Gravitationally sensitive structured x-ray optics using nuclear resonances
Abstract
Einstein’s general theory of relativity not only revolutionized human understanding of the universe, but also brought many gravitational applications in large scale, such as gravitational-wave astronomy [1], gravitational lensing [2], and the operation of the global positioning system [3]. However, it still remains a challenge to implement applications for gravitational effects at small spacial extensions on Earth. Here, we investigate a structured waveguide system that allows for the control of an x-ray profile at altitude separations of millimeters and even shorter using the nuclear resonant scattering of x rays [4, 5]. Our present results suggest a potential compact scheme for turning the Earth’s gravity into a practical application of x-ray optics.
The Pound–Rebka experiment [6] has demonstrated a unique system for probing the gravitational red-shift effect by exploiting an extremely narrow nuclear linewidth in combination of a high x-ray energy in the Mössbauer effect [7]. Moreover, advances in modern x-ray light sources and optics have raised the field of x-ray-nuclei interactions to a new level of accuracy where coherent quantum control comes into play [8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21]. A combination of nuclear quantum coherence and its sensitivity to gravity will potentially lead to a new type of x-ray optics whose performance depends on the gravitational red-shift in addition to the typical Zeeman shift [8, 12] and Doppler shift [14, 17, 18, 19]. In this context, we investigate a system that emulates the Schrödinger equation [22] and is sensitive to gravity. The present scheme allows for a systematic generation of structured x rays [23, 24, 25, 26] via changing the altitude, the x-ray photon energy, or the external magnetic field of a our system for a given waveguide structure.
(keV) | (MHz) | |||||||
---|---|---|---|---|---|---|---|---|
45Sc | 12.4 | 2.18 | 3.84 | 2.97 | 2.091 | 131.9 | 1.78 | 2.737 |
57Fe | 14.413 | 7.05 | 7.43 | 2.20 | 1.607 | 338.9 | 0.93 | 2.49 |
73Ge | 13.275 | 0.24 | 5.41 | 2.59 | 1.622 | 508.9 | 1.32 | 2.947 |
181Ta | 6.238 | 0.11 | 67.74 | 11.77 | 8.731 | 7987.3 | 32.76 | 12.616 |
182Ta | 16.273 | 2.45 | 10.41 | 1.72 | 1.304 | 1062.2 | 0.55 | 1.642 |




The system is depicted in Figure 1 with the description as follows. An x ray drives a nuclear transition from the ground state to the excited state with the total detuning in Fig. 1(a). We emphasize that the gravitational red shift has to be taken into account when the system is located at different altitude relative to where x rays are emitted. Here is the nuclear transition energy, is the gravitational constant, is the mass of the Earth, and is the average radius of the Earth. is the x-ray detuning, and the spontaneous decay rate of the excited state . Our system of nuclear resonant scattering of x rays can be described by the optical-Bloch equation [9, 21]:
(1) | |||||
(2) |
where is the coherence of a nuclear two-level system, is the x-ray electric field strength, is the reduced Planck constant, is the x-ray wavenumber, and is the index of refraction from electrons. Further, we denote the transverse Laplacian , and the coupling constant , where the length of the waveguide, is the nuclear transition dipole moment, and is the nuclear resonant thickness. The steady state of Eqs. (1-2) leads to the analytic solution and the optical Schrödinger equation [22] (see supplemental information)
(3) |
with the effective mass . Equation (3) suggests a structured x-ray waveguide (SXWG) system composed of resonant nuclei, as depicted in Fig. 1(b), with a high degree of freedom for simulating different quantum systems via spatial engineering of and . Fig. 1(c) displays the altitude dependent real part (blue-solid) and imaginary part (red-dashed line) of with for the isotope 45Sc. describes the refractive index from nuclei and results in the gravitational effects in our system as revealed by the last term of Eq. (3). In cantrast, of Lorentzian line shape represents the nuclear absorption of x rays and leads to the Pound–Rebka experiment [6].
