34
Elementary Multiphoton Processes in Multimode Scenarios

Nils Trautmann and Gernot Alber

Technische Universität Darmstadt, Institut für Angewandte Physik, Hochschulstrasse 4a, 64289 Darmstadt, Germany

The field of quantum optics has experienced remarkable experimental developments during the past decades (13). Progress in controlling single quantum emitters, such as trapped atoms or ions, and the ability to tailor the mode structure of the electromagnetic radiation field using high finesse cavities has enabled new possibilities in studying resonant light‐matter interactions. This led to a variety of remarkable experiments (46) probing the interaction between single quantum emitters and selected modes of the radiation field and demonstrating quantum communication and quantum information processing (711). However, the implementation of quantum networks based on high finesse cavities coupled to suitable waveguides is still challenging due to lossy connections between cavities and waveguides.

A new approach for harnessing the nonlinear interaction between light and single quantum emitters is to enhance matter‐field couplings in the absence of a strongly mode‐selective optical resonator by confining the photons to subwavelength length scales. This can be achieved by suitable one‐dimensional waveguides, such as nanowires (1216), nanofibers (17,18), in coplanar waveguides (circuit quantum electrodynamics (QED)) (19,20), or even in free space (21) by focusing the light using a parabolic mirror. However, these approaches are inherently connected to multimode scenarios in which a large number of field modes participates in the coupling of the quantum emitters to the radiation field. This vast number of degrees of freedom complicates the theoretical investigation especially if highly nonclassical multiphoton states, such as photon number states, are involved in the systems dynamics. Such states have already been realized in experiment (2224) and are of significant interest for applications in quantum information processing and quantum communication. Hence, there is a need for developing suitable theoretical methods to treat such matter‐field interactions involving highly nonclassical multiphoton states in multimode scenarios.

In recent years several methods addressing this issue have been developed. The Bethe‐ansatz (25) and the input‐output formalism (26) have been used to analyze photon transport in waveguides with an embedded qubit and one‐ and two‐photon scattering matrix elements have been evaluated (27,28). With similar techniques scattering matrix elements for even higher photon number states have also been evaluated (2932). Recently, the input‐output formalism has been generalized to treat many spatially distributed atoms coupled to a common waveguide (33). Particular interesting phenomena arise if non‐Markovian processes are investigated. Recently, a multiphoton scattering theory has been developed to treat (34) these kinds of situations and has been used to evaluate the scattering matrix elements for several scenarios of interest. Starting from initially prepared coherent states and analogously to the technique developed by Mollow (35), displacement transformations are applied, and generalized master equations have been derived for describing the dynamics.

In this chapter, we focus on this line of research and discuss a systematic diagrammatic method for evaluating the time evolution of highly nonclassical multiphoton number states interacting with multiple quantum emitters in multimode scenarios. It allows the interpretation of the system's dynamics in terms of sequences of spontaneous photon emission and absorption processes interconnected by photon propagation between quantum emitters or involving reflection by the boundary of a waveguide or a mirror. This photon path representation for multiphoton states not only allows us to evaluate transition amplitudes between initial and final states in the form of a scattering matrix but also enables us to study the full‐time evolution of the quantum state describing the closed system consisting of emitters, the radiation field, and possible boundary surfaces. This photon path representation is not only restricted to the description of one‐dimensional waveguides but can also be used to evaluate the time evolution of several quantum emitters interacting with the radiation field in large or half‐open cavities or even in free space. For the sake of simplicity, however, we restrict our subsequent considerations to two‐level systems. But it is straightforward to generalize this multiphoton path representation also to more general multilevel systems.

A major advantage of this photon path representation for multiphoton states is that only a finite number of diagrams has to be taken into account for determining the time evolution of finitely many photons over a finite time interval. This is achieved by exploiting the retardation effects caused by the multimode radiation field and basic properties of initially prepared photon number states. The accuracy of this diagrammatic method is only limited by the typical quantum optical approximations, namely the dipole approximation and the assumption that the timescale induced by the atomic transition frequencies is by far the shortest one. Thus, this method offers a systematic possibility to study nonlinear and non‐Markovian processes induced by resonant matter‐field interactions involving highly nonclassical multiphoton states and the full multimode description of the radiation field. This is not only interesting from an applied perspective in order to accomplish tasks relevant for quantum information processing, for example, but also from a fundamental point of view.

