User:Kchinni/sandbox/Dicke Model

The Dicke Hamiltonian describes   two-level atoms, with energy splitting  , interacting with a single mode   of the electromagnetic field. The Hamiltonian is given by where  are the creation and the annihilation operators for the field satisfying   and   ,  are the collective spin operators of the atoms satisfying the usual angular momentum commutation relations   and  .

Since the original dipole coupling strength is proportional to  , where   is the volume of the cavity, one can factor out   from the original coupling constant by substituting   where   is the density of the atoms in the cavity. This is the reason for the presence of the factor   in the above Hamiltonian. This Dicke Hamiltonian commutes with  , and also commutes with the parity operator  , where   is the number of excitations in the system. It is straight-forward to check that   using the relations   and  . Notice that the   atoms in this model are not interacting with each other. The atoms are also considered to be identical, so each atom in the ensemble has the same energy splitting   and the same coupling constant   with the field. However, they are considered to be distinguishable, which means that each atom can be given a label.

Some of the interesting phenomena shown by this model include the divergence of atom-field entanglement at   and a quantum phase transition at   between the “normal phase” and the “super-radiant phase” in the thermodynamic limit [1]. In addition, it has been shown this Hamiltonian in the rotating-wave approximation shows a thermal phase transition between the normal phase and the super-radiant phase[2]. For  , the system undergoes the thermal phase transition at the temperature,  . Below this temperature, the system is in super-radiant phase where the atomic ensemble emits light with the intensity proportional to  , which is stronger emission compared to the case where all atoms emit independently. Above this temperature  , the system is in normal phase. For  , the system doesn't exhibit a thermal phase transition. Finally, this Hamiltonian also shows a change in the nearest-neighbor level spacing distribution from the Poissonian distribution to the Wigner-Dyson distribution as   goes through  , indicative of the presence of chaos in the system for  [3]

Rotating-wave approximation

edit

In the quantum-optical systems, the frequencies   and   are, in general, very large compared to the coupling strength  . Therefore, the terms that oscillate at the frequency   average to zero on the time scale set by  [1]. For example, the system realized in [4] has   compared to  . Therefore, one could neglect the rapidly oscillating terms(rotating-wave approximation) in the Dicke Hamiltonian to obtain the following Hamiltonian

 

Also this Hamiltonian has an additional symmetry. The number of excitations   are conserved in this case.

Quantum Phase Transition

edit

The Dicke Hamiltonian shows a   order QPT at   in the thermodynamic limit,  . The presence of this QPT can be seen by solving the Dicke Hamiltonian in the thermodynamic limit, and looking at the second derivative of the ground state energy. Since  , one knows that the ground state of the system is always in the   subspace. This can understood based on the following thought experiment. If  , then the ground state of the system is  . Suppose the value of   is now adiabatically increased with | , then the state of the system at the later time   will always be the instantaneous ground state of the Hamiltonian. We also know that  . These two conditions imply that the ground state of the Hamiltonian will always remain in the  subspace. Therefore,   will be assumed for the rest of the discussion associated with the QPT. In this case, the system can be treated as a single large pseudo-spin system consisting of   levels.

The spin operators in the Hamiltonian can be expressed in terms of the bosonic operators using the Holstein-Primakoff mapping: 

Substituting these expressions into the Dicke Hamiltonian, the Hamiltonian becomes

  (1)

This Hamiltonian can be now diagonalized in the thermodynamic limit. Two effective Hamiltonians,   and  , will be derived: one for   and another for  with each one valid in their respective regimes.

Normal Phase

edit

Taylor expanding the square root function in the Hamiltonian mentioned in 1, and then neglecting all the terms with any powers of   in the denominator will produce a Hamiltonian that is bilinear in the bosonic operators. That bilinear Hamiltonian can then be transformed using the Bogoliubov transformation to the following Hamiltonian[3]: where

   

and

 

From the above equation, one can notice that the excitation energy  is real only when  . Hence, this Hamiltonian is valid only when  . Also, the ground state energy in this phase is  .

Super-radiant Phase

edit

Since the field and the atomic ensemble acquire macroscopic occupations in the super-radiant phase, the bosonic modes needs to be displaced in one of the following ways:

  (2)

and where   and   are assumed to be of the  . After substituting these displaced modes(choosing the positive displacement) in equation 1, the Hamiltonian becomes  where   and  . Now, Taylor expanding the   and neglecting all the terms with any powers of   in the denominator will result in a Hamiltonian that contains terms both linear and bilinear in the bosonic modes. Then setting   and   will eliminate the linear terms, and the Hamiltonian can then be diagonalized using the Bogoliubov transformation[3]. The final Hamiltonian is shown below[3].

 

where

 

 

  ,  

and

 

Notice that   is real only if  , and hence the Hamiltonian is valid only for  . The ground state energy is phase is then given by  . If one had chosen the displacements with the negative signs in equation 2, then the spectrum of   would have been identical to the one obtained above. This implies that the energy spectrum of the Hamiltonian above   becomes degenerate. Also, note that this Hamiltonian   doesn't commute with  ,  , but   where  . This means the parity symmetry   becomes spontaneously broken and two new symmetries appeared,  (one associated with the positive displacement of the mode and the other one associated with the negative displacement of the mode). 

 
The second-derivative of the ground state energy is shown here for  , which shows a discontinuity at  .

Properties of the QPT

edit

From the above analysis, the ground state energy is given by  and the second derivative of the ground state energy is given by  

 
The plot of the excitation energy is shown above for  

It can be seen from the above expression and the plot that there is a discontinuity in the second-derivative of the ground state energy at  . The plot of the excitation energies indicate that   is the energy gap between the ground state and the first excited state, and this gap goes to zero at   as expected for a second-order QPT. Specifically, this gap closes as [3]  , and the characteristic length scale diverges as[3]  [3]. These expressions imply that the critical exponents are given by   and  , as the characteristic length diverges as   and the energy gap closes as  .

Chaos

edit

We know that the Dicke Hamiltonian is integrable in the thermodynamic limit,  , as it was explicitly diagonalized above in the quantum phase transition section. The Dicke Hamiltonian is also integrable in the   limit because in that case, the Hamiltonian can be written as follows: and, this Hamiltonian can be diagonalized by simply displacing the bosonic modes. The eigenvalues in this case are given by   where   and   can be  [3]. For the general case, the Dicke Hamiltonian can be numerically diagonalized in the basis  . To determine the chaoticity in this general case, the authors in [3] have looked at the nearest-neighbor level spacing distribution,  , where  . Then, the chaoticity of the system was quantified using the quantity  , which is defined as   where   is the Poissonian distribution,   is the Wigner-Dyson distribution, and  , which is the value at which   and   first intersect[3]. Notice that   if   and   if  , so the system becomes more chaotic as   goes towards 0. If one plots the quantity   as a function of   for the case of the Dicke Hamiltonian, it can be seen from the plot that the value of   will be decreasing as   is increased   to  , and above  , it is almost zero[3]. In summary, the Hamiltonian seems to become chaotic as   passes through  . This behavior of the chaoticity seems to agree well with the chaoticity of the semi-classical Hamiltonian.

Semi-classical Analysis

edit

As there is no direct analog for the spin operators in the classical mechanics, there are many different approaches to perform semiclassical analysis. Here two different approaches are presented.

Method 1: In this method, the differential equations associated with the expectation values of the operators are obtained, and these expectation values are replaced by the classical variables to obtain a set of nonlinear equations of motion for these observables. More specifically, the time evolution of the expectation values of the observables is given by the equation  . Using this equation, one can obtain 

Assuming that the expectation values of the product of the observables is the product of expectation values of the observables, we have  where  ,   and  . The above set of equations also conserve the pseudo angular momentum,  . For  , there are two fixed points given by  and   with the positive population inverse being unstable and the negative population inverse being stable, which could have been expected by looking at the ground state of the Dicke Hamiltonian at  . For  , the fixed points   and   become unstable and two new stable fixed points appear  ,   and  [1]. These stable points could have also been expected as well because the Hamiltonian in the   has a degenerate spectrum with the eigenvectors that are superpositions of   and  .

Method 2: Another method to perform semiclassical analysis is to obtain the semiclassical Hamiltonian by writing the bosonic modes in equation 1 in terms of the corresponding position and the momentum operators, and then setting the commutators between the position and the corresponding momentum operators to zero. This will result in the Hamiltonian given by[3]

 (3)

where   and  . The Hamilton's equations of motion associated with the above Hamiltonian are then given by

 (4)

where  . One of the fixed point is given by  , which is equivalent to the fixed point   and   obtained by the first method. However, this fixed point is stable only for   unlike the condition   in method 1. The other two fixed points   and  

and  [3]. These two fixed point exist and are stable only for

 .

From the Hamiltonian in equation 3, it can be noticed that this semiclassical system can be treated as a particle moving in the following potential well:   .The plots of this potential well are shown for  .

 
Potential energy well for the Dicke Hamiltonian with   and  

As mentioned in [3], these contour plots don't show bifurcation for all the values of  . Using the solutions obtained from the Hamilton's equations in equation 4, one can plot the Poincare' sections by setting   to a certain value. Note that the value of   is fixed by the energy. The Poincare' sections have been plotted in [3]. In these figures, one can notice that as the value of   is increased from 0 to 1, the trajectories in the phase space appear to become chaotic. Note that this increase in chaoticity as   crosses through   is also expected from the results of the random matrix theory for the corresponding quantum Hamiltonian.

Experimental Implementation

edit

The Dicke model could be implemented using multiple methods. To realize the Dicke model in the absence of a cavity, one would need a large sample of atoms, and the distance between the atoms should be smaller than the wavelength of the field for the collective spontaneous emission[5]. Since all the atoms in the Dicke model should experience the same amount of electric field, large number of atoms has to be arranged in a regular pattern within a small distance for the realization of the model. On the other hand, the atoms should not be too close else there will be dipole-dipole interactions between the atoms. However, in the presence of the cavity, high cavity finesse allows one to use small sample of atoms to achieve superradiance[5].

 
Experimental implementation of the Dicke model using the ring cavity and neutral atoms

Ring Cavity and Neutral Atoms

edit

This experimental scheme is discussed in [1]. In this implementation scheme, an ensemble of atoms are confined in the ring cavity. Two laser fields with Rabi frequencies   and   enter through one cavity mirror, and the quantized field enter and exit through one other cavity mirror as shown in the figure. The two level atom is realized through a pair of Raman channels, whose atomic excitation scheme is shown below.

 
Atomic excitation scheme for the realization of two-level atom

In [1] , it is shown that this system evolves under the effective Dicke Hamiltonian: 

where  ,  ,  ,   and  . Here   and   are the laser frequencies, and  ,  and   are the atomic frequencies. Notice that this model also takes cavity loss into account, but the parameters  ,   and   can be chosen appropriately for the Hamiltonian dynamics to dominate[1].

Circuit QED

edit

 Circuit QED has been used to realize the Dicke model in the RWA for  [6]. In this system, superconducting qubits act as atoms and the cavity mode will be the resonant waveguide mode. With CQED, one can precisely place the effective atoms and cavity frequency is also very controllable. For large  , the physical size of the device could be a problem[7]

References

edit
  1. ^ a b c d e f Dimer, F.; Estienne, B.; Parkins, A. S.; Carmichael, H. J. (2007-01-08). "Proposed realization of the Dicke-model quantum phase transition in an optical cavity QED system". Physical Review A. 75 (1): 013804. doi:10.1103/PhysRevA.75.013804. S2CID 5513429.
  2. ^ Hepp, Klaus; Lieb, Elliott H. (1973-04-01). "On the superradiant phase transition for molecules in a quantized radiation field: the dicke maser model". Annals of Physics. 76 (2): 360–404. doi:10.1016/0003-4916(73)90039-0. ISSN 0003-4916.
  3. ^ a b c d e f g h i j k l m n Emary, Clive; Brandes, Tobias (2003-06-12). "Chaos and the quantum phase transition in the Dicke model". Physical Review E. 67 (6): 066203. arXiv:cond-mat/0301273. doi:10.1103/PhysRevE.67.066203. PMID 16241322. S2CID 13865871.
  4. ^ Baumann, Kristian. "Experimental Realization of the Dicke Quantum Phase Transition" (PDF).
  5. ^ a b Gross, M.; Haroche, S. (1982-12-01). "Superradiance: An essay on the theory of collective spontaneous emission". Physics Reports. 93 (5): 301–396. doi:10.1016/0370-1573(82)90102-8. ISSN 0370-1573.
  6. ^ Fink, J. M.; Bianchetti, R.; Baur, M.; Göppl, M.; Steffen, L.; Filipp, S.; Leek, P. J.; Blais, A.; Wallraff, A. (2009-08-17). "Dressed Collective Qubit States and the Tavis-Cummings Model in Circuit QED". Physical Review Letters. 103 (8): 083601. arXiv:0812.2651. doi:10.1103/PhysRevLett.103.083601. PMID 19792728. S2CID 18510324.
  7. ^ Garraway, Barry M. (2011-03-28). "The Dicke model in quantum optics: Dicke model revisited". Philosophical Transactions of the Royal Society of London A: Mathematical, Physical and Engineering Sciences. 369 (1939): 1137–1155. doi:10.1098/rsta.2010.0333. ISSN 1364-503X. PMID 21320910. S2CID 8603278.
edit