In the following, we use the SXWG to simulate Rabi oscillations of x ray in a finite square well of width . As illustrated in Fig. 1(d), we introduce a platinum cladding for to constitute a finite square well potential, which provides with a transverse confinement to the x ray propagation direction [21]. The cladding material leads to the energy eigenfunctions of the field in Eq. (3) which read as (see supplemental information)
(4) |
with the eigen angular frequencies Inside the square well , we perturb the system by a periodic (gradient) particle distribution of isotope and carbon along the y direction (x direction), where and (see supplemental information for the form of ). In Fig. 1(d) subscripts Pt, C, and X represent platinum, carbon and resonant nucleus, respectively. We list other relevant material parameters in Table 1. With the above density modulation, the last term in Eq. (3) effectively becomes the electric dipole Hamiltonian in an oscillating field which perturbes the square well potential. This plays the key role to drive the x-ray Rabi oscillation with gravitational sensitivity. When the resonant condition is fulfilled, the periodic structure of the refractive index drives the dipole transition with the effective Rabi frequency (see supplemental information)
(5) |
A propagating x ray experiences a constant SXWG-induced Rabi frequency, and the condition for having a pulse is . We use to demonstrate the Rabi oscillation between the x-ray ground state and the first excited state by numerically solving Eq. (3).
The solution of Eq. (3) in Fig. 2 represents an x ray which propagates through an SXWG with with Sc, , natural scandium particle density m-3, nuclear resonance absorption cross section (kbarn), nm, mm, rad/m, , and . The process is visualized in terms of the fidelity in Fig. 2(a), and the normalized intra-waveguide x-ray intensity distribution in Fig. 2(b). The alternation of and in Fig. 2(a) clearly demonstrates that the ground-state x ray enters the SXWG at , and is then coherently promoted to the first excited state when approaching mm. After that, the x ray returns to the ground state and finishes a full Rabi cycle at mm. One can also observe the same phenomenon at the intra-waveguide x-ray intensity which evolves back and forth between states and in Fig.2(b).
The above x-ray Rabi oscillation suggests a systematic way to generate the mode with a sequence of SXWGs driving transitions, where one can raise the quantum number one by one. Specifically, one can connect two different SXWGs to accomplish a transition, where the upstream SXWG drives a transition, and the downstream SXWG achieves a promotion. Moreover, one can even design multiple SXWG modules for or any dipole forbidden transitions. Thus, all combinations of SXWG modules open the capability to generate high order x-ray modes starting from the ground state . It is worth mentioning another possible application using dual SXWGs with a gap in between as an x-ray interferometer. While the upstream SXWG causes the transition as a beam splitter, the downstream SXWG leads to the return of as a beam combiner. Furthermore, in the gap between two SXWGs one can introduce a phase modulator to impose a phase shift at one branch of the split state , e.g., at mm in Fig. 2. A controllable interference due to the phase modulation is expected to happen at the end of the downstream SXWG.
We are ready to demonstrate the Earth’s gravitational effect on our SXWG system. Given that the gravitational redshift significantly changes the nuclear coherence and Rabi frequency in Eq. (5) in two millimeter on Earth as demonstrated in Fig. 1(c), this sensitivity potentially allows for turning gravity into a practical use, e.g., gravitationally sensitive x-ray optics. For illustrating the effect, we numerically solve Eq. (3) and use the isotope 45Sc in an SXWG with parameters , nm, mm, rad/m, and to show the gravitational effect. Fig. 4(a) illustrates three cases where the above discussed SXWG is located at cm, cm, and cm from the top down. An incident x ray with the transverse mode is deflected upward and experiences a gravitational redshift (vertical upward arrow with color gradient). The will change when the x ray illuminates the SXWG at different altitudes. The total detuning for each case is specified at the level-scheme plot, namely, , , and from the top down. We emphasize that the periodic particle density modulation effectively plays the role of a resonant field, and it always resonantly drives a transition between the x-ray modes in a cladding waveguide for three cases. However, various change the effective coupling strength and result in different outputs. The scattered/split x rays reflect the output mode and can be measured by a downstream position-sensitive detector. The -dependent photon number counts show the output and reveal the Earth’s gravitational effect. We depict the normalized for cm, cm, and cm in Fig. 4(c, e, and g), respectively. The intra-waveguide intensity shows that the x ray significantly gets split in Fig. 4(e) under a half Rabi cycle in the SXWG at cm. In contrast, Fig. 4(c and g) depicts only a transverse broadening of the x ray due to a small . Fig. 4(b,d, and f) illustrate (red-dashed line) and (green-solid line) for cm, cm, and cm, respectively. One can clearly see that the x ray experiences Rabi flopping and becomes at the resonant altitude cm as also pointed out by Fig. 4(d and e). Given that decreases when the SXWG leaves the resonant altitude, the x ray mostly remains in the initial mode , namely, at cm and cm. We depict the output altitude-dependent at mm in Fig. 4(a). As a result, different x-ray splitting is expected to occur when lifting an SXWG composed of 45Sc at only a millimeter altitude change.