This chapter is organized as follows. In Section 34.1 we introduce a generic theoretical model and discuss the main approximations. The multiphoton path representation for describing the time evolution of relevant quantum mechanical transition amplitudes is presented in Section 34.2 and is applied to physical scenarios in Section 34.3.

34.1 A Generic Quantum Electrodynamical Model

We investigate the dynamics of images quantum emitters, for example, atoms or ions, situated at the positions images (images ) interacting almost resonantly with the radiation field in a large or half‐open cavity or in free space. For the sake of simplicity, we assume that the quantum emitters can be modeled by identical two‐level atoms or qubits whose center of mass motion is negligible. The dipole matrix element of atom images is denoted by images , and the corresponding transition frequency is images . In the following equation, we assume that the dipole and the rotating‐wave approximation (RWA) are applicable. For justifying the RWA, we assume that the timescale induced by the atomic transition frequency images is by far the shortest one. The interaction between the two‐level atoms and the quantized (transverse) electromagnetic radiation field is described by the Hamiltonian

34.1 equation

with

34.2 equation

and with the dipole transition operator

34.3 equation

of atom images . The coupling to the radiation field is modeled by introducing the electric field operators images of the transverse modes of the radiation field. In the Schrödinger picture they are given by

34.4 equation

with the orthonormal mode functions images . The mode function images solves the Helmholtz equation and fulfills the boundary conditions modeling the presence of a possible cavity or a wave guide.

34.2 The Multiphoton Path Representation

A solution of the time‐dependent Schrödinger equation of the generic quantum electrodynamical model with Hamiltonian 34.1 can be obtained conveniently with the help of a photon path representation.

34.2.1 Analytical Solution of the Schrödinger Equation

Let us consider the time evolution of an initially prepared quantum state with images photonic and images atomic excitations, that is,

34.5 equation

with images denoting the vacuum state of the radiation field and with images denoting the ground state of all two‐level atoms. Each of the sums images represents a single‐photon wave packet and the amplitudes images fulfill the normalization condition images .

The Schrödinger equation with Hamiltonian 34.1 fulfilling this initial condition is equivalent to an integral equation whose solution can be obtained with the help of a fix‐point iteration procedure. In the interaction picture with Hamiltonian images and quantum state images , this Schrödinger equation and its associated integral equation are given by

34.6 equation

In order to develop an iteration procedure for solving this equation, which terminates after a finite number of iterations for any given finite time interval of duration images , it is necessary to take into account directly all processes describing spontaneous photon emission and reabsorption before a photon has had time to leave the atom. These processes take place during a time interval of the order of images and are responsible for spontaneous decay of an excited atom and for a small level shift of its transition frequency images ( 35,36). It turns out that all the other possible photon emission and absorption processes are delayed by retardation effects caused by photon propagation and characterized by the finite speed of light in vacuum images . These retardation effects cause the corresponding iteration procedure to terminate after a finite number of iterations in any finite interval or for a finite number of initially prepared photons.

For this iteration procedure, the solution of the integral equation 34.6 is split into two parts according to

34.7 equation

with images . Inserting Eq. 34.7 into the Schrödinger equation 34.6 yields

34.8 equation

By applying the definition of images , we get

34.9 equation

with images denoting the electric field operators in the interaction picture and images denoting normal ordering. The commutator in the last term of Eq. 34.9 can be associated with the propagation of a photon emitted by atom images to atom images where it is absorbed again. As outlined in Appendix 34.A for photon propagating in vacuum, this commutator can be related to a dyadic Green operator of the d'Alembert equation. As the dispersion relation of the radiation field is linear, this commutator can be evaluated in a straightforward way yielding the result

34.10 equation

for all images and images being the spontaneous decay rate of an atom in free space with the dielectric constant of the vacuum images . The constant images is the time a photon emitted by atom images needs to propagate to atom images . In the special case images , it is the time a photon emitted by images needs to return again to the same atom after being reflected by the boundary of a waveguide or by the surface of a cavity. In free space, such a recurrence is impossible, i.e., images . The delta distribution appearing in Eq. 34.10 originates from the RWA in which physical processes taking place during timescales of the order of images are approximated by instantaneous processes (35). Thus, Eq. 34.10 reflects the fact that spontaneous emission and reabsorption of a photon before it has left the atom again requires a timescale of the order of images and is responsible for the spontaneous decay of an atom in free space. Furthermore, Eq. 34.10 assumes that the small shift of the transition frequency (Lamb shift) has already been incorporated in a properly renormalized atomic transition frequency images .

