval this implies that the change in the density operator is (to first order in )

If the system begins in a pure state, and a continuous measurement is performed to determine whether the system is in any of the states , then this equation says that

with probability

When this happens it is referred to as a jump. Alternatively, with probability

the state is

since, to first order in

The normalization factors are necessary because neither of the time evolution options are non-unitary. It may seem mysterious that the evolution in non-unitary even in the case where no jump occurs. This can be understood in terms of the information gained about the system in that time. Put simply, the lack of a jump provides information that leads the observer to assign a lower probability to the system being in a state that can be effected by . The non-Hermitian part of accomplishes this.

A possible combination of non-unitary time evolution operators and jump operators applied to the state is referred to as a trajectory. Averaging over many trajectories recovers the master equation dynamics. If a measurement with back action of the system corresponding to the operators is actually being preformed by the experimenters then the time of evolution of the quantum state does follow some trajectory although exactly which trajectory is a matter of probability.

Alternative Unravelings of the Master Equation

The choice of jump operators is not unique. For example, the master equation is invariant under the substitutions[6]

where . The quantum trajectories produced by using these new jump operators is referred to as a different unraveling of the master equation. If many trajectories are averaged then the various possible unravelings are all equivalent, i.e. they all reproduce the master equation dynamics. However, if the measurement record is going to be kept then the jump operators which correspond to the particular measurement being preformed must be used to obtain accurate predictions about possible trajectories.

The Quantum Jump Method

This perspective leads to a numerical algorithm that can be used to simulate quantum dynamics. This algorithm is sometimes referred to as the quantum jump method, sometimes as Monte Carlo wavefunctions, and sometimes simply as quantum trajectories. A summary of the algorithm is as follows. If at a time the system is in the state then the time step from to occurs in five steps. First, the probably of each type of collapse is calculated as described in the theory section. Second, The interval is divided up as follows

Third, a random number in the interval is drawn. Fourth, if

then the jump operator is applied. If

Then time evolution occurs according to the non-unitary time evolution operator. Finally, the state must be normalized since in general the jump operators are also non-unitary.

In addition to physical insight about the behavior of quantum systems in the continuous measurement regime, this algorithm offers some efficiency advantages in the case that the state may always be represented by a pure state. In that case, only complex numbers must be stored in contrast to the that must be stored in the entire density matrix in going to be kept track of.

Example: The Jaynes-Cummings Model with Photon Counting

Comparison of master equation and quantum trajectory time evolution for the Jaynes-Cummings model. In units where time is measured in units of . The cavity is tuned to atomic resonance . The coupling strength is also set to . The atomic decay rate out of the cavity and the cavity decay rate are both given by . These plots were produced using The Quantum Toolbox in Python (QuTiP)[7][8].

As a simple example lets consider the dynamics of the Jaynes-Cummings model. The Jaynes-Cummings model describes the interaction of a two level atom with a single cavity mode. The Hamiltonian is given by

where is the atomic resonance frequency, is the resonance frequency of the cavity, is the coupling between the cavity and the atom, is the photon annihilation operator, and is the atomic lower operator. The jump operators used in this example, which correspond to photon being observed outside the cavity, are given by

where is the atomic decay rate out of the cavity and is that cavity decay rate. The Figure displays the photon number and atomic state expectation values as a function of time for (a) master equation evolution, (b) trajectory evolution averaged over many trajectories, and (c) a single trajectory. First, notice that (a) and (b) are basically identical. The single trajectory on the other hand appears quite different. The first two time evolutions are damped oscillations. In the case of the single trajectory the oscillations appear virtually undamped and then suddenly the oscillations stop completely. This end of the oscillations corresponds to a photon being detected outside of the cavity. We should also note that in the single trajectory case the oscillations actually are slightly damped due to the non-Hermitian nature of the effective Hamiltonian but it is not noticeable on this time scale.

Trajectories in Experiments

Sketch of the florescence observed by Nagourney et al. which indicated quantum jumps in a trapped Barium ion.

The idea of a quantum jump dates back to Bohr's 1913 proposal[9] to explain atomic spectra and as discussed above the dynamics of open quantum systems can always be understood in terms of an average over many trajectories. In a modern context, Nagourney et al.[10] observed quantum jumps in a trapped ion in 1986. This was done by coupling the state strongly to the state and weakly to the state. They then observed the florescence from the ion. The florescence jump in an apparently discontinuous fashion from being more or less constant to being essentially zero. This corresponded to the ion jumping between a bright state in which it was Rabi flopping between the two strongly coupled states and a dark state in which it was in the weakly coupled state.

References

  1. ^ Dalibard, Jean; Castin, Yvan; Mølmer, Klaus (February 1992). "Wave-function approach to dissipative processes in quantum optics". Physical Review Letters. 68 (5): 580–583. arXiv:0805.4002. Bibcode:1992PhRvL..68..580D. doi:10.1103/PhysRevLett.68.580. PMID 10045937.
  2. ^ Mølmer, Klaus; Castin, Yvan; Dalibard, Jean (1993). "Monte Carlo wave-function method in quantum optics". Journal of the Optical Society of America. 10 (3): 524. Bibcode:1993JOSAB..10..524M. doi:10.1364/JOSAB.10.000524.
  3. ^ Carmichael, Howard (1993). An Open Systems Approach to Quantum Optics. Springer-Verlag. ISBN 978-0-387-56634-4.
  4. ^ Dum, R.; Zoller, P.; Ritsch, H. (1992). "Monte Carlo simulation of the atomic master equation for spontaneous emission". Physical Review A. 45 (7): 4879–4887. Bibcode:1992PhRvA..45.4879D. doi:10.1103/PhysRevA.45.4879. PMID 9907570.
  5. ^ Hegerfeldt, G. C.; Wilser, T. S. (1992). "Ensemble or Individual System, Collapse or no Collapse: A Description of a Single Radiating Atom". In H.D. Doebner; W. Scherer; F. Schroeck, Jr. (eds.). Classical and Quantum Systems (PDF). Proceedings of the Second International Wigner Symposium. World Scientific. pp. 104–105.
  6. ^ Wiseman, Howard M.; Milburn, Gerard J. (2010). Quantum Measurement and Control. Cambridge University Press. ISBN 978-0-521-80442-4.
  7. ^ Johansson, J. R.; Nori, F. (2012). "QuTiP: An open-source Python framework for the dynamics of open quantum systems". Comp. Phys. Comm. 183: 1760-1772. doi:10.1016/j.cpc.2012.02.021.
  8. ^ Johansson, J. R.; Nori, F. (2013). "QuTiP 2: An open-source Python framework for the dynamics of open quantum systems". Comp. Phys. Comm. 184: 1234. doi:10.1016/j.cpc.2012.11.019.
  9. ^ . Bohr, N. On the constitution of atoms and molecules. Part I. Binding of electrons by positive nuclei. Phil. Mag. 26, 1–25 (1913).
  10. ^ Nagourney, Warren; Sandberg, Jon; Dehmelt, Hans (May 1986). "Shelved Optical Electron Amplifier: Observation of Quantum Jumps". Physical Review Letters. 56 (26): 2797–2799. doi:10.1103/PhysRevLett.56.2797.