We quantify the gravitational sensitivity of the SXWG by the full altitude width at the half maximum
(6) |
as indicated by the black-horizontal double arrow in Fig. 1(c). The introduced is a measure for the sensitivity of the x-ray-nucleus coupling to the change of the SXWG vertical location. With the definition of the quality factor of a nuclear resonance , we can see that is proportional to . Fig. 4(b) exemplifies the implication of Eq. (6) for our system on Earth in a double-logarithmic plot, where we mark the isotopes 45Sc, 57Fe, 67Zn, 73Ge, 103Rh, 107Ag, 109Ag, 181Ta, 182Ta, 229Th, and 249Bk, according their factor. Some of the nuclear parameters are listed in Table 1. Remarkably, the advantage of a very high of 45Sc or 182Ta nuclear resonance endues an SXWG with a gravitational sensitivity to only millimeter altitude change. Notably, it is also possible to get sub-millimeter using 229Th whose [29], and micron with 107Ag and 109Ag whose . It deserves to mention that 103Rh whose [27] even results in a nanometer , and it may lead to gravitational application in mesoscopic scale.
In conclusion we have put forward a controllable SXWG system that potentially turns gravity into an application of x-ray optics. A periodic intra-waveguide structure, e.g., nuclear optical lattice 57Fe/56Fe bilayers in Ref.[30], can drive a transition between x-ray modes. The x-ray transverse mode experiences Rabi oscillation when propagating in an SXWG. Our scheme allows for applications like a systematic production of structured x rays and an x-ray interferometer without any beam splitter. Remarkably, a significant change of a gravitionally induced splitting of x rays could happen by lifting our SXWG made of, e.g., 45Sc or 182Ta, at only a millimeter scale.
S.-Y. L. and W.-T. L. are supported by the National Science and Technology Council of Taiwan (Grant No. 110-2112-M-008-027-MY3, 110-2639-M-007 -001-ASP, 111-2923-M-008-004-MY3 & 111-2639-M-007-001-ASP). S. A. is supported by National Science Foundation of China (Grant No. 11975155).
References
- Abbott and et al. [2016] B. P. Abbott and et al. (LIGO Scientific Collaboration and Virgo Collaboration), Observation of gravitational waves from a binary black hole merger, Phys. Rev. Lett. 116, 061102 (2016).
- Bartelmann and Schneider [2001] M. Bartelmann and P. Schneider, Weak gravitational lensing, Physics Reports 340, 291 (2001).
- Ashby [2002] N. Ashby, Relativity and the global positioning system, Physics Today 55, 41 (2002).
- Hastings et al. [1991] J. B. Hastings, D. P. Siddons, U. van Bürck, R. Hollatz, and U. Bergmann, Mössbauer spectroscopy using synchrotron radiation, Phys. Rev. Lett. 66, 770 (1991).
- Röhlsberger [2004] R. Röhlsberger, Nuclear Condensed Matter Physics With Synchrotron Radiation: Basic Principles, Methodology and Applications (Springer-Verlag, 2004).
- Pound and Rebka [1960] R. V. Pound and G. A. Rebka, Apparent weight of photons, Phys. Rev. Lett. 4, 337 (1960).
- Kalvius and Kienle [2012] M. Kalvius and P. Kienle, The Rudolf Mössbauer Story: His Scientific Work and Its Impact on Science and History (Springer Science & Business Media, 2012).
- Shvyd’ko et al. [1996] Y. V. Shvyd’ko, T. Hertrich, U. van Bürck, E. Gerdau, O. Leupold, J. Metge, H. D. Rüter, S. Schwendy, G. V. Smirnov, W. Potzel, and P. Schindelmann, Storage of nuclear excitation energy through magnetic switching, Phys. Rev. Lett. 77, 3232 (1996).
- Shvyd’ko [1999] Y. V. Shvyd’ko, Nuclear resonant forward scattering of x rays: Time and space picture, Phys. Rev. B 59, 9132 (1999).
- Pálffy et al. [2009] A. Pálffy, C. H. Keitel, and J. Evers, Single-photon entanglement in the kev regime via coherent control of nuclear forward scattering, Phys. Rev. Lett. 103, 017401 (2009).
- Röhlsberger et al. [2012] R. Röhlsberger, H. C. Wille, K. Schlage, and B. Sahoo, Electromagnetically induced transparency with resonant nuclei in a cavity, Nature 482, 199 (2012).
- Liao et al. [2012] W.-T. Liao, A. Pálffy, and C. H. Keitel, Coherent storage and phase modulation of single hard-x-ray photons using nuclear excitons, Phys. Rev. Lett. 109, 197403 (2012).