Using Eq. 34.10 and choosing images for all images , the Schrödinger equation 34.6 simplifies to

34.11 equation

with images . Together with the initial condition 34.5, Eq. 34.11 is equivalent to the integral equation

34.12 equation

which can be solved using a fix‐point iteration starting with images . In the limit images in the physical sense of images , its solution is given by the multiphoton path representation

34.13 equation

with images denoting the time‐ordering operator. In this solution it has been taken into account that the contributions from the last line of Eq. 34.12 vanish in the physically relevant limit images . The sum of normally ordered terms appearing in Eq. 34.13 can be evaluated by introducing the functions

34.14 equation

They describe the retardation effects arising from spontaneous photon emission and reabsorption processes. Thus, only finitely many terms contribute to the sum of Eq. 34.12, if a finite time interval and an initial photon state with a finite number of photons are considered.

34.2.2 Graphical Representation of the Multiphoton Path Representation

For applying the previously derived multiphoton path representation of Eq. 34.13 and for giving a physical interpretation in terms of subsequent photon emission and absorption processes, a diagrammatic method can be developed. Thereby, each term generated by applying Eq. 34.14 in order to bring Eq. 34.13 into a normally ordered form is represented graphically by a diagram. By generating the finite number of all possible diagrams and summing up their contributions allows to determine the time evolution of the quantum state images for any finite time. In the following discussion, we list the basic elements constituting such a diagram, provide a list of rules for generating all possible diagrams, and discuss the connection between these diagrams and the corresponding analytical expressions in the multiphoton path representation of Eq. 34.13.

Let us start with the graphical representation of the initial state images of Eq. 34.5. An initial atomic excitation of an atom images is represented by a graphical element of the form depicted in Figure 34.1a, and an initial photonic excitation corresponding to a term images is represented by an element of the form depicted in Figure 34.1b. Correspondingly, the initial state defined in Eq. 34.5 is represented by the diagram depicted in Figure 34.1c.

Diagram for the excitations contributing to the initial state.

Figure 34.1 Diagrammatic representation of the excitations contributing to the initial state images of Eq. 34.5: representation of (a) an initial atomic excitation, (b) an initial photonic excitation, and (c) the initial state images .

We can also represent the excitations contributing to the state images of Eq. 34.13 in a similar way. Thereby, each atomic excitation of the state images is represented by a graphical element of the form depicted in Figure 34.2a and denotes an outgoing atomic excitation. Each photonic excitation of the state images is represented by an element of the form depicted in Figure 34.2b and denotes an outgoing photonic excitation. Correspondingly, the state images is represented by the diagram of Figure 34.2c.

Image described by caption and surrounding text.

Figure 34.2 Diagrammatic representation of the excitations contributing to the state images of Eq. 34.13: representation of (a) an outgoing atomic excitation, (b) an outgoing photonic excitation, and (c) the excitations of the state images .

Photon emission and absorption processes, involving an atom images at the intermediate time step images (with images ), are represented in Figure 34.3a,b. The propagation of atomic or photonic excitations during these processes are represented by the diagrams depicted in Figure 34.3c,d. These atomic and photonic excitation lines connect emission processes, absorption processes, and initial and outgoing excitations. In a diagram, an atomic excitation line refers to a single atom only, that is, its beginning and its end connect the same atom.

Image described by caption and surrounding text.

Figure 34.3 Diagrammatic representation of basic processes: representation of (a) an emission of a photon by an excited atom, (b) an absorption of a photon by an atom in the ground state, (c) propagation of an atomic excitation (atomic excitation line), and (d) propagation of a photonic excitation (photonic excitation line).

These graphical elements are assembled to a complete diagram according to a set of rules. For a process involving images absorption and emission processes taking place at intermediate time steps images with images , these rules are as follows:

  1. At each emission process, exactly one photon line starts and exactly one atomic excitation line ends.
  2. At each absorption process, exactly one atomic excitation line starts and exactly one photon line ends.
  3. Each atomic excitation line starts either at an initial atomic excitation or at an absorption process and it ends either at an outgoing atomic excitation or at an emission process.
  4. Each photon line starts either at an initial photonic excitation or at an emission process and it ends either at an outgoing photonic excitation or at an absorption process.
  5. Atomic excitation lines corresponding to the same atom cannot coexist.

