Skip to main content
Log in

An excited atom interacting with a Chern insulator: toward a far-field resonant Casimir–Polder repulsion

  • Regular Article – Optical Phenomena and Photonics
  • Published:
The European Physical Journal D Aims and scope Submit manuscript

Abstract

We investigate the resonant Casimir–Polder interaction of an excited atom which has a single (electric dipole) transition with a Chern insulator, using the approach of quantum linear response theory. The Chern insulator has a nonzero, time-reversal symmetry breaking Hall conductance, leading to an additional contribution to the resonant Casimir–Polder interaction which depends on the coupling between the Hall conductance and the circular polarization state of the atomic transition. We find that the resonant Casimir–Polder shift can be significantly enhanced if the atomic de-excitation frequency is near a value associated with a van Hove singularity of the Chern insulator. Furthermore, we find that the resonant Casimir–Polder force can become monotonically decaying and repulsive for a relatively large atom–surface distance. This happens if the atomic dipole transition is right (left) circularly polarized and the Chern number of the Chern insulator is \(-1\) (\(+1\)), and the atomic de-excitation energy is comparable to the bandgap energy of the Chern insulator. This has potential implications for the design of atom–surface interaction which can be tuned repulsive over a relatively large range of separations.

Graphic abstract

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

Data availability statement

This manuscript has no associated data or the data will not be deposited. The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

Notes

  1. The reflection coefficients in Equation (2.16) in Ref. [37] are valid only for the case where \(\omega \) is set to \(\omega _{10}\). Equations (2.9) and the reflection coefficients in Appendix A are valid for general values of \(\omega \).

  2. See e.g. Eqs. (3.34) and (3.36) of S. Fuchs, Control of dispersion interactions of atoms near surfaces. (Dissertation zur Erlangung des Doktorgrades des Fakultät für Mathematik und Physik der Albert-Ludwigs-Universität Freiburg im Breisgau, Dezember 2018).

References

  1. H.B.G. Casimir, D. Polder, The influence of retardation on the London-van der Waals forces. Phys. Rev. 73, 360 (1948)

    ADS  MATH  Google Scholar 

  2. D. Bloch, M. Ducloy, Atom-wall interaction, in Advances in Atomic, Molecular and Optical Physics, vol. 50, ed. by B. Bederson, H. Walther (Elsevier Academic, Amsterdam, 2005), p.91

    Google Scholar 

  3. A. Laliotis, B.-S. Lu, M. Ducloy, D. Wilkowski, Atom-surface physics: A review. AVS Quantum Sci. 3, 043501 (2021)

    ADS  Google Scholar 

  4. M. Bordag, G. Klimchitskaya, U. Mohideen, V. Mostepanenko, Advances in the Casimir Effect, 1st edn. (Oxford University Press, UK, 2009)

    MATH  Google Scholar 

  5. F. Shimizu, Specular reflection of very slow metastable neon atoms from a solid surface. Phys. Rev. Lett. 86, 987 (2001)

    ADS  Google Scholar 

  6. H. Oberst, Y. Tashiro, K. Shimizu, F. Shimizu, Quantum reflection of He on silicon. Phys. Rev. A 71, 052901 (2005)

    ADS  Google Scholar 

  7. D.M. Harber, J.M. Obrecht, J.M. McGuirk, E.A. Cornell, Measurement of the Casimir–Polder force through center-of-mass oscillations of a Bose–Einstein condensate. Phys. Rev. A 72, 033610 (2005)

    ADS  Google Scholar 

  8. P. Wolf et al., From optical lattice clocks to the measurement of forces in the Casimir regime. Phys. Rev. A 75, 063608 (2007)

    ADS  Google Scholar 

  9. Jan Chwedeńczuk, Luca Pezzé, Francesco Piazza, Augusto Smerzi, Rabi interferometry and sensitive measurement of the Casimir–Polder force with ultracold gases. Phys. Rev. A 82, 032104 (2010)

    ADS  Google Scholar 

  10. R. Bennett, D.H.J. O’Dell, Revealing short-range non-Newtonian gravity through Casimir–Polder shielding. New J. Phys. 21, 033032 (2019)

    ADS  Google Scholar 

  11. E.A. Chan, S.A. Aljunid, G. Adamo, A. Laliotis, M. Ducloy, D. Wilkowski, Tailoring optical metamaterials to tune the atom-surface Casimir–Polder interaction. Sci. Adv. 4, eaao4223 (2018)

    ADS  Google Scholar 

  12. S. Ribeiro, S. Scheel, Shielding vacuum fluctuations with graphene. Phys. Rev. A 88, 042519 (2013) [Erratum: Phys. Rev. A 89, 039904 (2014)]

  13. Y.V. Churkin, A.B. Fedortsov, G.L. Klimchitskaya, V.A. Yurova, Comparison of hydrodynamic and Dirac models of dispersion interaction between graphene and H, He\(^*\), or Na atoms. Phys. Rev. B 82, 165433 (2010)

    ADS  Google Scholar 

  14. T. Cysne, W.J.M. Kort-Kamp, D. Oliver, F.A. Pinheiro, F.S.S. Rosa, C. Farina, Tuning the Casimir–Polder interaction via magneto-optical effects in graphene. Phys. Rev. A 90, 052511 (2014)

    ADS  Google Scholar 

  15. N. Khusnutdinov, R. Kashapov, L.M. Woods, The Casimir–Polder effect for a stack of conductive planes. Phys. Rev. A 94, 012513 (2016)

    ADS  Google Scholar 

  16. V.M. Marachevsky, Y.M. Pis’mak, Casimir–Polder effect for a plane with Chern–Simons interaction. Phys. Rev. D 81, 065005 (2010)

    ADS  Google Scholar 

  17. S.Y. Buhmann, V.N. Marachevsky, S. Scheel, Charge-parity-violating effects in Casimir–Polder potentials. Phys. Rev. A 98, 022510 (2018)

    ADS  Google Scholar 

  18. J.A. Crosse, S. Fuchs, S.Y. Buhmann, Electromagnetic Green’s function for layered topological insulators. Phys. Rev. A 92, 063831 (2015)

    ADS  Google Scholar 

  19. S. Fuchs, J.A. Crosse, S.Y. Buhmann, Casimir-Polder shift and decay rate in the presence of nonreciprocal media. Phys. Rev. A 95, 023805 (2017)

    ADS  Google Scholar 

  20. M.G. Silveirinha, S.A.H. Gangaraj, G.W. Hanson, M. Antezza, Fluctuation-induced forces on an atom near a photonic topological material. Phys. Rev. A 97, 022509 (2018)

    ADS  Google Scholar 

  21. C. Eberlein, R. Zietal, Casimir-Polder interaction between a polarizable particle and a plate with a hole. Phys. Rev. A 83, 052514 (2011)

    ADS  Google Scholar 

  22. K.V. Shajesh, M. Schaden, Repulsive long-range forces between anisotropic atoms and dielectrics. Phys. Rev. A 85, 012523 (2012)

    ADS  Google Scholar 

  23. K.A. Milton, P. Parashar, N. Pourtolami, Casimir-Polder repulsion: polarizable atoms, cylinders, spheres and ellipsoids. Phys. Rev. D 85, 025008 (2012)

    ADS  Google Scholar 

  24. M. Antezza, L.P. Pitaevskii, S. Stringari, New asymptotic behavior of the surface-atom force out of thermal equilibrium. Phys. Rev. Lett. 95, 113202 (2005)

    ADS  Google Scholar 

  25. M.-P. Gorza, S. Saltiel, H. Failache, M. Ducloy, Quantum theory of van der Waals interactions between excited atoms and birefringent dielectric surfaces. Eur. Phys. J. D 15, 113 (2001)

    ADS  Google Scholar 

  26. X.-L. Qi, Y.-S. Wu, S.-C. Zhang, Topological quantization of the spin Hall effect in two-dimensional paramagnetic semiconductors. Phys. Rev. B 74, 085308 (2006)

    ADS  Google Scholar 

  27. J.K. Asboth, L. Oroszlány, A. Pályi, A Short Course on Topological Insulators (Springer, Heidelberg, 2016)

    MATH  Google Scholar 

  28. B.-S. Lu, The Casimir effect in topological matter. Universe 7, 237 (2021)

    ADS  Google Scholar 

  29. J.M. Wylie, J.E. Sipe, Quantum electrodynamics near an interface. Phys. Rev. A 30, 1185 (1984)

    ADS  Google Scholar 

  30. J.M. Wylie, J.E. Sipe, Quantum electrodynamics near an interface. II Phys. Rev. A 32, 2030 (1985)

    ADS  Google Scholar 

  31. R.R. Chance, A. Prock, R. Silbey, Frequency shifts of an electric-dipole transition near a partially reflecting surface. Phys. Rev. A 12, 1448 (1975)