- Adams et al. [2013] B. W. Adams, C. Buth, S. M. Cavaletto, J. Evers, Z. Harman, C. H. Keitel, A. Pálffy, A. Picón, R. Röhlsberger, Y. Rostovtsev, and K. Tamasaku, X-ray quantum optics, Journal of modern optics 60, 2 (2013).
- Vagizov et al. [2014] F. Vagizov, V. Antonov, Y. Radeonychev, R. N. Shakhmuratov, and O. Kocharovskaya, Coherent control of the waveforms of recoilless -ray photons, Nature 508, 80 (2014).
- Liao and Ahrens [2015] W.-T. Liao and S. Ahrens, Gravitational and relativistic deflection of x-ray superradiance, Nature Photonics 9, 169 (2015).
- Heeg et al. [2015] K. P. Heeg, J. Haber, D. Schumacher, L. Bocklage, H.-C. Wille, K. S. Schulze, R. Loetzsch, I. Uschmann, G. G. Paulus, R. Rüffer, R. Röhlsberger, and J. Evers, Tunable subluminal propagation of narrow-band x-ray pulses, Phys. Rev. Lett. 114, 203601 (2015).
- Heeg et al. [2017] K. P. Heeg, A. Kaldun, C. Strohm, P. Reiser, C. Ott, R. Subramanian, D. Lentrodt, J. Haber, H.-C. Wille, S. Goerttler, R. Rüffer, C. H. Keitel, R. Röhlsberger, T. Pfeifer, and J. Evers, Spectral narrowing of x-ray pulses for precision spectroscopy with nuclear resonances, Science 357, 375 (2017).
- Zhang et al. [2019] X. Zhang, W.-T. Liao, A. Kalachev, R. Shakhmuratov, M. Scully, and O. Kocharovskaya, Nuclear quantum memory and time sequencing of a single photon, Phys. Rev. Lett. 123, 250504 (2019).
- Radeonychev et al. [2020] Y. V. Radeonychev, I. R. Khairulin, F. G. Vagizov, M. Scully, and O. Kocharovskaya, Observation of acoustically induced transparency for -ray photons, Phys. Rev. Lett. 124, 163602 (2020).
- Heeg et al. [2021] K. P. Heeg, A. Kaldun, C. Strohm, C. Ott, R. Subramanian, D. Lentrodt, J. Haber, H.-C. Wille, S. Goerttler, R. Rüffer, et al., Coherent x-ray- optical control of nuclear excitons, Nature 590, 401 (2021).
- Chen et al. [2022] Y.-H. Chen, P.-H. Lin, G.-Y. Wang, A. Pálffy, and W.-T. Liao, Transient nuclear inversion by x-ray free electron laser in a tapered x-ray waveguide, Phys. Rev. Research 4, L032007 (2022).
- Marte and Stenholm [1997] M. A. M. Marte and S. Stenholm, Paraxial light and atom optics: The optical schrödinger equation and beyond, Phys. Rev. A 56, 2940 (1997).
- Seipt et al. [2014] D. Seipt, A. Surzhykov, and S. Fritzsche, Structured x-ray beams from twisted electrons by inverse compton scattering of laser light, Phys. Rev. A 90, 012118 (2014).
- Rubinsztein-Dunlop et al. [2016] H. Rubinsztein-Dunlop, A. Forbes, M. V. Berry, M. R. Dennis, D. L. Andrews, M. Mansuripur, C. Denz, C. Alpmann, P. Banzer, T. Bauer, et al., Roadmap on structured light, Journal of Optics 19, 013001 (2016).
- MacDonald [2017] C. A. MacDonald, Structured x-ray optics for laboratory-based materials analysis, Annual Review of Materials Research 47, 115 (2017).
- Forbes et al. [2021] A. Forbes, M. de Oliveira, and M. R. Dennis, Structured light, Nature Photonics 15, 253 (2021).
- [27] National Nuclear Data Center, https://www.nndc.bnl.gov/ensdf/.
- [28] The Center for X-Ray Optics, https://www.cxro.lbl.gov/.
- Seiferle et al. [2019] B. Seiferle, L. von der Wense, P. V. Bilous, I. Amersdorffer, C. Lemell, F. Libisch, S. Stellmer, T. Schumm, C. E. Düllmann, A. Pálffy, et al., Energy of the 229th nuclear clock transition, Nature 573, 243 (2019).
- Haber et al. [2016] J. Haber, K. S. Schulze, K. Schlage, R. Loetzsch, L. Bocklage, T. Gurieva, H. Bernhardt, H.-C. Wille, R. Rüffer, I. Uschmann, et al., Collective strong coupling of x-rays and nuclei in a nuclear optical lattice, Nature Photonics 10, 445 (2016).