In particular, the last rule encodes effects originating from the saturation of an atomic transition. Thus, diagrams containing parts, such as the one depicted in Figure 34.4, are forbidden. Ignoring this latter rule would result in a time evolution in which atoms would behave similarly as harmonic oscillators that do not show any saturation effects.

Geometry for a forbidden diagram.

Figure 34.4 A forbidden diagram: The diagram describes a process in which an already excited atom images absorbs a photon. By excluding such diagrams all saturation effects are taken into account that characterize the excitation of two‐level atoms.

The rules connecting each diagram of this graphical representation with a corresponding term of the multiphoton path representation of images of Eq. 34.13) are as follows:

  1. To a photon line connecting an emission process of atom images at time images with an absorption process at time images (images ) by atom images , we associate the term images .
  2. To a photon line connecting an initial photonic excitation images with an absorption process at atom images and time images , we associate a term images .
  3. To a photon line connecting an initial photonic excitation images with an outgoing photonic excitation, we associate a term images .
  4. To a photon line connecting an emission process of atom images at time images with an outgoing photonic excitation, we associate a term images .
  5. To an atomic excitation line not ending at an outgoing atomic excitation and starting and ending at times images and images , we associate a term images .
  6. To an outgoing atomic excitation and starting at time images , we associate a term images .

The expression assigned to a complete diagram is given by the product of all these terms acting on the state images and being integrated over all intermediate time steps images , with images . The quantum state at time images , that is, images , is obtained by summing over all possible equivalence classes of diagrams that can be constructed by these rules. Thereby, each equivalence class of diagrams appears in this sum only once. Two diagrams are considered to be equivalent if the corresponding photon and atomic excitation lines connect emission and absorption processes that involve the same atoms at the same time steps images and the same initial and final excitations.

So far, we have restricted our discussion to identical two‐level systems. However, it is straightforward to generalize this multiphoton path representation also to multilevel atoms by following the steps of Section 34.2.1. This way an expression quite similar to Eq. 34.13 can be derived and can be represented by an analogous diagrammatic procedure.

34.3 Examples

34.3.1 Processes Involving Only a Single Excitation

In order to discuss the basic features of the multiphoton path representation and the corresponding diagrammatic representation, let us consider the simplest quantum electrodynamical processes involving a single excitation only. This way a direct connection can be established between this multiphoton path representation and the photon path representations that have been discussed in the literature previously in connection with single‐photon processes (3739).

Let us consider the spontaneous decay of a single initially excited atom coupled to the radiation field in free space or in an open waveguide. In free space, this process is described by the diagrams depicted in Figure 34.5a,b. According to the rules of the previous section, the diagram depicted in Figure 34.5a is associated with the contribution

34.15 equation

to Eq. 34.13. It describes the decay of the excited atomic state due to the spontaneous emission of a photon. The emitted single‐photon wave packet is described by the contribution to Eq. 34.13 associated with the diagram of Figure 34.5b, that is,

34.16 equation

The diagram of next higher order is depicted in Figure 34.5c and corresponds to the term

34.17 equation

with images describing the return and reabsorption of a photon by atom 1 after having being emitted by the same atom. In general, such a process gives rise to non‐Markovian effects. In free space or in an open waveguide in which a spontaneously emitted photon cannot return again to the same atom, such a recurrence contribution is impossible so that images and the term 34.17 both vanish. The same argument applies to all other diagrams of higher order. Thus, only the diagrams depicted in Figures 34.5a,b contribute to the pure quantum state describing this process, that is,

equation
Image described by caption and surrounding text.

Figure 34.5 (a)‐(c) Diagrammatic representation of the spontaneous photon emission of a single atom (atom 1) in free space or in an open waveguide. (d) Diagram describing a transfer of the excitation from the initially excited atom images to atom images mediated by a single photon.