    ADS  Google Scholar 

  32. M. Fichet, F. Schuller, D. Bloch, M. Ducloy, van der Waals interactions between excited-state atoms and dispersive dielectric surfaces. Phys. Rev. A 51, 1553 (1995)

    ADS  Google Scholar 

  33. H. Failache, S. Saltiel, M. Fichet, D. Bloch, M. Ducloy, Resonant van der Waals repulsion between excited cs atoms and sapphire surface. Phys. Rev. Lett. 83, 5467 (1999)

    ADS  Google Scholar 

  34. H. Failache, S. Saltiel, M. Fichet, D. Bloch, M. Ducloy, Resonant coupling in the van der Waals interaction between an excited alkali atom and a dielectric surface: an experimental study via stepwise selective reflection spectroscopy. Eur. Phys. J. D 23, 237 (2003)

    ADS  Google Scholar 

  35. M.-P. Gorza, M. Ducloy, van der Waals interactions between atoms and dispersive surfaces at finite temperature. Eur. Phys. J. D 40, 343 (2006)

    ADS  Google Scholar 

  36. H. Nha, W. Jhe, Cavity quantum electrodynamics between parallel dielectric surfaces. Phys. Rev. A 54, 3505 (1996)

    ADS  Google Scholar 

  37. B.-S. Lu, K.Z. Arifa, X.R. Hong, Spontaneous emission of a quantum emitter near a Chern insulator: Interplay of time-reversal symmetry breaking and van Hove singularity. Phys. Rev. B 101, 205410 (2020)

    ADS  Google Scholar 

  38. G. Czycholl, Theoretische Festkörperphysik, vol. 2 (Springer-Verlag GmbH, Deutschland, 2017)

  39. J. Cayssol, Introduction to Dirac materials and topological insulators. Comptes Rendus Physique 14, 760 (2013)

    ADS  Google Scholar 

  40. B.A. Bernevig, T.L. Hughes, Topological Insulators and Topological Superconductors (Princeton University Press, Princeton, 2013)

    MATH  Google Scholar 

  41. H. Weng, R. Yu, X. Hu, X. Dai, Z. Fang, Quantum anomalous Hall effect and related topological electronic states. Adv. Phys. 64, 227 (2015)

    ADS  Google Scholar 

  42. C.-X. Liu, S.-C. Zhang, X.-L. Qi, The quantum anomalous Hall effect: theory and experiment. Annu. Rev. Condens. Matter Phys. 7, 301 (2016)

    ADS  Google Scholar 

  43. Y. Ren, Z. Qiao, Q. Niu, Topological phases in two-dimensional materials: a review. Rep. Prog. Phys. 79, 066501 (2016)

    ADS  Google Scholar 

  44. J. Zhang, B. Zhao, T. Zhou, Z. Yang, Quantum anomalous Hall effect in real materials. Chin. Phys. B 25, 117308 (2016)

  45. W. C. Chew, Waves and Fields in Inhomogeneous Media (Wiley-IEEE Press, New York, USA (1999)), Sec. 2.2

  46. A. Laliotis, M. Ducloy, Casimir–Polder effect with thermally excited surfaces. Phys. Rev. A 91, 052506 (2015)

    ADS  Google Scholar 

  47. V.M. Fain, Ya.I. Khanin, Quantum Electronics Volume 1: Basic Theory (Pergamon Press, Oxford, 1969)

Download references

Acknowledgements

One of the authors (BSL) thanks David Wilkowski for constructive discussions. This work was supported by funding from the Research Development Funding of Xi’an Jiaotong-Liverpool University (XJTLU) under the grant number RDF-21-02-005.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed equally to the paper.

Corresponding author

Correspondence to Bing-Sui Lu.

Appendices

Appendix A: Fresnel coefficients

In this appendix, we present the formulas (with dimensions restored) for the Fresnel coefficients obtained for the case of a single Chern insulator [37]. We denote by the symbols \(r_{ss}\), \(r_{sp}\), \(r_{ps}\) and \(r_{pp}\), respectively, the reflection coefficients for an incident s-polarized wave which is reflected as an s-polarized wave, an incident s-polarized wave which is reflected as an p-polarized wave, an incident p-polarized wave which is reflected as an s-polarized wave, and an incident p-polarized wave which is reflected as an p-polarized wave. For the transmission coefficients, we replace the symbol r by the symbol t.

$$\begin{aligned} r_{ss}= & {} -\frac{1}{\Delta } \bigg ( \frac{4\pi ^2}{c^2} (\sigma _{xx}^2 + \sigma _{xy}^2) + \frac{2\pi }{c} \big ( 1 - (ck_\parallel /\omega )^2 \big )^{-1/2} \sigma _{xx} \bigg ),\nonumber \\ \end{aligned}$$
(A1a)
$$\begin{aligned} r_{ps}= & {} r_{sp} = - \frac{2\pi }{c} \frac{\sigma _{xy}}{\Delta }, \end{aligned}$$
(A1b)
$$\begin{aligned} r_{pp}= & {} \frac{1}{\Delta } \bigg ( \frac{4\pi ^2}{c^2} (\sigma _{xx}^2 + \sigma _{xy}^2) + \frac{2\pi }{c} \big ( 1 - (ck_\parallel /\omega )^2 \big )^{1/2} \sigma _{xx} \bigg ),\nonumber \\ \end{aligned}$$
(A1c)
$$\begin{aligned} t_{ss}= & {} \frac{1}{\Delta } \bigg ( 1 + \frac{2\pi }{c} \big ( 1 - (ck_\parallel /\omega )^2 \big )^{1/2} \sigma _{xx} \bigg ), \end{aligned}$$
(A1d)
$$\begin{aligned} t_{ps}= & {} t_{sp} = - \frac{2\pi }{c} \frac{\sigma _{xy}}{\Delta }, \end{aligned}$$
(A1e)
$$\begin{aligned} t_{pp}= & {} \frac{1}{\Delta } \bigg ( 1 + \frac{2\pi }{c} \big ( 1 - (ck_\parallel /\omega )^2 \big )^{-1/2} \sigma _{xx} \bigg ), \end{aligned}$$
(A1f)

where \(\Delta \equiv 1 + \frac{2\pi }{c} \Big ( \big ( 1 - (ck_\parallel /\omega )^2 \big )^{1/2} + \big ( 1 - (ck_\parallel /\omega )^2 \big )^{-1/2} \Big ) \sigma _{xx} + \frac{4\pi ^2}{c^2} \big ( \sigma _{xx}^2 + \sigma _{xy}^2 \big )\). By defining a dimensionless conductivity \(\widetilde{\sigma }\equiv 2\pi \sigma /c\), we can put the reflection coefficients above in the form Eqs. (2.9).

Appendix B: derivation of energy shifts in the presence of a nonreciprocal medium

In Ref. [30], it was shown using perturbation theory methods that for a quantum emitter interacting with the radiation field via the dipole interaction \(-\varvec{\mu }\cdot \textbf{D}\), the shift \(\delta E_m'\) in the energy \(E_m\) of atomic state \(|m\rangle \) is given by

$$\begin{aligned}&\delta E_m' = - \mathcal {P} \left\{ \sum _{n} \int _{-\infty }^\infty \textrm{d}\omega \, \frac{\mu _a^{mn} \mu _b^{nm} }{\omega + \omega _{nm}}\right. \nonumber \\&\left. \left[ \frac{1}{\hbar } \sum _{B,N} p(B) D_a^{BN}({\textbf{r}}_0) D_b^{NB}({\textbf{r}}_0) \, \delta (\omega - \omega _{NB}) \right] \right\} . \nonumber \\\end{aligned}$$
(B1)

Here, \(\mathcal {P}\) denotes the principal value, \(a, b = 1,2,3\) label Cartesian axes, mn label atomic states, BN label field states, p(B) denotes the prior probability to prepare the field in state \(|B\rangle \), \(\omega _{nm} \equiv (E_n - E_m)/\hbar \), \(\mu _a^{mn} \equiv \langle m | \mu _a | n \rangle \) denotes the dipole transition matrix element from atomic state \(|n\rangle \) to \(|m\rangle \), and \(D_a^{BN} \equiv \langle B | D_a | N \rangle \) denotes the dipole transition matrix element from field state \(|N\rangle \) to \(|B\rangle \). As we are interested in the surface-induced correction to the energy shift, we have ignored a contribution which arises purely from the electromagnetic fluctuations in free space.

In the presence of a nonreciprocal medium, it turns out that the term enclosed by the square brackets in Eq. (B1) is proportional to the anti-Hermitian part of the dyadic Green function. In what follows, we are going to derive this result. We first recall that the dyadic Green tensor \(\mathbb {G}({\textbf{r}},{\textbf{r}}',t)\) connects the dipole \(\textbf{p}(t')\) at time \(t'\) and position \({\textbf{r}}'\) to the expectation value of its displacement field response, \(\langle \textbf{D}({\textbf{r}},t) \rangle \), via

$$\begin{aligned} \langle D_a({\textbf{r}},t) \rangle = \int _{-\infty }^t \textrm{d}t' \, G_{ab}({\textbf{r}},{\textbf{r}}',t-t') \, p_b(t'), \end{aligned}$$
(B2)

with the dyadic Green tensor being given by [47]

$$\begin{aligned}&G_{ab}({\textbf{r}},{\textbf{r}}',t) = \frac{i}{\hbar } \sum _{B,N} p(B) \big [ \langle B | D_a({\textbf{r}},t) | N \rangle \langle N | D_b({\textbf{r}}',0) |B\rangle \nonumber \\&\quad - \langle B | D_b({\textbf{r}}' , 0) | N \rangle \langle N | D_a({\textbf{r}},t) |B\rangle \big ] \Theta (t), \end{aligned}$$
(B3)

where \(\Theta (t)\) is the Heaviside step function, being equal to 1 if \(t > 0\) and 0 if \(t < 0\), and the operator \(D_a({\textbf{r}},t)\) is in the interaction picture. The operator in the interaction picture is related to its counterpart \(D_a({\textbf{r}})\) in the Schrödinger picture by \(D_a({\textbf{r}},t) = e^{iH_0 t/\hbar } D_a({\textbf{r}}) \, e^{-iH_0 t/\hbar }\), where \(H_0\) is the free Hamiltonian for the field in the absence of the dipole interaction. In terms of the Schrödinger picture, the dyadic Green tensor is given by

$$\begin{aligned}{} & {} G_{ab}({\textbf{r}},{\textbf{r}}',t) = \frac{i}{\hbar } \sum _{B,N} p(B)\nonumber \\{} & {} \quad \big [ e^{-i\omega _{NB}t} D_a^{BN}({\textbf{r}}) D_b^{NB}({\textbf{r}}') \nonumber \\ {}{} & {} \quad - e^{i\omega _{NB}t} D_b^{BN}({\textbf{r}}') D_a^{NB}({\textbf{r}}) \big ] \Theta (t).\nonumber \\ \end{aligned}$$
(B4)

By relabeling \(B \leftrightarrow N\) in the second term, we obtain

$$\begin{aligned}{} & {} G_{ab}({\textbf{r}},{\textbf{r}}',t)\nonumber \\{} & {} \quad = \frac{i}{\hbar } \sum _{B,N} e^{-i\omega _{NB}t} \big [ p(B) - p(N) \big ] D_a^{BN}({\textbf{r}}) D_b^{NB}({\textbf{r}}') \Theta (t).\nonumber \\ \end{aligned}$$
(B5)

The Fourier transform is given by

$$\begin{aligned}{} & {} G_{ab}({\textbf{r}},{\textbf{r}}',\omega ) = \int _{-\infty }^{\infty } \textrm{d}t \, e^{i\omega t} G_{ab}({\textbf{r}},{\textbf{r}}',t) \nonumber \\{} & {} \quad =-\frac{1}{\hbar } \sum _{B,N} \big [ p(B) - p(N) \big ] \frac{D_a^{BN}({\textbf{r}}) D_b^{NB}({\textbf{r}}')}{\omega - \omega _{NB} + i0} \nonumber \\{} & {} \quad = -\frac{1}{\hbar } \sum _{B,N}\big [ p(B) - p(N) \big ] D_a^{BN}({\textbf{r}}) D_b^{NB}({\textbf{r}}')\nonumber \\{} & {} \qquad \times \left[ \mathcal {P} \left( \frac{1}{\omega - \omega _{NB}} \right) - i\pi \delta (\omega - \omega _{NB}) \right] . \end{aligned}$$
(B6)

As the displacement field is a physical observable, \(\widehat{\textbf{D}} = \widehat{\textbf{D}}^\dagger \), which implies \((D_{a}^{BN})^*= D_{a}^{NB}\). This implies

$$\begin{aligned} G_{ba}^*({\textbf{r}}',{\textbf{r}},\omega )= & {} -\frac{1}{\hbar } \sum _{B,N}\big [ p(B) - p(N) \big ] D_a^{BN}({\textbf{r}}) D_b^{NB}({\textbf{r}}')\nonumber \\{} & {} \times \left[ \mathcal {P} \left( \frac{1}{\omega - \omega _{NB}} \right) + i\pi \delta (\omega - \omega _{NB}) \right] .\nonumber \\ \end{aligned}$$
(B7)

Noting that \(p(B) - p(N) = p(B) (1 - e^{-\beta \hbar \omega _{NB}})\) and making use of the Dirac delta to change \(\omega _{NB}\) to \(\omega \), we obtain

$$\begin{aligned}{} & {} G_{ab}({\textbf{r}},{\textbf{r}}',\omega ) - G_{ba}^*({\textbf{r}}',{\textbf{r}},\omega ) = \frac{2\pi i}{\hbar } \sum _{B,N} p(B) D_a^{BN}({\textbf{r}}) \nonumber \\{} & {} \quad \times D_b^{NB}({\textbf{r}}') \big ( 1 - e^{-\beta \hbar \omega } \big ) \delta (\omega - \omega _{NB}), \end{aligned}$$
(B8)

which implies

$$\begin{aligned}&\frac{1}{\hbar } \sum _{B,N} p(B) D_a^{BN}({\textbf{r}}) D_b^{NB}({\textbf{r}}') \delta (\omega - \omega _{NB}) \nonumber \\&= \frac{1}{2\pi i} \frac{G_{ab}({\textbf{r}},{\textbf{r}}',\omega ) - G_{ba}^*({\textbf{r}}',{\textbf{r}},\omega )}{1 - e^{-\beta \hbar \omega }}. \end{aligned}$$
(B9)

We can rewrite the denominator of the rightmost term in terms of the Bose-Einstein distribution function, \(n(T,\omega )=(e^{\beta \hbar \omega }-1)^{-1}\) as follows

$$\begin{aligned}&\frac{1}{\hbar } \sum _{B,N} p(B) D_a^{BN}({\textbf{r}}) D_b^{NB}({\textbf{r}}') \delta (\omega - \omega _{NB}) \nonumber \\&=-\frac{n(T,-\omega )}{2\pi i} \bigg (G_{ab}({\textbf{r}},{\textbf{r}}',\omega ) - G_{ba}^*({\textbf{r}}',{\textbf{r}},\omega )\bigg ). \end{aligned}$$
(B10)

When the temperature of the system is at or close to \(T=0K\), the Bose-Einstein distribution function can be approximated in terms of Heaviside function as \(n(T,\omega )\rightarrow -\Theta (-\omega )\), and hence we have that \(n(T,-\omega )\rightarrow -\Theta (\omega )\):

$$\begin{aligned}&\frac{1}{\hbar } \sum _{B,N} p(B) D_a^{BN}({\textbf{r}}) D_b^{NB}({\textbf{r}}') \delta (\omega - \omega _{NB}) \nonumber \\&= \frac{\Theta (\omega )}{2\pi i} \bigg (G_{ab}({\textbf{r}},{\textbf{r}}',\omega ) - G_{ba}^*({\textbf{r}}',{\textbf{r}},\omega )\bigg ). \end{aligned}$$
(B11)

Using this expression, we can rewrite the energy shift \(\delta E'_m\) in Eq. (B1) as

$$\begin{aligned}{} & {} \delta E_m'=-\frac{1}{2\pi i} \mathcal {P} \left( \sum _n \mu _{a}^{mn} \mu _{b}^{nm} \int _{-\infty }^{\infty } \textrm{d}\omega \,\right. \nonumber \\{} & {} \quad \left. \frac{\Theta (\omega )\big (G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; \omega ) - G_{ba}^*({\textbf{r}}_0,{\textbf{r}}_0; \omega )\big )}{\omega + \omega _{nm}} \right) .\nonumber \\ \end{aligned}$$
(B12)

Since the Heaviside function \(\Theta (\omega )\) has values of 1 and 0 if its argument \(\omega \) takes on positive and negative values respectively, only the positive frequency integral will survive. Hence we get

$$\begin{aligned} \delta E_m'= & {} -\frac{1}{2\pi i} \mathcal {P} \left( \sum _n \mu _{a}^{mn} \mu _{b}^{nm} \int _0^\infty \textrm{d}\omega \right. \nonumber \\{} & {} \left. \frac{G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; \omega ) - G_{ba}^*({\textbf{r}}_0,{\textbf{r}}_0; \omega )}{\omega + \omega _{nm}} \right) \nonumber \\= & {} -\frac{1}{2\pi i} \sum _n \mu _{a}^{mn} \mu _{b}^{nm} \int _0^\infty \textrm{d}\omega \nonumber \\{} & {} \frac{G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; \omega ) - G_{ba}^*({\textbf{r}}_0,{\textbf{r}}_0; \omega )}{\omega + \omega _{nm} + i\eta } \nonumber \\{} & {} -\frac{1}{2} \sum _n \mu _{a}^{mn} \mu _{b}^{nm} \int _0^\infty \textrm{d}\omega \big ( G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; \omega )\nonumber \\{} & {} - G_{ba}^*({\textbf{r}}_0,{\textbf{r}}_0; \omega ) \big ) \delta (\omega + \omega _{nm}). \end{aligned}$$
(B13)

In the above, on going to the second step, we have made use of the Plemelj relation, \(\mathcal {P}(1/x) = 1/(x+i\eta ) + i\pi \delta (x)\). Next we make use of the Schwarz reflection property \(G_{ba}^*(\omega ) = G_{ba}(-\omega ^*)\) (which is simply a statement about the reality of the Green function in real time domain) to obtain

$$\begin{aligned} \delta E_m'= & {} -\frac{1}{2\pi i} \sum _n \mu _{a}^{mn} \mu _{b}^{nm} \int _0^\infty \textrm{d}\omega \frac{G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; \omega )}{\omega - \omega _{mn} + i\eta } \nonumber \\{} & {} -\frac{1}{2\pi i} \sum _n \mu _{a}^{mn} \mu _{b}^{nm} \int _{-\infty }^0 \textrm{d}\omega \frac{G_{ba}({\textbf{r}}_0,{\textbf{r}}_0; \omega )}{\omega + \omega _{mn} - i\eta } \nonumber \\{} & {} -\frac{1}{2} \sum _n \mu _{a}^{mn} \mu _{b}^{nm} \int _0^\infty \textrm{d}\omega \, \delta (\omega - \omega _{mn}) G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; \omega )\nonumber \\{} & {} +\frac{1}{2} \sum _n \mu _{a}^{mn} \mu _{b}^{nm} \int _{-\infty }^0 \textrm{d}\omega \, \delta (\omega + \omega _{mn}) G_{ba}({\textbf{r}}_0,{\textbf{r}}_0; \omega )\nonumber \\ \end{aligned}$$
(B14)
Fig. 11
figure 11

Integration contours on the complex frequency plane. There are two poles, which are located at \(\omega = \pm (\omega _{mn} - i\eta )\)

To evaluate the first term, we consider integrating around the contour \(C_+\) in Fig. 11, which does not enclose any pole, and the segment \(R_+\) tends radially to infinity. By the residue theorem we have

$$\begin{aligned}{} & {} \oint _{C_+} \frac{G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; \omega ) \textrm{d}\omega }{\omega - \omega _{mn} + i\eta }\nonumber \\{} & {} \quad = \Big ( \int _0^\infty + \int _{R_+} - \int _0^{i\infty } \Big ) \frac{G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; \omega ) \textrm{d}\omega }{\omega - \omega _{mn} + i\eta } = 0.\nonumber \\ \end{aligned}$$
(B15)

By Jordan’s Lemma, the integral over \(R_+\) vanishes as \(|\omega | \rightarrow \infty \) for \(G_{ab}\) becomes exponentially suppressed. Thus,

$$\begin{aligned} \int _0^\infty \frac{G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; \omega ) \, \textrm{d}\omega }{\omega - \omega _{mn} + i\eta } = i \int _0^\infty \frac{G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; i\xi ) \, \textrm{d}\xi }{- \omega _{mn} + i \xi + i\eta }. \end{aligned}$$
(B16)

Next, to evaluate the second term in Eq. (B14), we consider integrating around contour \(C_-\) in Fig. 11, which encloses a pole at \(\omega = -\omega _{mn} + i\eta \), and the segment \(R_-\) tends radially to infinity. By the residue theorem we have

$$\begin{aligned}{} & {} \oint _{C_-} \frac{G_{ba}({\textbf{r}}_0,{\textbf{r}}_0; \omega ) \, \textrm{d}\omega }{\omega + \omega _{mn} - i\eta } \\{} & {} \quad = 2\pi i \, G_{ba}({\textbf{r}}_0,{\textbf{r}}_0; -\omega _{mn} + i\eta ) \, \Theta (\omega _{mn}) \\{} & {} \quad = \left( \int _{-\infty }^0 + \int _0^{i\infty } + \int _{R_-} \right) \frac{G_{ba}({\textbf{r}}_0,{\textbf{r}}_0; \omega ) \textrm{d}\omega }{\omega + \omega _{mn} - i\eta }. \end{aligned}$$

Again, Jordan’s Lemma implies we can neglect the integral over \(R_-\), and we thus have

$$\begin{aligned}{} & {} \int _{-\infty }^0 \frac{G_{ba}({\textbf{r}}_0,{\textbf{r}}_0; \omega ) \, \textrm{d}\omega }{\omega + \omega _{mn} - i\eta } \nonumber \\{} & {} \quad = - i \int _0^{\infty } \frac{G_{ba}({\textbf{r}}_0,{\textbf{r}}_0; i\xi ) \, \textrm{d}\xi }{\omega _{mn} + i\xi - i\eta }\nonumber \\{} & {} \qquad + 2\pi i \, G_{ba}({\textbf{r}}_0,{\textbf{r}}_0; -\omega _{mn} + i\eta ) \, \Theta (\omega _{mn}) .\nonumber \\ \end{aligned}$$
(B17)

As \(\int _0^\infty \textrm{d}\omega \, \delta (\omega - \omega _{mn}) G_{ab}(\omega ) = G_{ab}(\omega _{mn})\Theta (\omega _{mn})\) and \(\int _{-\infty }^0 \textrm{d}\omega \, \delta (\omega + \omega _{mn}) G_{ba}(\omega ) = G_{ba}(- \omega _{mn}) \Theta (\omega _{mn})\), and using Eqs. (B16) and (B17), we can express Eq. (B14) as

$$\begin{aligned} \delta E_m'= & {} \frac{1}{\pi } \sum _n \mu _a^{mn} \mu _b^{nm} \int _0^\infty \textrm{d}\xi \nonumber \\{} & {} \quad \left( \frac{\omega _{mn} \left( G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; i\xi ) + G_{ba}({\textbf{r}}_0,{\textbf{r}}_0; i\xi ) \right) }{2(\omega _{mn}^2 + \xi ^2)} \right. \nonumber \\{} & {} \left. - \frac{\xi \left( G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; i\xi ) - G_{ba}({\textbf{r}}_0,{\textbf{r}}_0; i\xi ) \right) }{2i(\omega _{mn}^2 + \xi ^2)} \right) \nonumber \\{} & {} - \frac{1}{2} \sum _n \mu _a^{mn} \mu _b^{nm} \Theta (\omega _{mn}) ( G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{mn})\nonumber \\{} & {} + G_{ba}({\textbf{r}}_0,{\textbf{r}}_0; -\omega _{mn})) \nonumber \\= & {} \frac{1}{\pi } \sum _n \mu _a^{mn} \mu _b^{nm} \int _0^\infty \textrm{d}\xi \nonumber \\{} & {} \quad \left( \frac{\omega _{mn} \left( G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; i\xi ) + G_{ba}^*({\textbf{r}}_0,{\textbf{r}}_0; i\xi ) \right) }{2(\omega _{mn}^2 + \xi ^2)} \right. \nonumber \\{} & {} \left. - \frac{\xi \left( G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; i\xi ) - G_{ba}^*({\textbf{r}}_0,{\textbf{r}}_0; i\xi ) \right) }{2i(\omega _{mn}^2 + \xi ^2)} \right) \nonumber \\{} & {} - \frac{1}{2} \sum _n \mu _a^{mn} \mu _b^{nm} \Theta (\omega _{mn})\nonumber \\{} & {} \quad \left( G_{ab}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{mn}) + G_{ba}^*({\textbf{r}}_0,{\textbf{r}}_0; \omega _{mn}) \right) .\nonumber \\ \end{aligned}$$
(B18)

On going to the second equality, we have again made use of the Schwarz reflection property: \(G_{ba}^*(\omega ) = G_{ba}(-\omega ^*)\). For (real) imaginary frequencies, this implies \(G_{ba}(-\omega ) = G_{ba}^*(\omega )\) (\(G_{ba}(i\xi ) = G_{ba}^*(i\xi )\)). Equation (B18) generalizes the formula for the atomic energy level shift to nonreciprocal media.

The first two terms of Eq. (B18) represent the nonresonant contribution to the Casimir–Polder shift, while the third term represents the resonant contribution. For a two-level atom with a single (dipole) transition, the third term leads to \(\delta E_1^{\textrm{cl}}\) and \(\delta E_0^{\textrm{cl}}\) in Eq. (2.2).

We can check that for reciprocal normal insulators (for which \(G_{ba}^*({\textbf{r}}_0,{\textbf{r}}_0; \omega ) = G_{ab}^*({\textbf{r}}_0,{\textbf{r}}_0; \omega )\)), we recover Eq. (2.22) of Ref. [30]. Our result (B18) also agrees with the result obtained from an atomic dynamics-based approach.Footnote 2

Appendix C: Green tensor contractions for various dipole configurations

In this appendix, we collect the results for the contractions of the reflection Green tensor with the dipole orientation vector for three different dipole configurations.

1.1 1 dipole polarization perpendicular to the surface

For the case \(\textbf{n}^{\textrm{T}} = (0,0,1)\), we find

$$\begin{aligned} n_a \big ( G_{ab}^R + G_{ba}^{R*} \big ) n_b^*= & {} 2 {\textrm{Re}}\, \big ( G_{zz}^R \big ), \end{aligned}$$
(C1a)
$$\begin{aligned} n_a^*\big ( G_{ab}^R + G_{ba}^{R*} \big ) n_b= & {} 2 {\textrm{Re}}\, \big ( G_{zz}^R \big ), \end{aligned}$$
(C1b)
$$\begin{aligned} n_a \big ( G_{ab}^R - G_{ba}^{R*} \big ) n_b^*= & {} 2i {\textrm{Im}}\, \big ( G_{zz}^R \big ), \end{aligned}$$
(C1c)
$$\begin{aligned} n_a^*\big ( G_{ab}^R - G_{ba}^{R*} \big ) n_b= & {} 2i {\textrm{Im}}\, \big ( G_{zz}^R \big ). \end{aligned}$$
(C1d)

1.2 2 dipole polarization parallel to the surface

For the case \(\textbf{n}^{\textrm{T}} = (1,0,0)\), we find

$$\begin{aligned} n_a \big ( G_{ab}^R + G_{ba}^{R*} \big ) n_b^*= & {} 2 {\textrm{Re}}\, \big ( G_{xx}^R \big ), \end{aligned}$$
(C2a)
$$\begin{aligned} n_a^*\big ( G_{ab}^R + G_{ba}^{R*} \big ) n_b= & {} 2 {\textrm{Re}}\, \big ( G_{xx}^R \big ), \end{aligned}$$
(C2b)
$$\begin{aligned} n_a \big ( G_{ab}^R - G_{ba}^{R*} \big ) n_b^*= & {} 2i {\textrm{Im}}\, \big ( G_{xx}^R \big ), \end{aligned}$$
(C2c)
$$\begin{aligned} n_a^*\big ( G_{ab}^R - G_{ba}^{R*} \big ) n_b= & {} 2i {\textrm{Im}}\, \big ( G_{xx}^R \big ). \end{aligned}$$
(C2d)

1.3 3 right circular dipole polarization

Finally, for the case \(\textbf{n}^{\textrm{T}} = (1,i,0)/\sqrt{2}\), we find

$$\begin{aligned} n_a \big ( G_{ab}^R + G_{ba}^{R*} \big ) n_b^*= & {} 2 \big ( {\textrm{Re}}\, G_{xx}^R + {\textrm{Im}}\, G_{xy}^R \big ), \end{aligned}$$
(C3a)
$$\begin{aligned} n_a^*\big ( G_{ab}^R + G_{ba}^{R*} \big ) n_b= & {} 2 \big ( {\textrm{Re}}\, G_{xx}^R - {\textrm{Im}}\, G_{xy}^R \big ), \end{aligned}$$
(C3b)
$$\begin{aligned} n_a \big ( G_{ab}^R - G_{ba}^{R*} \big ) n_b^*= & {} 2i \big ( {\textrm{Im}}\, G_{xx}^R - {\textrm{Re}}\, G_{xy}^R \big ), \end{aligned}$$
(C3c)
$$\begin{aligned} n_a^*\big ( G_{ab}^R - G_{ba}^{R*} \big ) n_b= & {} 2i \big ( {\textrm{Im}}\, G_{xx}^R + {\textrm{Re}}\, G_{xy}^R \big ). \end{aligned}$$
(C3d)

Appendix D: Enhancement of the resonant Casimir–Polder shift at \(\widetilde{\omega }_{10}=1.9\)

Fig. 12
figure 12

a Frequency dependences of the imaginary part of \(\widetilde{\sigma }_{xx}\) (red circles) and the real part of \(\widetilde{\sigma }_{xy}\) (blue squares); b the \(\widetilde{k}_\parallel \)-dependence of \(|r_{ss}|\) (blue, dashed), \(|r_{pp}|\) (red, dot-dashed) and \(|r_{ps}|=|r_{sp}|\) (green, dotted) for the dispersive case with \(\widetilde{\omega }_{10} = 1.9\). The behaviors of the corresponding reflection coefficients in the nondispersive limit are plotted in black: \(|r_{ss}|\) (dashed), \(|r_{pp}|\) (dot-dashed), and \(|r_{ps}|= |r_{sp}|\) (dotted)

In this appendix, we consider a circularly polarized dipole near a \(C = -1\) Chern insulator, and look more deeply into how the resonant Casimir–Polder shift \(\delta \widetilde{\omega }_{10}^{\textrm{res}}\) becomes greatly enhanced at \(\widetilde{\omega }_{10} = 1.9\), compared to the nondispersive limit. From Eq. (2.6), the magnitude of \(\delta \widetilde{\omega }_{10}^{\textrm{res}}\) is determined by the magnitude of the reflection Green tensor, and from Eq. (2.7), the latter’s magnitude is in turn determined by the magnitude of the reflection coefficients.

The magnitudes of the reflection coefficients depend on the frequency in the conductances \(\widetilde{\sigma }_{xx}\) and \(\widetilde{\sigma }_{xy}\). The frequency dependence of the conductances is studied and plotted in Sec. II C of Ref. [37], where we found that for \(\widetilde{\omega }_{10} < 2\), the real part of \(\widetilde{\sigma }_{xx}\) and the imaginary part of \(\widetilde{\sigma }_{xy}\) both vanish. For the reader’s convenience, we quote here the formulaic results we found using the Kubo formula for the conductivity tensor [37]:

$$\begin{aligned} \sigma _{xx}(\omega )= & {} -\frac{1}{\hbar } \int _{BZ} \frac{d^2\textbf{k}}{(2\pi )^2} {\textrm{Re}} \, [\langle + | j_x | - \rangle \langle - | j_x | + \rangle ] \frac{f(d) - f(-d)}{2d} \nonumber \\{} & {} \times \bigg ( \pi \big ( \delta (\hbar \omega + 2d) + \delta (\hbar \omega - 2d) \big ) + \frac{2 i \hbar \omega }{\hbar ^2\omega ^2 - 4d^2} \bigg ),\nonumber \\ \end{aligned}$$
(D1a)
$$\begin{aligned} \sigma _{xy}(\omega )= & {} \frac{1}{\hbar } \int _{BZ} \frac{d^2\textbf{k}}{(2\pi )^2} \, {\textrm{Im}} \, [\langle + | j_x | - \rangle \langle - | j_y | + \rangle ] \frac{f(d) - f(-d)}{2d} \nonumber \\{} & {} \times \bigg ( \frac{4d}{4d^2-\hbar ^2\omega ^2} + i \pi \big ( \delta (\hbar \omega - 2d) - \delta (\hbar \omega + 2d) \big ) \bigg ),\nonumber \\ \end{aligned}$$
(D1b)

where \(d = \sqrt{d_x^2 + d_y^2 + d_z^2}\), \(d_x = t \sin k_x a\), \(d_y = t \sin k_y a\), \(d_z = t (\cos k_x a + \cos k_y a) + u\), \(f(d) = (\exp (\beta d) + 1)^{-1}\), and

$$\begin{aligned}{} & {} {\textrm{Im}}\, [\langle + | j_x | - \rangle \langle - | j_y | + \rangle ] \nonumber \\{} & {} \quad = - {\textrm{Im}}\, [\langle - | j_x | + \rangle \langle + | j_y | - \rangle ] \nonumber \\{} & {} \quad =\frac{t^3 (ae)^2}{4\,d (d_x^2+d_y^2)} \left( \left( \cos k_x a \sin ^2 k_y a + \cos k_y a \sin ^2 k_x a \right) \right. \nonumber \\{} & {} \quad \qquad \left( 2(d_x^2+d_y^2 + t^2) - t^2 (\cos 2k_x a + \cos 2k_y a) \right) \nonumber \\{} & {} \qquad \left. + 4td_z \cos k_x a \cos k_y a \left( \sin ^2 k_x a + \sin ^2 k_y a \right) \right) ,\nonumber \\ \end{aligned}$$
(D2a)
$$\begin{aligned}{} & {} \qquad \quad {\textrm{Re}}\, [\langle + | j_x | - \rangle \langle - | j_x | + \rangle ] \nonumber \\{} & {} \quad =\frac{(tae)^2}{16\,d^2 (d_x^2+d_y^2)} \left( \sin ^2 k_x a \left( t^2 (\cos 2k_x a + \cos 2k_y a) \right. \right. \nonumber \\{} & {} \qquad \left. - 2((d_x^2+d_y^2) + t^2) \right) \nonumber \\{} & {} \qquad \times \left( t^2 (\cos 2k_x a + \cos 2k_y a) - 2(d_x^2+d_y^2 + t^2) \right. \nonumber \\{} & {} \qquad \left. - 8 t d_z \cos k_x a) + 4 t^2 d_z^2 \sin ^2 2k_x a \right. \nonumber \\{} & {} \qquad \left. + 16 t^2 d^2 \cos ^2 k_x a \sin ^2 k_y a \right) . \end{aligned}$$
(D2b)

As in Ref. [37], we show in Fig. 12(a) the plots of the frequency dependence of the conductivity components which are non-vanishing, i.e., the imaginary part of \(\widetilde{\sigma }_{xx}\) (red circles) and the real part of \(\widetilde{\sigma }_{xy}\) (blue squares). As we explained in Ref. [37], for \(\widetilde{\omega }_{10} < 2\), \({\textrm{Im}} \, \widetilde{\sigma }_{xx}\) is zero and \({\textrm{Re}} \, \widetilde{\sigma }_{xy} = -\alpha \approx -1/137\), a feature that we also see in the plot.

Figure 12(b) shows plots for the \(\widetilde{k}_\parallel \)-dependence of the magnitudes of the reflection coefficients for the dispersive case where \(\widetilde{\omega }=1.9\), and also plots for the magnitudes of the reflection coefficients corresponding to the nondispersive limit. In the nondispersive limit, the reflection coefficients have constant values, with \(r_{ss} = - r_{pp} \approx -\alpha ^2 \sim \mathcal {O}(10^{-4})\) and \(r_{ps} = r_{sp} \approx \alpha \sim \mathcal {O}(10^{-2})\). On the other hand, we see that the reflection coefficients for the dispersive case with \(\widetilde{\omega }_{10} = 1.9\) depend on \(\widetilde{k}_\parallel \), and are generally much larger than those in the nondispersive limit.

Finally, let us express \(\widetilde{G}_{xx}^R\) and \(\widetilde{G}_{xy}^R\) in the form

$$\begin{aligned} \widetilde{G}_{xx}^R = \int _0^\infty \textrm{d}\widetilde{k}_\parallel \, g_1(\widetilde{k}_\parallel ), \quad \widetilde{G}_{xy}^R = \int _0^\infty \textrm{d}\widetilde{k}_\parallel \, g_2(\widetilde{k}_\parallel ), \end{aligned}$$
(D3)

which allows us to express Eq. (3.1) in the form

$$\begin{aligned} \delta \widetilde{\omega }_{10}^{\textrm{res}} = - \frac{3}{4} \int _0^\infty \textrm{d}\widetilde{k}_\parallel \Big ( {\textrm{Re}}\, g_1(\widetilde{k}_\parallel ) + {\textrm{Im}}\, g_2(\widetilde{k}_\parallel ) \Big ). \end{aligned}$$
(D4)

In Fig. 13, we plot the behavior of \({\textrm{Re}}\, g_1(\widetilde{k}_\parallel ) + {\textrm{Im}}\, g_2(\widetilde{k}_\parallel )\) for positive values of \(\widetilde{k}_\parallel \), for both the dispersive case with \(\widetilde{\omega }_{10} = 1.9\) (blue) and the nondispersive limit (black). We see that over the range plotted, \({\textrm{Re}}\, g_1(\widetilde{k}_\parallel ) + {\textrm{Im}}\, g_2(\widetilde{k}_\parallel )\) for the dispersive case generally has a much larger magnitude than the nondispersive limit. For values of \(\widetilde{k}_\parallel \) much larger than 1, \({\textrm{Re}}\, g_1(\widetilde{k}_\parallel ) + {\textrm{Im}}\, g_2(\widetilde{k}_\parallel )\) becomes rapidly exponentially suppressed.

To summarize: in the nondispersive limit, \(\widetilde{\sigma }_{xx} = 0\) and \(\widetilde{\sigma }_{xy} \approx -1/137\), whereas the magnitudes of both \(\widetilde{\sigma }_{xx}\) and \(\widetilde{\sigma }_{xy}\) are of the order of unity in the dispersive case with \(\widetilde{\omega }_{10} = 1.9\). This dispersion contributes to an enhancement in the reflection coefficients, which directly results in the enhancement of \(\delta \widetilde{\omega }_{10}^{\textrm{res}}\).

Fig. 13
figure 13

Behavior of \({\textrm{Re}}\, g_1 + {\textrm{Im}}\, g_2\) as a function of \(\widetilde{k}_\parallel \) for the dispersive case with \(\widetilde{\omega }_{10} = 1.9\) (blue) and the nondispersive limit (black). The functions \(g_1\) and \(g_2\) are defined in Eq. (D3)

Appendix E: Asymptotic regimes for the resonant Casimir–Polder shift

In this appendix, we obtain the asymptotic behavior of the resonant Casimir–Polder shift \(\delta \widetilde{\omega }_{10}^{\textrm{res}}\) in the near-field (\(\eta \rightarrow 0\)) and far-field limit \(\eta \gg 1\). For the dipole transitions which are polarized perpendicular to the surface, parallel to the surface, and right circularly polarized in the plane of the surface, this amounts to looking at the asymptotic behavior of \({\textrm{Re}}\,G_{xx}^R\), \({\textrm{Re}}\,G_{zz}^R\) and \({\textrm{Im}}\,G_{xy}^R\). For our asymptotic analysis, we follow the same procedure as the one described in Ref. [37]. Firstly, we scale out the dependence on \(\eta \) in the exponential factors in Eqs. (2.8) by defining a new dimensionless variable \(t \equiv {\widetilde{k}}_z \eta \) for the range \(0 \le {\widetilde{k}}_\parallel < 1\) (recalling that \({\widetilde{k}}_z = (1 - {\widetilde{k}}_\parallel ^2)^{1/2}\)), whence \({\widetilde{k}}_\parallel \textrm{d}{\widetilde{k}}_\parallel = - {\widetilde{k}}_z \textrm{d}{\widetilde{k}}_z = - t \, \textrm{d}t/\eta ^2\). For this range, \(0 \le {\widetilde{k}}_z < 1\) and thus \(0 \le t < \eta \). Then for the range \(1 \le {\widetilde{k}}_\parallel < \infty \), we have \(0 \le {\widetilde{k}}_z < i\infty \). If we define \({\widetilde{k}}_z \equiv i \ell \), then \(0 \le \ell < \infty \), and we can define \(t \equiv \ell \eta \), whereupon \({\widetilde{k}}_\parallel \textrm{d}{\widetilde{k}}_\parallel = \ell \, \textrm{d}\ell = t \, \textrm{d}t/\eta ^2\). After rescaling and using Eqs. (2.9), Eqs. (2.8) become

$$\begin{aligned}{} & {} {\widetilde{G}}_{xx}^R({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) = g_{1}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10}) + h_{1}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10}), \nonumber \\{} & {} g_{1}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10}) = - \frac{i}{2\eta } \int _0^\eta \textrm{d}t\nonumber \\{} & {} \quad \frac{ [(1 + (\frac{t}{\eta })^2)(\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2) + (\frac{\eta }{t} + (\frac{t}{\eta })^3)\widetilde{\sigma }_{xx}]e^{it} }{1 + \widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + (\frac{\eta }{t} + \frac{t}{\eta })\widetilde{\sigma }_{xx}}, \nonumber \\{} & {} h_{1}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10}) =- \frac{1}{2\eta } \int _0^\infty \textrm{d}t \nonumber \\{} & {} \quad \frac{[(1 - (\frac{t}{\eta })^2)(\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2) - i ((\frac{t}{\eta })^3 + \frac{\eta }{t})\widetilde{\sigma }_{xx}]e^{-t}}{1 + \widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + i (\frac{t}{\eta } - \frac{\eta }{t})\widetilde{\sigma }_{xx}}; \end{aligned}$$
(E1a)
$$\begin{aligned}{} & {} {\widetilde{G}}_{xy}^R({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) = {\widetilde{G}}_{yx}^R({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \nonumber \\{} & {} \quad = g_{2}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10}) + h_{2}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10}), \nonumber \\{} & {} g_{2}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10}) = - \frac{i}{\eta ^2} \int _0^\eta \textrm{d}t \, t\nonumber \\{} & {} \quad \frac{\widetilde{\sigma }_{xy}}{1 + \widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + (\frac{\eta }{t} + \frac{t}{\eta })\widetilde{\sigma }_{xx}} e^{it}, \nonumber \\{} & {} h_{2}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10}) = - \frac{i}{\eta ^2} \int _0^\infty \textrm{d}t \, t \nonumber \\{} & {} \quad \frac{\widetilde{\sigma }_{xy}}{1 + \widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + i (\frac{t}{\eta } - \frac{\eta }{t})\widetilde{\sigma }_{xx}} e^{-t}; \end{aligned}$$
(E1b)
$$\begin{aligned}{} & {} {\widetilde{G}}_{zz}^R({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) = g_{3}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10}) + h_{3}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10}), \nonumber \\{} & {} g_{3}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10}) = \frac{i}{\eta } \int _0^\eta \textrm{d}t \, \bigg ( 1 - \frac{t^2}{\eta ^2} \bigg ) \nonumber \\{} & {} \quad \frac{\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + \frac{t}{\eta } \widetilde{\sigma }_{xx}}{1 + \widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + (\frac{\eta }{t} + \frac{t}{\eta })\widetilde{\sigma }_{xx}} e^{it}, \nonumber \\{} & {} h_{3}({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10}) = \frac{1}{\eta } \int _0^\infty \textrm{d}t \, \bigg ( 1 + \frac{t^2}{\eta ^2} \bigg )\nonumber \\{} & {} \quad \frac{\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + i \frac{t}{\eta } \widetilde{\sigma }_{xx}}{1 + \widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + i (\frac{t}{\eta } - \frac{\eta }{t})\widetilde{\sigma }_{xx}} e^{-t}. \end{aligned}$$
(E1c)

In the above, we have expressed each Green tensor component as the sum of a contribution with an oscillatory integrand, \(g({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10})\), and a contribution with an exponentially decaying integrand, \(h({\textbf{r}}_0,{\textbf{r}}_0; \omega _{10})\).

1.1 1 Near-field asymptotics

Let us consider the asymptotic behavior of the resonant Casimir–Polder shift in the near-field limit in the low and high frequency regimes. This amounts to considering the behavior of the contributions \({\textrm{Re}}\,{\widetilde{G}}_{xx}^R = {\textrm{Re}}\, g_1 + {\textrm{Re}}\, h_1\), \({\textrm{Im}}\,{\widetilde{G}}_{xy}^R = {\textrm{Im}}\, g_2 + {\textrm{Im}}\, h_2\) and \({\textrm{Re}}\,{\widetilde{G}}_{zz}^R = {\textrm{Re}}\, g_3 + {\textrm{Re}}\, h_3\) to the terms in Eq. (E1) in the limit that \(\eta \rightarrow 0\).

As the resonant Casimir–Polder shift for our considered dipole polarizations involves \({\widetilde{G}}_{xx}^R\), \({\widetilde{G}}_{xy}^R\) and \({\widetilde{G}}_{zz}^R\), let us derive the near-field limits for the functions \(g_1\), \(g_2\), \(g_3\), \(h_1\), \(h_2\) and \(h_3\), for the case that the de-excitation frequency is not close to frequencies associated with van Hove singularities. Since \(\eta \) is the upper limit in the integration over t in the functions \(g_1\), \(g_2\) and \(g_3\) in Eqs. (E1), we have that \(t/\eta \rightarrow 1\), \(\eta /t \rightarrow 1\) and \(e^{it} \rightarrow 1\) as \(\eta \rightarrow 0\). We obtain

$$\begin{aligned} g_1({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10})\approx & {} -\frac{i}{2\eta } \int _0^\eta \textrm{d}t\, \frac{2(\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2) + 2\widetilde{\sigma }_{xx}}{1 + \widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + 2\widetilde{\sigma }_{xx}} \nonumber \\{} & {} \quad =- \frac{i(\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + \widetilde{\sigma }_{xx})}{1 + \widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + 2\widetilde{\sigma }_{xx}}, \end{aligned}$$
(E2a)
$$\begin{aligned} g_2({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10})\approx & {} - \int _0^\eta \textrm{d}t \frac{i}{\eta ^2} \frac{t \widetilde{\sigma }_{xy}}{1 + \widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + 2\widetilde{\sigma }_{xx}} \nonumber \\{} & {} \quad =-\frac{i}{2} \frac{\widetilde{\sigma }_{xy}}{1 + \widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + 2\widetilde{\sigma }_{xx}}, \end{aligned}$$
(E2b)
$$\begin{aligned} g_3({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10})\approx & {} \frac{i}{\eta } \int _0^\eta \textrm{d}t \, \frac{\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + \widetilde{\sigma }_{xx}}{1 + \widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + 2\widetilde{\sigma }_{xx}} \Big (1 - \frac{t^2}{\eta ^2}\Big ) \nonumber \\{} & {} \quad =\frac{2i}{3} \frac{\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + \widetilde{\sigma }_{xx}}{1 + \widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2 + 2\widetilde{\sigma }_{xx}}. \end{aligned}$$
(E2c)

Next, let us consider the near-field limits for the functions \(h_1\), \(h_2\) and \(h_3\). We can perform a perturbation expansion in powers of \(\eta /t\) in the integrand and retain the leading order term. This leads to

$$\begin{aligned} h_1({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10})\approx & {} \frac{1}{2\eta ^3} \int _0^\infty \textrm{d}t \, t^2 e^{-t} = \frac{1}{\eta ^3}, \end{aligned}$$
(E3a)
$$\begin{aligned} h_2({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10})\approx & {} - \frac{1}{\eta } \int _0^\infty \textrm{d}t \, \frac{\widetilde{\sigma }_{xy}}{\widetilde{\sigma }_{xx}} e^{-t} = - \left( \frac{\widetilde{\sigma }_{xy}}{\widetilde{\sigma }_{xx}} \right) \frac{1}{\eta },\nonumber \\ \end{aligned}$$
(E3b)
$$\begin{aligned} h_3({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10})\approx & {} \frac{1}{\eta ^3} \int _0^\infty \textrm{d}t\, t^2 e^{-t} = \frac{2}{\eta ^3}, \end{aligned}$$
(E3c)

From the above, we see again that the near-field limiting values of \({\textrm{Re}}\,h_1\), \({\textrm{Im}}\,h_2\) and \({\textrm{Re}}\,h_3\) dominate the limiting values of \({\textrm{Re}}\,g_1\), \({\textrm{Im}}\,g_2\) and \({\textrm{Re}}\,g_3\), which implies that \({\textrm{Re}}\,{\widetilde{G}}_{xx}^R \approx {\textrm{Re}}\, h_1\), \({\textrm{Im}}\,{\widetilde{G}}_{xy}^R \approx {\textrm{Im}}\, h_2\) and \({\textrm{Re}}\,{\widetilde{G}}_{zz}^R \approx {\textrm{Re}}\, h_3\).

Using Eq. (3.7), we obtain the near-field limit of the resonant Casimir–Polder shift for a dipole transition polarized perpendicular to the surface:

$$\begin{aligned} \delta \widetilde{\omega }_{10}^{\textrm{res}} \approx -\frac{3}{4} {\textrm{Re}} \, h_3({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \approx -\frac{3}{2\eta ^3} \,, \end{aligned}$$
(E4)

Similarly, using Eq. (3.13) we obtain the near-field limit of the resonant Casimir–Polder shift for a dipole transition polarized parallel to the surface:

$$\begin{aligned} \delta \widetilde{\omega }_{10}^{\textrm{res}} \approx -\frac{3}{4} {\textrm{Re}} \, h_1({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \approx -\frac{3}{4\eta ^3} \,. \end{aligned}$$
(E5)

Finally, the near-field limit of the resonant Casimir–Polder shift for a right circular dipole polarization, Eq. (3.1), is

$$\begin{aligned} \delta \widetilde{\omega }_{10}^{\textrm{res}}\approx & {} -\frac{3}{4} \left( {\textrm{Re}} \, h_1({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) + {\textrm{Im}} \, h_2({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \right) \nonumber \\\approx & {} -\frac{3}{4\eta ^3}. \end{aligned}$$
(E6)

The above near-field limits are the same for all the frequency regimes.

1.2 2 Far-field asymptotics

In the far-field limit, \(t/\eta \ll 1\). We can perform a perturbation expansion in powers of \(t/\eta \) in the integrands of the functions \(g_1\), \(g_2\), \(g_3\), \(h_1\), \(h_2\) and \(h_3\) from Eqs. (E1), and retain the leading order term. We obtain

$$\begin{aligned}{} & {} g_1({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \approx -\frac{i}{2\eta } \int _0^\eta \textrm{d}t \, e^{it} = \frac{1}{2\eta } \left( 1 - \cos \eta - i \sin \eta \right) ,\nonumber \\ \end{aligned}$$
(E7a)
$$\begin{aligned}{} & {} g_2({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \approx -\frac{i}{\eta ^3} \int _0^\eta \textrm{d}t \, t^2 e^{it} \frac{\widetilde{\sigma }_{xy}}{\widetilde{\sigma }_{xx}}\nonumber \\{} & {} \quad = - \frac{1}{\eta ^3} \frac{\widetilde{\sigma }_{xy}}{\widetilde{\sigma }_{xx}} \left( 2 + \left( \eta ^2 -2 + 2 i \eta \right) e^{i \eta } \right) , \end{aligned}$$
(E7b)
$$\begin{aligned}{} & {} g_3({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \approx \frac{i}{\eta ^2} \int _0^\eta \textrm{d}t \, t \, e^{it} \left( \frac{\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2}{\widetilde{\sigma }_{xx}} \right) \nonumber \\{} & {} \quad =- \frac{i}{\eta ^2} \left( \frac{\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2}{\widetilde{\sigma }_{xx}} \right) \left( 1 - \left( 1 - i \eta \right) e^{i \eta } \right) , \end{aligned}$$
(E7c)
$$\begin{aligned}{} & {} h_1({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \approx -\frac{1}{2\eta } \int _0^\infty \textrm{d}t \, e^{-t} = - \frac{1}{2\eta }, \end{aligned}$$
(E7d)
$$\begin{aligned}{} & {} h_2({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \approx \frac{1}{\eta ^3} \int _0^\infty \textrm{d}t \, t^2 e^{-t} \left( \frac{\widetilde{\sigma }_{xy}}{\widetilde{\sigma }_{xx}} \right) = \frac{2}{\eta ^3} \left( \frac{\widetilde{\sigma }_{xy}}{\widetilde{\sigma }_{xx}} \right) ,\nonumber \\ \end{aligned}$$
(E7e)
$$\begin{aligned}{} & {} h_3({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \approx \frac{i}{\eta ^2} \int _0^\infty \textrm{d}t \, t \, e^{-t} \left( \frac{\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2}{\widetilde{\sigma }_{xx}} \right) \nonumber \\{} & {} \quad = \frac{i}{\eta ^2} \left( \frac{\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2}{\widetilde{\sigma }_{xx}} \right) . \end{aligned}$$
(E7f)

As \({\widetilde{G}}^R = g + h\), we have

$$\begin{aligned} {\widetilde{G}}_{xx}^R({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10})\approx & {} - \frac{\cos \eta + i \sin \eta }{2\eta } , \end{aligned}$$
(E8a)
$$\begin{aligned} {\widetilde{G}}_{xy}^R({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10})\approx & {} - \frac{1}{\eta ^3} \frac{\widetilde{\sigma }_{xy}}{\widetilde{\sigma }_{xx}} \left( \eta ^2 -2 + 2 i \eta \right) e^{i \eta } , \end{aligned}$$
(E8b)
$$\begin{aligned} {\widetilde{G}}_{zz}^R({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10})\approx & {} \frac{i}{\eta ^2} \left( \frac{\widetilde{\sigma }_{xx}^2 + \widetilde{\sigma }_{xy}^2}{\widetilde{\sigma }_{xx}} \right) \left( 1 - i \eta \right) e^{i \eta }. \nonumber \\\end{aligned}$$
(E8c)

In the low frequency regime (\(\omega _{10} < 2(2t-|u|)/\hbar \)) and high frequency regime (\(\omega _{10} > 2(2t+|u|)/\hbar \)), \(\widetilde{\sigma }_{xx} = i\widetilde{\sigma }_{xx}''\) and \(\widetilde{\sigma }_{xy} = \widetilde{\sigma }_{xy}'\) [37]. The resonant Casimir–Polder shifts for our considered dipole polarizations involve \({\textrm{Re}}\, {\widetilde{G}}_{xx}^R\), \({\textrm{Im}}\, {\widetilde{G}}_{xy}^R\) and \({\textrm{Re}}\, {\widetilde{G}}_{zz}^R\). Using (E8), we obtain their far-field limits:

$$\begin{aligned}{} & {} {\textrm{Re}}\, {\widetilde{G}}_{xx}^R({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \approx - \frac{\cos \eta }{2\eta } , \end{aligned}$$
(E9a)
$$\begin{aligned}{} & {} {\textrm{Im}}\, {\widetilde{G}}_{xy}^R({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \nonumber \\{} & {} \quad \approx - \frac{1}{\eta ^3} \frac{\widetilde{\sigma }_{xy}'}{\widetilde{\sigma }_{xx}''} \left( 2\cos \eta + 2\eta \sin \eta - \eta ^2\cos \eta \right) , \end{aligned}$$
(E9b)
$$\begin{aligned}{} & {} {\textrm{Re}}\, {\widetilde{G}}_{zz}^R({\textbf{r}}_0,{\textbf{r}}_0;\omega _{10}) \nonumber \\{} & {} \quad \approx - \frac{1}{\eta ^2} \left( \frac{(\widetilde{\sigma }_{xx}'')^2 - (\widetilde{\sigma }_{xy}')^2}{\widetilde{\sigma }_{xx}''} \right) \left( \cos \eta + \eta \sin \eta \right) .\nonumber \\ \end{aligned}$$
(E9c)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lu, BS., Arifa, K.Z. & Ducloy, M. An excited atom interacting with a Chern insulator: toward a far-field resonant Casimir–Polder repulsion. Eur. Phys. J. D 76, 210 (2022). https://doi.org/10.1140/epjd/s10053-022-00544-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epjd/s10053-022-00544-x

Navigation