If many atoms are present, the excitation of one atom can be transferred to another atom by the exchange of a photon that is emitted spontaneously by an excited atom and absorbed again later by an unexcited atom. In general, such an excitation transfer from one atom to another mediated by the exchange of a single‐photon wave packet leads to non‐Markovian effects, especially if the distance between the two atoms is larger than the characteristic length images of the photon wave packet. A diagram describing such a process is depicted in Figure 34.5d. This diagram describing the excitation transfer from atom 1 to atom 2 is associated with the term

34.18 equation

34.3.2 Scattering of Two Photons by a Single Atom

Diagram for Two photons with initial states f (1)(t0) and f (2)(t0) propagating in a one-dimensional waveguide and interact with a two-level atom.

Figure 34.6 A schematic setup: Two photons with initial states images and images propagate in a one‐dimensional waveguide and interact with a two‐level atom.

The photon path representation of Eq. 34.13 also describes saturation effects properly, which come into play as soon as more than a single excitation is present in the atom‐field system. In the following discussion, we investigate the scattering of two photons propagating in free space or in a waveguide by a single two‐level atom at the fixed position images . We assume that the atom is initially prepared in its ground state images and that two initial photonic excitations are present in the system. Thus, the initial state is given by

equation

A corresponding sketch of a possible experimental setup using a one‐dimensional waveguide is depicted in Figure 34.6. The five diagrams contributing to the particular part of the quantum state images , which describes two outgoing photons, are depicted in Figure 34.7. By adding up the associated terms, we obtain the result

equation

Thereby, the diagram depicted in Figure 34.7a corresponds to the term

equation

and describes the unperturbed time evolution of both incoming photons. The diagrams in Figure 34.7b,c describe scattering processes in which one of the two photons is absorbed by the atom at time images , and the atom emits the photon again spontaneously at the later time images . The other photon is propagating in an unperturbed way. These diagrams correspond to the terms

equation

and

equation

The diagrams in Figure 34.7d,e correspond to the terms

equation

and

equation

They describe scattering processes in which the atom absorbs and re‐emits both of the photons one after the other. Thereby, the nonlinear features of these processes induced by saturation effects originate from the rule that the atom can only absorb a second photon after the first absorbed photon has already been re‐emitted again.

Diagram for describing the scattering of two photons by a single two-level atom.

Figure 34.7 Diagrams describing the scattering of two photons by a single two‐level atom.

34.3.3 Dynamics of Two Atoms

The photon path representation of Eq. 34.13 can also describe the dynamics of many atoms interacting with a radiation field or the non‐Markovian retardation effects arising from the presence of a cavity. Such processes share the characteristic feature that a photon emitted by one atom can return again to the same atom at a later time or it may interact later with one of the residual atoms. In the following discussion, we investigate such a situation involving two two‐level atoms as depicted schematically in Figure 34.8a for a waveguide or in Figure 34.8b for free‐space scenario with two half‐open parabolic cavities. Both cases result in the same dynamics.

Diagram for Two atoms coupled to a multimode radiation field.

Figure 34.8 Two atoms coupled to a multimode radiation field: Photonic excitations can propagate from one atom to the other through a one‐dimensional waveguide (a) and photons are guided by two parabolic mirrors in free space (b). Both setups lead to the same atomic dynamics.

The setup depicted in Figure 34.8 a consists of two atoms coupled to a common waveguide, which forms a loop. Consequently, a photon emitted by one of the atoms can travel to the other atom or it can return again to the original atom. We assume that the atoms couple on to the modes of the radiation field that are guided by the one‐dimensional waveguide. The corresponding free‐space setup is depicted in Figure 34.8b. It consists of two parabolic mirrors facing each other and two atoms. Each of these atoms is supposed to be trapped close to the focal points images and images of these parabolic mirrors. For the sake of simplicity we also assume that the dipole matrix elements of these atoms are oriented along the axis of symmetry of the setup. The ideally conducting parabolic mirrors enhance the matter‐field interactions of the two atoms. In this case the exclusive coupling to the radiation field guided by the one‐dimensional waveguide of Figure 34.8a corresponds to the limit that the mirrors cover almost the full solid angle around the atoms.

In the following paragraph, we discuss the time evolution of the initial state images with the radiation field in its vacuum state and the two atoms being in their excited states. The waveguide as well as the free‐space scenario can be described using the relations

34.19 equation

and

34.20 equation

The constant images denotes the typical time a photon needs to propagate from atom 1 to atom 2 (40). With the help of the path representation and the relations of Eqs. 34.19 and 34.20, the time evolution of the matter‐field system can be evaluated. A major difficulty is caused by the nonlinear behavior originating from the saturation effects of the two excited atoms. However, using the previously discussed diagrammatic Method, the probability of finding both atoms in their excited states at a later time can be determined in a straightforward way. The corresponding results are depicted in Figure 34.9a,b. It is worth comparing these results with the ones in which the nonlinear behavior of the atoms is neglected. In such a harmonic approximation, the two atoms can be replaced by harmonic oscillators according to the substitutions

34.21 equation

with images and images denoting the creation and annihilation operators of a harmonic oscillator. In such a harmonic approximation, the evaluation of the time evolution is simplified significantly because the Hamiltonian operator describes a system of coupled harmonic oscillators. Comparing the situations depicted in Figure 34.9a,b, one realizes that the harmonic approximation is appropriate in the case of Figure 34.9a, but it fails completely in the case of Figure 34.9b. This can be understood in a simple way because in the case of Figure 34.9a, we have images so that the probability, for example, that atom 2 is still excited before the photon emitted by atom 1 can reach it is very small. Consequently, saturation effects are negligible. In the case of Figure 34.9b, we have images so that this probability is no longer negligible. As a result, saturation effects are significant.

Graphical illustration of Time dependence of the probability of exciting both atoms.

Figure 34.9 Time dependence of the probability of exciting both atoms: Exact solution obtained from the diagrammatic method (solid line), and harmonic approximation replacing both two‐level atoms by harmonic oscillators (dashed line). The parameters are images with images (a) and images (b).

34.4 Conclusion

We have developed a diagrammatic method suitable for investigating the time evolution of highly nonclassical multiphoton number states interacting with multiple quantum emitters in extreme multimode scenarios. This method can be applied to study numerous cases of interest in quantum information processing, such as the dynamics of quantum emitters coupled to one‐dimensional waveguides or to the radiation field in large or half‐open cavities or even in free space. Thereby, each term of this photon path representation can be represented by a descriptive photon path involving sequences of spontaneous photon emission and absorption processes involving multiple atoms and multiple photons simultaneously. The accuracy of this diagrammatic method is only limited by the main standard quantum optical approximations, namely the dipole approximation and the assumption that the timescale induced by the atomic transition frequencies is by far the shortest one. Furthermore, it offers the unique feature that in order to obtain exact analytical expressions for a finite time interval, only a finite number of diagrams has to be taken into account. By applying this diagrammatic method we are able to study the matter‐field interaction of single quantum emitters with highly nonclassical multiphoton field states in scenarios ranging from free space or half‐open cavities to waveguides. In particular, our method allows us to study nonlinear and non‐Markovian effects induced by matter‐field interactions on the single‐photon level. The investigation of these effects is interesting not only for possible applications in quantum information processing and quantum communication but also from the fundamental point of view. Thus, our method could be used to design suitable protocols for quantum information processing and quantum communication in a variety of architectures ranging from metallic nanowires coupled to quantum dots to possible applications in free space.

Appendix Evaluation of the Field Commutator

In this section, we evaluate the commutator

34.A.1 equation

which is identical to

equation

apart from terms negligible under the assumption that the timescale induced by images is by far the shortest for the system's dynamics. Thus, in this approximation, we conclude

34.A.2 equation

Furthermore, we have the relation

34.A.3 equation

with images denoting the dyadic Green operator of the electromagnetic radiation field. It satisfies the defining equation

34.A.4 equation

with images denoting the transversal delta distribution. This equation has to be solved under the boundary conditions modeling a possible cavity. Combining Eqs. 34.A.2 and 34.A.3, we obtain the relation

equation

Due to the finite speed of light in vacuum images , the dyadic Green operator images exhibits retardation effects. These retardation effects are inherited by the commutator images and lead to the properties described in Eq. 34.10. Eq. 34.10 can be derived using the well‐known expression for the dyadic Green operator images in free space. In fact, Eq. 34.10 contains an additional purely imaginary term which reflects a level shift (Lamb shift) and which can be incorporated into a properly renormalized atomic transition frequency images .

References

  1. 1 Berman, P. (ed.) (1994) Cavity Quantum Electrodynamics, Academic Press, San Diego, CA.
  2. 2 Walther, H., Varcoe, B.T., Englert, B.G., and Becker, T. (2006) Cavity quantum electrodynamics. Prog. Phys., 69, 1325.
  3. 3 Haroche, S. and Raimond, J.M. (2006) Exploring the Quantum: Atoms, Cavities and Photons, Oxford University Press, Oxford.
  4. 4 Goy, P., Raimond, J., Gross, M., and Haroche, S. (1983) Observation of cavity‐enhanced single‐atom spontaneous emission. Phys. Rev. Lett., 50, 1903–1906.
  5. 5 Meschede, D., Walther, H., and Müller, G. (1985) One‐atom maser. Phys. Rev. Lett., 54 (6), 551.
  6. 6 McKeever, J., Boca, A., Boozer, A.D., Buck, J.R., and Kimble, H.J. (2003) Experimental realization of a one‐atom laser in the regime of strong coupling. Nature, 425, 268–271.
  7. 7 Reiserer, A., Ritter, S., and Rempe, G. (2013) Nondestructive detection of an optical photon. Science, 342 (6164), 1349–1351.
  8. 8 Reiserer, A., Kalb, N., Rempe, G., and Ritter, S. (2014) A quantum gate between a flying optical photon and a single trapped atom. Nature, 508 (7495), 237–240.
  9. 9 Kalb, N., Reiserer, A., Ritter, S., and Rempe, G. (2015) Heralded storage of a photonic quantum bit in a single atom. Phys. Rev. Lett., 114, 220 501.
  10. 10 Boozer, A.D., Boca, A., Miller, R., Northup, T.E., and Kimble, H.J. (2007) Reversible state transfer between light and a single trapped atom. Phys. Rev. Lett., 98 (19), 193 601.
  11. 11 Hofheinz, M., Wang, H., Ansmann, M., Bialczak, R.C., Lucero, E., Neeley, M., O'Connell, A., Sank, D., Wenner, J., Martinis, J.M. et al. (2009) Synthesizing arbitrary quantum states in a superconducting resonator. Nature, 459 (7246), 546–549.
  12. 12 Chang, D., Sørensen, A.S., Hemmer, P., and Lukin, M. (2006) Quantum optics with surface plasmons. Phys. Rev. Lett., 97 (5), 053 002.
  13. 13 Akimov, A., Mukherjee, A., Yu, C., Chang, D., Zibrov, A., Hemmer, P., Park, H., and Lukin, M. (2007) Generation of single optical plasmons in metallic nanowires coupled to quantum dots. Nature, 450 (7168), 402–406.
  14. 14 Schuller, J.A., Barnard, E.S., Cai, W., Jun, Y.C., White, J.S., and Brongersma, M.L. (2010) Plasmonics for extreme light concentration and manipulation. Nat. Mater., 9 (3), 193–204.
  15. 15 Babinec, T.M., Hausmann, B.J., Khan, M., Zhang, Y., Maze, J.R., Hemmer, P.R., and Lončar, M. (2010) A diamond nanowire single‐photon source. Nat. Nanotechnol., 5 (3), 195–199.
  16. 16 Claudon, J., Bleuse, J., Malik, N.S., Bazin, M., Jaffrennou, P., Gregersen, N., Sauvan, C., Lalanne, P., and Gérard, J.M. (2010) A highly efficient single‐photon source based on a quantum dot in a photonic nanowire. Nat. Photonics, 4 (3), 174–177.
  17. 17 Vetsch, E., Reitz, D., Sagué, G., Schmidt, R., Dawkins, S., and Rauschenbeutel, A. (2010) Optical interface created by laser‐cooled atoms trapped in the evanescent field surrounding an optical nanofiber. Phys. Rev. Lett., 104 (20), 203 603.
  18. 18 Goban, A., Choi, K.S., Alton, D.J., Ding, D., Lacroûte, C., Pototschnig, M., Thiele, T., Stern, N.P., and Kimble, H.J. (2012) Demonstration of a state‐insensitive, compensated nanofiber trap. Phys. Rev. Lett., 109, 033 603.
  19. 19 You, J. and Nori, F. (2011) Atomic physics and quantum optics using superconducting circuits. Nature, 474 (7353), 589–597.
  20. 20 Devoret, M. and Schoelkopf, R. (2013) Superconducting circuits for quantum information: an outlook. Science, 339 (6124), 1169–1174.
  21. 21 Maiwald, R., Golla, A., Fischer, M., Bader, M., Heugel, S., Chalopin, B., Sondermann, M., and Leuchs, G. (2012) Collecting more than half the fluorescence photons from a single ion. Phys. Rev. A, 86 (4), 043 431.
  22. 22 Varcoe, B.T., Brattke, S., Weidinger, M., and Walther, H. (2000) Preparing pure photon number states of the radiation field. Nature, 403 (6771), 743–746.
  23. 23 Hofheinz, M., Weig, E., Ansmann, M., Bialczak, R.C., Lucero, E., Neeley, M., O'connell, A., Wang, H., Martinis, J.M., and Cleland, A. (2008) Generation of Fock states in a superconducting quantum circuit. Nature, 454 (7202), 310–314.
  24. 24 Waks, E., Diamanti, E., and Yamamoto, Y. (2006) Generation of photon number states. New J. Phys., 8 (1), 4.
  25. 25 Bethe, H. (1931) Zur theorie der metalle. Z. Phys., 71 (3‐4), 205–226.
  26. 26 Gardiner, C.W. and Collett, M.J. (1985) Input and output in damped quantum systems: quantum stochastic differential equations and the master equation. Phys. Rev. A, 31, 3761–3774.
  27. 27 Shen, J.T. and Fan, S. (2007) Strongly correlated two‐photon transport in a one‐dimensional waveguide coupled to a two‐level system. Phys. Rev. Lett., 98, 153 003.
  28. 28 Fan, S., Kocabaş, c.S.E., and Shen, J.T. (2010) Input‐output formalism for few‐photon transport in one‐dimensional nanophotonic waveguides coupled to a qubit. Phys. Rev. A, 82 (6), 063 821.
  29. 29 Shi, T. and Sun, C.P. (2009) Lehmann‐symanzik‐zimmermann reduction approach to multiphoton scattering in coupled‐resonator arrays. Phys. Rev. B, 79, 205 111.
  30. 30 Zheng, H., Gauthier, D.J., and Baranger, H.U. (2010) Waveguide QED: many‐body bound‐state effects in coherent and Fock‐state scattering from a two‐level system. Phys. Rev. A, 82 (6), 063 816.
  31. 31 Roy, D. (2011) Two‐photon scattering by a driven three‐level emitter in a one‐dimensional waveguide and electromagnetically induced transparency. Phys. Rev. Lett., 106, 053 601.
  32. 32 Xu, S. and Fan, S. (2015) Input‐output formalism for few‐photon transport: a systematic treatment beyond two photons. Phys. Rev. A, 91, 043 845.
  33. 33 Caneva, T., Manzoni, M.T., Shi, T., Douglas, J.S., Cirac, J.I., and Chang, D.E. (2015) Quantum Dynamics of Propagating Photons with Strong Interactions: A Generalized Input‐Output Formalism. arXiv preprint arXiv:1501.04427.
  34. 34 Shi, T., Chang, D.E., and Cirac, J.I. (2015) Multiphoton‐scattering theory and generalized master equations. Phys. Rev. A, 92, 053 834.
  35. 35 Mollow, B. (1975) Pure‐state analysis of resonant light scattering: radiative damping, saturation, and multiphoton effects. Phys. Rev. A, 12 (5), 1919.
  36. 36 Weisskopf, V.F. and Wigner, E.P. (1930) Calculation of the natural brightness of spectral lines on the basis of Dirac's theory. Z. Phys., 63, 54–73.
  37. 37 Alber, G., Bernád, J.Z., Stobińska, M., Sánchez‐Soto, L.L., and Leuchs, G. (2013) QED with a parabolic mirror. Phys. Rev. A, 88, 023 825.
  38. 38 Alber, G. (1992) Photon wave packets and spontaneous decay in a cavity. Phys. Rev. A, 46 (R5338), R5338.
  39. 39 Milonni, P.W. and Knight, P.L. (1974) Retardation in the resonant interaction of two identical atoms. Phys. Rev. A, 10, 1096–1108.
  40. 40 Trautmann, N., Bernád, J.Z., Sondermann, M., Alber, G., Sánchez‐Soto, L.L., and Leuchs, G. (2014) Generation of entangled matter qubits in two opposing parabolic mirrors. Phys. Rev. A, 90, 063 814.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset