Rydberg Single Photon Detection for Probing 0.1–10 meV Dark Matter with BREAD

Abhishek Banerjee ID [email protected] Maryland Center for Fundamental Physics, University of Maryland, College Park, MD 20742, USA    Reza Ebadi ID [email protected] Department of Physics & Astronomy, The Johns Hopkins University, Baltimore, MD 21218, USA Department of Physics and Astronomy, University of Delaware, Newark, Delaware 19716, USA Department of Physics, University of Maryland, College Park, Maryland 20742, USA    Surjeet Rajendran ID [email protected] Department of Physics & Astronomy, The Johns Hopkins University, Baltimore, MD 21218, USA
Abstract

We introduce a Rydberg–based single photon detector (SPD) for probing dark matter in the 0.1–10 meV mass range (20 GHz–2 THz). The Rydberg SPD absorbs photons produced and focused by the BREAD dish antenna and trades them for free, detectable electrons. At the lower end of the mass range, photons drive Rydberg–Rydberg transitions, which are read out via state-selective ionization. At higher masses, they directly ionize the Rydberg atoms.

Light bosonic dark matter candidates (e.g. axion, CP-even scalars, and dark photon) are well described as coherent classical fields Kolb and Turner (2019); Cheong et al. (2025). Through their coupling to electromagnetism, a small fraction of the background dark matter (DM) converts into coherent photons. The non-relativistic nature of DM implies that the photon frequencies are narrowly distributed around the dark matter mass ωmDM\omega\simeq m_{\rm DM}. Numerous existing and proposed experiments are designed to search for such DM–induced ac electromagnetic signal Adams et al. (2022); Sushkov (2023).

Refer to caption
Figure 1: Schematic of the Rydberg SPD setup. (a) BREAD uses a parabolic reflector to focus photons emitted perpendicular to the barrel surface onto a focal point, increasing the photon intensity at the focus. (b) A Rydberg beam is prepared by exciting electrons to a high principal quantum number. When the photon energy exceeds the ionization threshold ω>In\omega>I_{n} the photons directly ionize the Rydberg atoms, producing detectable free electrons. (c) The target photon energy matches the transition energy between Rydberg states, ω=ωn\omega=\omega_{n}. During the sensing stage, the Rydberg beam is exposed to photons from the BREAD setup. Downstream, the Rydberg state is read out using selective field ionization (SFI). In SFI, an external electric field is scanned across the ionization threshold of the relevant states, and the resulting ionization is detected.

The Broadband Reflector Experiment for Axion Detection (BREAD) employs a dish antenna with a parabolic reflector to focus photons generated by DM interactions onto a focal point Liu et al. (2022). BREAD is a modular broadband signal-enhancement platform that must be supplemented with a high-precision readout module. Developing efficient detectors continues to be an active area of research. A key focus is the development of single photon detectors (SPDs), as photon-counting is expected to achieve higher sensitivity in the low-signal regime compared to bolometric detection Sushkov (2023).

SPDs at optical and infrared frequencies are mature technologies Hadfield (2009). However, efficient single photon detection in the THz range remains challenging Todorov et al. (2024). Photons in this frequency range carry energies on the meV scale, far lower than the eV-scale energies of optical photons. Thus, well-established optical SPDs lack sensitivity in this range. Addressing this challenge requires development of new detection technologies. The search for meV-scale dark matter has recently driven further development of THz SPDs Fan et al. (2025).

Graham et al. proposed a Rydberg-based SPD for the HAYSTAC experiment to probe DM in the 40–200 μeV{\rm\mu eV} mass range (10–50 GHz) Graham et al. (2024). In this setup, DM–induced photons drive transitions between Rydberg states, and the resulting state population is read out via selective field ionization (SFI). Inspired by this approach, we introduce a Rydberg-atom–based SPD operating in the 20 GHz–2 THz range as a complementary component of the BREAD experiment. Rydberg electronic energy levels span this frequency range, and their large dipole coupling enhances photon absorption efficiency. A photon absorbed by a Rydberg atom can either directly ionize it or excite it to a higher Rydberg state (detectable via state-selective field ionization); see Fig. 1.

One of our proposed detection schemes (Rydberg transitions induced by photons followed by state readout via SFI) is similar to that in Ref. Graham et al. (2024). However, our approach differs in several key aspects. First, the direct ionization scheme we introduce is entirely new and enables sensitivity to DM masses substantially higher than those accessible through transition-based methods. Second, while Ref. Graham et al. (2024) considers a microwave cavity setup with separate cavities for signal buildup and detection, our design employs the broadband reflector configuration of BREAD. This setup is considerably simpler, as it removes the need for frequency scanning and allows searches at higher frequencies using Rydberg ionization, since the signal strength is no longer limited by the cavity size.

Photons from dark matter.— We focus on two specific DM candidates: the axion and the dark photon. The relevant electromagnetic interaction Lagrangian takes the form gaγγaFF~/4=gaγγa𝐄𝐁\mathcal{L}\supset-g_{a\gamma\gamma}aF\tilde{F}/4=g_{a\gamma\gamma}a\mathbf{E}\cdot\mathbf{B} for the axion, and εFF/2\mathcal{L}\supset\varepsilon FF^{\prime}/2 for the dark photon. Here, aa is the pseudoscalar axion field, gaγγg_{a\gamma\gamma} the axion–photon coupling, ε\varepsilon the dimensionless dark photon–photon mixing parameter, FF the electromagnetic field strength, and FF^{\prime} the dark photon field strength. When a large bias magnetic field 𝐁0\mathbf{B}_{0} is applied, ambient axion DM produces an oscillating electric field with amplitude EagaγγB0ρDM/mDME_{\rm a}\sim g_{a\gamma\gamma}B_{0}\sqrt{\rho_{\rm DM}}/m_{\rm DM}. For the dark photon, the mixing automatically produces a similar oscillating electric field EAερDME_{A^{\prime}}\sim\varepsilon\sqrt{\rho_{\rm DM}} where ρDM\rho_{\rm DM} denotes the local DM energy density.

Coherent detection of the ac electric field is linearly sensitive to the couplings and has been proposed for DM detection using electro-optics Ebadi et al. (2024). In the mass range of interest for this work, however, the photon production by BREAD occurs incoherently. For mDM0.1meVm_{\rm DM}\gtrsim 0.1\,{\rm meV}, the DM de Broglie wavelength is smaller than the size of the BREAD setup Liu et al. (2022). As a result, photons emitted from different parts of the barrel surface are out of phase, producing an incoherent signal. We will therefore focus on single photon detection, which has quadratic sensitivity to the couplings. While the standard quantum limit sets the fundamental limit for coherent detection, the dark count rate (DCR) is the primary figure of merit for SPDs.

The BREAD setup will have a dish area of Adish10m2A_{\rm dish}\sim 10\,{\rm m}^{2}. The focus area AfocusA_{\rm focus} is smeared out compared to an ideal point due to the momentum spread of the emitted photons, inherited from the momentum spread of the DM. With a typical momentum spread of ΔpDM/mDM103\Delta p_{\rm DM}/m_{\rm DM}\sim 10^{-3} and a physical BREAD setup size of about a meter, the resulting focus area is estimated as Afocusπmm2A_{\rm focus}\sim\pi\,{\rm mm}^{2} Liu et al. (2022). Note that the photon field amplitude at the focal point is enhanced by a factor of Adish/Afocus𝒪(103)\sqrt{A_{\rm dish}/A_{\rm focus}}\sim\mathcal{O}(10^{3}) relative to its value away from the focus.

We compute the signal photon rate by estimating the emitted power PDM=EDM2Adish/2P_{\rm DM}=E_{\rm DM}^{2}A_{\rm dish}/2 where AdishA_{\rm dish} is the effective emission area. The corresponding photon rate is RDM=PDM/mDMR_{\rm DM}=P_{\rm DM}/m_{\rm DM}. Therefore, the expected signal photon rates are Ragaγγ2B02ρDMAdish/2ma3R_{a}\simeq g_{a\gamma\gamma}^{2}B_{0}^{2}\rho_{\rm DM}A_{\rm dish}/2m_{a}^{3} and RAε2ρDMAdish/2mAR_{A^{\prime}}\simeq\varepsilon^{2}\rho_{\rm DM}A_{\rm dish}/2m_{A^{\prime}}. For fixed external parameters B0=10TB_{0}=10~\mathrm{T}, ρDM=0.45GeV/cm3\rho_{\mathrm{DM}}=0.45~\mathrm{GeV/cm^{3}}, and Adish=10m2A_{\mathrm{dish}}=10~\mathrm{m^{2}}, these expressions provide numerical estimates of the photon rates as a function of couplings and masses Liu et al. (2022):

Ra\displaystyle R_{a} 0.55Hz(gaγγ1011GeV1)2(mameV)3,\displaystyle\simeq 0.55~\mathrm{Hz}\bigg(\frac{g_{a\gamma\gamma}}{10^{-11}~\mathrm{GeV^{-1}}}\bigg)^{2}\bigg(\frac{m_{a}}{\mathrm{meV}}\bigg)^{-3},
RA\displaystyle R_{A^{\prime}} 0.14Hz(ε1014)2(mAmeV)1.\displaystyle\simeq 0.14~\mathrm{Hz}\bigg(\frac{\varepsilon}{10^{-14}}\bigg)^{2}\bigg(\frac{m_{A^{\prime}}}{\mathinner{\mathrm{meV}}}\bigg)^{-1}.

DM–photon conversion can be further enhanced by introducing a stack of NlN_{l} dielectric layers, which effectively increases the number of conversion surfaces Jaeckel and Redondo (2013); Caldwell et al. (2017); Baryakhtar et al. (2018). In the BREAD setup, this approach can enhance the photon rate by a factor of Nl2/2N_{l}^{2}/2, where the optimal number of layers is Nl80mDM/meVN_{l}\simeq 80\sqrt{m_{\rm DM}/{\rm meV}} Fan et al. (2025). Thus, the signal photon rate will be enhanced by a factor of 103(mDM/meV)10^{3}(m_{\rm DM}/{\rm meV}). At the same time, the required semi-resonant buildup across the layers limits the emitted radiation to QNl/NsQ\sim N_{l}/N_{s}, where NsN_{s} is the number of stacks with different spacings. Ref. Fan et al. (2025) identified that for the optimal case Q17Q\sim 17, well below the intrinsic quality factor of the DM wave QDM106Q_{\rm DM}\sim 10^{6}.

Table 1: Scaling behavior of Rydberg quantities.
Quantity Symbol Scaling
Typical orbital size rn\langle r\rangle_{n} n2/meα\sim n^{2}/m_{e}\alpha
Binding energy InI_{n} meα2/2n2\sim m_{e}\alpha^{2}/2n^{2}
Transition frequency ωn\omega_{n} meα2/n3\sim m_{e}\alpha^{2}/n^{3}
Radiative lifetime τn\tau_{n} {100(n/40)3μs,l=110(n/40)5ms,l=n1\sim\begin{cases}100\,(n/40)^{3}\,{\rm\mu s},&l=1\\ 10\,(n/40)^{5}\,{\rm ms},&l=n-1\end{cases}
dc ionizing field EionE_{\rm ion} 30n4GV/m\sim 30\,n^{-4}\,{\rm GV/m}

Rydberg SPD.— A Rydberg atom has an electron excited to a very high principal quantum number n1n\gg 1 Gallagher (1994). Rydberg atoms exhibit exaggerated physical properties; see Table 1. A defining feature is their large orbital size, which scales as n2a0\sim n^{2}a_{0}, where a0a_{0} is the Bohr radius 111For Rydberg atoms, the deviation from the pure Coulomb potential due to the presence of core electrons is incorporated through the quantum defect, δ\delta_{\ell}, which depends on \ell. Thus an effective description is obtained by replacing nn=nδl,=lδl+I(l)n\to n^{*}=n-\delta_{l},\ell\to\ell^{*}=l-\delta_{l}+I(l) where I()I(\ell) is some integer Kostelecký and Nieto (1985). This is a small correction to the hydrogenic values and do not alter the physics discussed here.. This leads to strong dipole couplings, both between Rydberg atoms and with photons. We exploit the strong photon–Rydberg interaction to detect DM–induced photons, for the same reason it has been used in similar applications Jau and Carter (2020); Meyer et al. (2020); Li et al. (2024); Nill et al. (2024); Liu et al. (2023).

We consider two Rydberg-based sensing modalities: Rydberg–Rydberg transitions and Rydberg–continuum ionization (see Fig. 1). Lower-energy photons (0.1–1 meV) can drive Rydberg–Rydberg transitions, while higher-energy photons (1–10 meV) can directly ionize electrons from the Rydberg state. In the Rydberg–Rydberg transition case, the final state can be read out via state-selective field ionization (SFI) Hollenstein et al. (2001); Tada et al. (2002); Gürtler and van der Zande (2004); Feynman et al. (2015); Gregoric et al. (2017); Alonso et al. (2018); Gregoric et al. (2018). In SFI, an external electric field is ramped up until the atom ionizes, and the free electron is detected. Different Rydberg levels ionize at distinct field strengths, providing a correspondence between the applied field and the state. In the direct ionization case, the electron is detected immediately.

Rydberg states with lifetimes τn10μs\tau_{n}\gtrsim 10\,\mu{\rm s} exhibit narrow transitions with quality factor QRyd=ωn/(2π/τn)107Q_{\rm Ryd}=\omega_{n}/(2\pi/\tau_{n})\gtrsim 10^{7}. Because this exceeds the DM-induced photon quality factor QDM106Q_{\rm DM}\sim 10^{6} for BREAD and 20\sim 20 for BREAD with dielectric layers, the effective search bin width is determined by the photon linewidth rather than the transition linewidth. In contrast, ionization occurs whenever the DM-induced photon energy exceeds the ionization threshold InI_{n} and lies within the effective ionization range. The ionization search is therefore broadband, determined by the readout scheme and independent of the signal linewidth.

The transition energy between neighboring Rydberg states is ωn27.2eV/n3\omega_{n}\simeq 27.2\,{\rm eV}/n^{3}, corresponding to n(30,65)n\in(30,65) for the 0.1–1 meV DM mass range. The binding energy is In13.6eV/n2I_{n}\simeq 13.6\,{\rm eV}/n^{2}, corresponding to n(37,120)n\in(37,120) for the 1–10 meV range. Note that not all of these nn values are required to cover the DM mass range of interest; we discuss the full mass coverage strategy later.

The rates of transition and ionization relative to the photon production rate serve as a figure of merit for how efficiently the BREAD-produced photons are collected. We compute these rates for low-ll states using Fermi’s golden rule; see the Supplemental Material for details. For a single Rydberg atom, the transition rate between a Rydberg state nn and a nearby state n+in+i, where ini\ll n, is γnn+i(2πα/3)EDM2n4a02min[τDM,τn,τsens.]\gamma_{n\rightarrow n+i}\simeq(2\pi\alpha/3)E_{\rm DM}^{2}\,n^{4}a_{0}^{2}\,{\rm min}[\tau_{\rm DM},\tau_{n},\tau_{\rm sens.}]. Here, τDM,τn\tau_{\rm DM},\tau_{n}, and τsens.\tau_{\rm sens.} are the DM-induced photon coherence time, the Rydberg state lifetime, and the time each atom spends in the sensing (focal) region, respectively. The Rydberg ionization rate is better understood away from the ionization threshold Bethe and Salpeter (1957), since Coulomb corrections become significant near the threshold. We adopt a conservative estimate γncont.(1024π/3)EDM2n4a03(In/mDM)6(mDM/In1)3/2\gamma_{n\rightarrow{\rm cont.}}\simeq(1024\pi/3)E_{\rm DM}^{2}n^{4}a^{3}_{0}\,(I_{n}/m_{\rm DM})^{6}\left(m_{\rm DM}/I_{n}-1\right)^{3/2}. The ionization rate has broad support in DM mass, with a power-law fall-off. The maximum value is γncont.max36EDM2n4a03\gamma_{n\rightarrow{\rm cont.}}^{\rm max}\simeq 36E_{\rm DM}^{2}n^{4}a^{3}_{0}. The estimated rates are consistent in order of magnitude with values reported in the literature, and show only mild ll-dependence for low-ll, high-nn states Ovsiannikov et al. (2011). The transition rates between nearby nn levels also exhibit only subleading dependence on ll. This is unlike the radiative lifetime that has a strong ll-dependence. The reason is that the dominant radiative decay channels for low-ll states correspond to transitions to much lower nn states (for which a larger phase space is available), and as a result, the dipole matrix element acquires a distinct overall nn-scaling behavior.

EDME_{\rm DM} is the photon field amplitude at the focusing area EDM2=2RDMmDM/AfocusE_{\rm DM}^{2}=2R_{\rm DM}m_{\rm DM}/A_{\rm focus}. Key advantage of a focusing setup such as BREAD is manifest in these rates: the photon field amplitude at the detector site is enhanced. Since RDMAdishR_{\rm DM}\propto A_{\rm dish}, the rate enhancement scales as Adish/Afocus3×106A_{\rm dish}/A_{\rm focus}\sim 3\times 10^{6} for typical BREAD parameters.

In our proposed method, a large number of Rydberg atoms simultaneously interact with photons in the focused area. Assuming a Rydberg beam with flux Φbeam1015cm2s1\Phi_{\rm beam}\sim 10^{15}\,{\rm cm^{-2}s^{-1}} and velocity vbeamkm/sv_{\rm beam}\sim{\rm km/s}, the total number of Rydberg atoms present at any instant in the sensing volume is NRydΦbeamAfocusτsens.3×107N_{\rm Ryd}\simeq\Phi_{\rm beam}A_{\rm focus}\tau_{\rm sens.}\sim 3\times 10^{7}, where τsens.=Afocus/π/vbeamμs\tau_{\rm sens.}=\sqrt{A_{\rm focus}/\pi}/v_{\rm beam}\simeq{\rm\mu s}.

The total photon absorption rate is given by the single-atom rate multiplied by the number of Rydberg atoms Γ=NRydγ\Gamma=N_{\rm Ryd}\gamma. We define efficiency of photon absorption as η=Γ/RDM\eta=\Gamma/R_{\rm DM}. For BREAD (QDM106Q_{\rm DM}\sim 10^{6}), the limiting timescale is τsens.\tau_{\rm sens.}: min[τDM,τn,τsens.]=τsens.{\rm min}[\tau_{\rm DM},\tau_{n},\tau_{\rm sens.}]=\tau_{\rm sens.}. For BREAD with the dielectric layers (QDM20Q_{\rm DM}\sim 20), the limiting timescale is instead the DM-induced photon coherence time: min[τDM,τn,τsens.]=τDM{\rm min}[\tau_{\rm DM},\tau_{n},\tau_{\rm sens.}]=\tau_{\rm DM}. The corresponding nominal efficiencies are:

ηnn+iBREAD\displaystyle\eta_{n\rightarrow n+i}^{\rm BREAD} 1.5×103(mDM0.3meV)1/3(Φbeam1015cm2s1),\displaystyle\simeq 1.5\times 10^{3}\bigg(\frac{m_{\rm DM}}{0.3\,{\rm meV}}\bigg)^{-1/3}\bigg(\frac{\Phi_{\rm beam}}{10^{15}\,{\rm cm^{-2}s^{-1}}}\bigg),
ηnn+iDielectric\displaystyle\eta_{n\rightarrow n+i}^{\rm Dielectric} 0.3(QDM17)(mDM0.3meV)4/3(Φbeam1015cm2s1),\displaystyle\simeq 0.3\bigg(\frac{Q_{\rm DM}}{17}\bigg)\bigg(\frac{m_{\rm DM}}{0.3\,{\rm meV}}\bigg)^{-4/3}\bigg(\frac{\Phi_{\rm beam}}{10^{15}\,{\rm cm^{-2}s^{-1}}}\bigg),
ηncont.max\displaystyle\eta_{n\rightarrow{\rm cont.}}^{\rm max} 3×105(mDM3meV)1(Φbeam1015cm2s1),\displaystyle\simeq 3\times 10^{-5}\bigg(\frac{m_{\rm DM}}{3\,{\rm meV}}\bigg)^{-1}\bigg(\frac{\Phi_{\rm beam}}{10^{15}\,{\rm cm^{-2}s^{-1}}}\bigg),

where we have used the scaling relations mDMn3m_{\rm DM}\propto n^{-3} for transition and mDMn2m_{\rm DM}\propto n^{-2} for ionization. The efficiency cannot exceed unity, so the effective efficiency is simply taken to be the smaller of the nominal value and one.

To summarize, in the high-Q resonant modality of BREAD, the Rydberg SPD can detect essentially all of the photons produced. In the lower-Q BREAD setup with dielectric layers, the Rydberg SPD remains highly efficient but may not absorb every photon. However, since this search is broadband compared to the high-Q modality, the overall sensitivity improves due to relatively longer averaging time available per experiment. In the ionization modality, the Rydberg SPD collects only a small fraction of the photons, but since the search is even broader in bin size, it still provides competitive sensitivity across the higher-mass parameter space.

Refer to caption
Figure 2: Dark matter mass range accessible using Rydberg states with principal quantum number nn. Top: Rydberg transition coverage (without scanning) for n=30,31,,65n=30,31,\cdots,65. We include transitions of nn+in\rightarrow n+i within the 0.1-1 meV range. The inset shows the distribution of mass gaps across all available transitions. The largest uncovered gap in mass coverage is 0.02 meV, which can be bridged using a perturbative scanning scheme in about 0.02meV×QDM/mDM4×1040.02\,{\rm meV}\times Q_{\rm DM}/m_{\rm DM}\sim 4\times 10^{4} steps for QDM106Q_{\rm DM}\sim 10^{6}. Bottom: Ionization is efficient for photons with frequencies above the ionization threshold, with a power-law fall-off at higher frequencies. This enables broadband sensitivity using only a small number of atoms. Using four different Rydberg states allows coverage of the 0.1–1 meV DM mass range.

Experimental feasibility.— Rydberg states with nn up to 100 can be readily prepared Jau and Carter (2020). Toward the upper end of this range, however, Rydberg states become increasingly fragile to stray electric fields. The ionizing dc electric field is 3V/cm\sim 3\,\mathrm{V/cm} for n100n\sim 100, which can be shielded against. At lower nn, the ionizing electric field amplitude increases rapidly as n4n^{-4}.

Rydberg states are relatively long-lived Branden et al. (2009); Mack et al. (2015); Hölzl et al. (2024). Their main decay channels are spontaneous emission and blackbody-radiation (BBR)–induced transitions. The BBR rate is suppressed relative to spontaneous decay by the Boltzmann factor 1/(eω/kBT1)1/(e^{\omega/k_{B}T}-1), which is 102\lesssim 10^{-2} at T1KT\sim 1\,{\rm K} in the frequency range of interest. At such low temperatures, the lifetime is limited by spontaneous decay. For circular states (l=n1l=n-1) it is expected to be 10(n/40)5ms10(n/40)^{5}\,{\rm ms}, while for PP states (l=1l=1) it is shorter, 100(n/40)3μs100(n/40)^{3}\,\mu{\rm s} Šibalić et al. (2017). The reduced lifetime of low-angular-momentum states arises from their larger overlap with low-nn core states.

In both detection modalities (direct ionization and transition followed by SFI) the freed electron must be collected. A channel electron multiplier (CEM) can be used for this purpose Graham et al. (2024). CEMs are widely used in similar contexts, for example in photoionization detection of a single atom Henkel et al. (2010). Imaging of electrons and ions produced by SFI has also been demonstrated Fahey et al. (2015) using microchannel plates Beam Imaging Solutions () (BIS). Commercial devices Dr. Sjuts Optotechnik GmbH achieve near-unity efficiency with dark count rate (DCR) of about 0.01Hz0.01\,{\rm Hz}. Typical operating bias voltage for CEMs is 100V\gtrsim 100\,{\rm V}. For the highest-nn states considered here, this corresponds to an electric field that can approach or exceed the ionization threshold. To avoid this, we propose using a low-voltage extraction electrode near the sensing region, followed by electrostatic steering of the electron toward the CEM. Local acceleration at the device’s operating voltage can then raise the electron to the detection threshold energy.

The Rydberg readout module operates at cryogenic temperatures to suppress BBR–induced backgrounds. On the lower end of the mass range of interest the required temperature to keep the BBR background rate below the electronic DCR of 0.01 Hz is approximately 50 mK Graham et al. (2024). At higher masses, this requirement becomes less stringent, scaling as 500mK(mDM/meV)500\,{\rm mK}\,(m_{\rm DM}/{\rm meV}).

A Rydberg beam can be prepared from a high-intensity atomic beam with fluxes in the range Φ1011\Phi\sim 10^{11}1016cm2s110^{16}\,{\rm cm^{-2}s^{-1}} Larsen et al. (1974); Catani et al. (2006); Wei et al. (2022). Preparation can be achieved using two- or three-photon transitions Tate and Gallagher (2018); Fahey et al. (2015) or stimulated Raman adiabatic passage (STIRAP) Cubel et al. (2005). We also note that Rydberg beams composed of low-ll states are the most straightforward to prepare, since with each absorbed photon only transitions with Δl=±1\Delta l=\pm 1 are allowed by the dipole selection rules. Our proposal therefore makes use of these readily accessible Rydberg states.

Given a Rydberg beam flux of Φ1015cm2s1\Phi\sim 10^{15}\,{\rm cm^{-2}s^{-1}} and a beam velocity vRydkm/sv_{\rm Ryd}\sim{\rm km/s}, the Rydberg atom number density is nRyd1010cm3n_{\rm Ryd}\sim 10^{10}\,{\rm cm^{-3}}. This corresponds to an average spacing of about 5μm5\,{\rm\mu m} between atoms, much larger than the size of the largest Rydberg state considered here (n=100n=100) which is 104a00.5μm10^{4}a_{0}\simeq 0.5\,{\rm\mu m}. Therefore, the dipole–dipole interactions are suppressed Saffman et al. (2010). Another potential factor that can contribute to efficiency reduction or backgrounds is atomic collisions within the Rydberg beam, which can be mitigated by engineering low-temperature beams. Ultimately, magneto-optical trap (MOT) cooling combined with conveyor-belt atomic transport featuring a continuously high reloading rate can be employed as the sensing beam in our proposal Kuppens et al. (2000); Klostermann et al. (2022); Matthies et al. (2024); Chiu et al. (2025).

The length of the Rydberg beam is τnvRyd10cm\tau_{n}v_{\rm Ryd}\gtrsim 10\,{\rm cm}, assuming a lifetime τn100μs\tau_{n}\gtrsim 100\,{\rm\mu s}. For both Rydberg preparation and SFI detection we assume operation times of about μs{\rm\mu s}, which are much shorter than τn\tau_{n}. Consequently, most of the beam length is available for sensing and transport between the different detection stages; see Fig. 1. Minimizing the separation between detection stages reduces the dead time of each measurement. This would also suppress the lost Rydberg population fraction from stochastic decays 1e3μs/τn1-e^{-3\,{\rm\mu s}/\tau_{n}} to the percent level or lower. This effect is negligible for our efficiency estimates. Also note that DM-induced transitions occur between nearby nn states, whereas radiative decays in low-ll states predominantly drive electrons to the lowest-nn levels. Therefore, we do not expect these decay processes to be a significant background for our nn+in\rightarrow n+i transition–based search.

As shown in Fig. 2, the DM mass range 0.1-1 meV can be covered using transitions with n(30,65)n\in(30,65). The largest gap between adjacent transitions is 0.02meV=2π(4.8GHz)0.02\,{\rm meV}=2\pi(4.8\,{\rm GHz}). To bridge this gap, the transition frequencies can be shifted using external Zeeman or Stark effects. The Zeeman (magnetic) shift is, to leading order, independent of nn: 2πMHz(Bext/10G)\sim 2\pi\,{\rm MHz}(B_{\rm ext}/10\,{\rm G}). In contrast, the Stark shift depends on nn quadratically: 2π(4GHz)(Eext/V/cm)(n/40)2\sim 2\pi(4\,{\rm GHz})(E_{\rm ext}/{\rm V/cm})(n/40)^{2}. Therefore, modest magnetic fields (particularly at the lower end of this mass range) and electric fields at the V/cm level (at the upper end) are sufficient to fully cover the target mass range.

Refer to caption
Figure 3: Projected Rydberg SPD sensitivities in the BREAD experiment for axion (left) and dark photon (right) dark matter. We assume SNR=5, DCR=0.01 Hz, 1000 days of measurement to cover the 0.1–1 meV range, and an additional 1000 days to probe the 1–10 meV range. Orange lines correspond to the BREAD dish antenna setup, and green lines include additional dielectric layers that further enhance the photon production rate and also broaden the photon linewidth. Details are discussed in the main text. Existing constraints are shown in light gray (astrophysical and cosmological) and light purple (laboratory-based). Existing limits are adapted from Ref. O’Hare (2020) and references therein. Dark purple constraints come from a pilot BREAD experiment, dubbed GigaBREAD, which used a custom coaxial horn antenna at the focal point with a low-noise radio-frequency receiver to probe dark photons Knirck et al. (2024) and axions Hoshino et al. (2025) in the 44–52 μeV{\rm\mu eV} range (10.7–12.5 GHz).

Projected sensitivity.—We compute the signal-to-noise ratio (SNR) using

SNR=ηRDMηRDM+DCRtexp,{\rm SNR}=\frac{\eta R_{\rm DM}}{\sqrt{\eta R_{\rm DM}+{\rm DCR}}}\sqrt{t_{\rm exp}}\,,

where η\eta is the efficiency associated with each of the relevant search modalities discussed above; RDMR_{\rm DM} is the photon production rate from the BREAD setup or the BREAD with additional dielectric layers; DCR0.01Hz{\rm DCR}\sim 0.01\,{\rm Hz} is the dark count rate; and texpt_{\rm exp} is the averaging time per experiment, defined by a unique set of experimental parameters and configurations.

We assume a total experimental runtime of 1000 days per decade in mass and adopt an SNR=5 criterion for projected sensitivities, following the conventions of the original BREAD proposal Liu et al. (2022). The results are presented in Fig. 3. We discuss some features of the projected sensitivities in the following.

0.1–1 meV with BREAD: The search is based on Rydberg–Rydberg transitions in a resonant scheme, with a bin width of ΔmDM/mDM=106\Delta m_{\rm DM}/m_{\rm DM}=10^{-6}. As shown in Fig. 2, covering the full mass range requires approximately 4×1044\times 10^{4} scanning steps when using Rydberg states with n=31,,65n=31,\cdots,65. Total number of experiments is therefore 35×4×10435\times 4\times 10^{4}, which results in an average measurement time of about texp=60t_{\rm exp}=60 seconds per experiment. The reported sensitivity lies precisely at the boundary between the background-free and DCR-limited regimes; that is, this sensitivity is achieved in the regime where it is 1/texp\propto 1/\sqrt{t_{\rm exp}}. However, for even longer averaging time the sensitivity scales as 1/texp1/41/t_{\rm exp}^{1/4}. The efficiency of detecting BREAD photons is η=1\eta=1 for this search.

0.1–1 meV with ‘BREAD+ dielectric layers’: This search is also based on Rydberg–Rydberg transitions, but the photon linewidth is broadened by adding dielectric layers, resulting in bin width of ΔmDM/mDM=1/17\Delta m_{\rm DM}/m_{\rm DM}=1/17. The entire mass range can be covered with 35 experiments using Rydberg states with n=31,,65n=31,\dots,65, eliminating the need for scanning. The measurement time per experiment is therefore approximately 28 days. In this mass range, the BREAD-produced photon rate is enhanced by a factor of about 300–3000. The detection efficiency for these photons decreases from 1 at the lower end of the mass range to 0.06 at the upper end. The reported sensitivity for this range lies well within the DCR-limited regime, where the sensitivity scales as 1/texp1/41/t_{\rm exp}^{1/4} for a longer averaging time.

1–10 meV with BREAD: This search is based on direct ionization of Rydberg atoms. As shown in Fig. 2, this mass range can be covered with four experiments using Rydberg states with n=55,75,100,and 125n=55,75,100,\,{\rm and}\,125. The measurement time per experiment is 250 days. The detection efficiency for these photons decreases from 10410^{-4} at the lower end of the mass range to 10510^{-5} at the upper end. The reported sensitivity is DCR-limited, i.e., the sensitivity scales as 1/texp1/41/t_{\rm exp}^{1/4} for a longer averaging time.

1–10 meV with ‘BREAD+ dielectric layers’: This search is also based on direct ionization of Rydberg atoms and can similarly be covered with four experiments of texp=250dayst_{\rm exp}=250\,{\rm days}. The photon production rate is enhanced by a factor of about 3×1033×1043\times 10^{3}–3\times 10^{4}. The detection efficiency is the same as the previous case without dielectric layers. Note that this differs from the transition-based search, where adding dielectric layers reduces efficiency because the photon linewidth broadens and its frequency shifts off resonance. The reported sensitivity is DCR-limited, i.e., the sensitivity scales as 1/texp1/41/t_{\rm exp}^{1/4} for a longer averaging time. The improvement in sensitivity is not as strong as in the lower-mass, transition-based search. This can be understood as follows. In the transition-based case, the improvement comes from two effects: a broader sensitivity bandwidth, which increases the effective integration time per experiment, and an enhancement in photon production. In contrast, the ionization-based search is already broadband in the BREAD setup without dielectric layers. Therefore, the improvement here comes only from the enhanced photon production.

Acknowledgments.—

We thank Masha Baryakhtar, Peter Graham, David E. Kaplan, Roni Harnik, Harikrishnan Ramani, Ron Walsworth, and Samuel Wong for useful discussions, and especially Don Fahey for a careful reading of the manuscript and for comments on various experimental aspects. AB is supported by the National Science Foundation under grant number PHY-2514660 and the Maryland Center for Fundamental Physics. R.E. is supported by the John Templeton Foundation Award No. 63595 and the University of Delaware Research Foundation NSF Grant No. PHY-2515007. The work of R.E. was also supported by the Grant 63034 from the John Templeton Foundation and the University of Maryland Quantum Technology Center. S.R. is supported in part by the U.S. National Science Foundation (NSF) under Grant No. PHY-1818899. S.R. is also supported by the DOE under a QuantISED grant for MAGIS. The work of S.R. was also supported by the Simons Investigator Award No. 827042.

References

Supplemental Material
Rydberg Single Photon Detection for Probing 0.1–10 meV Dark Matter with BREAD

Abhishek Banerjee, Reza Ebadi, and Surjeet Rajendran

I Rate calculations

In this section, we outline the calculation of a) the transition rate between two rydberg states ifi\to f and b) the ionization rate from a rydberg state ii in the presence of background electric field. As the dark matter (DM), which is coherently oscillating converted photon providing the background electric field, it is taken to be classical which oscillating in time with the same frequency as the DM as E(t)=EDMcos[mDMt+φ]\vec{E}(t)=E_{\rm DM}\cos[m_{\rm DM}t+\varphi] with an amplitude of EDME_{\rm DM} and a random phase φ\varphi. As the DM velocity |v|1031|v|\simeq 10^{-3}\ll 1, we omit the space dependent part. In the presence of a time dependent electric field, the Hamiltonian becomes time-dependent as Hint=eE(t)rH_{\rm int}=e\vec{E}(t)\cdot\vec{r} due to the dipole interaction. The problem is similar to calculate the transition due to radiation of frequency mDMm_{\rm DM}.

Transition.

To calculate the transition probability from ifi\to f state, we use Fermi’s golden rule and obtain

Pif=|eE(t)i|r|f/2|2t2(sin[(ωtmDM)t/2](ωtmDM)t/2)2,\displaystyle P_{i\to f}=\left|e\vec{E}(t)\bra{i}\vec{r}\ket{f}/2\right|^{2}t^{2}\left(\frac{\sin[(\omega_{t}-m_{\rm DM})t/2]}{(\omega_{t}-m_{\rm DM})t/2}\right)^{2}, (S1)

where ωt=EfEi\omega_{t}=E_{f}-E_{i} is the energy difference between two states. As, sinx/x1\sin x/x\to 1 for x1x\lesssim 1 and |sinx/x|1/x2|\sin x/x|\lesssim 1/x^{2} for x1x\gtrsim 1, for t2/(ωtmDM)t\lesssim 2/(\omega_{t}-m_{\rm DM}), we obtain a resonant enhancement. Thus for t2min[τcoh,τRyd]t\lesssim 2\,{\rm min}[\tau_{\rm coh},\tau_{\rm Ryd}], the transition probability can be written as

Pif|eEDMi|r|f/2|2(2min[τcoh,τRyd])2,\displaystyle P_{i\to f}\approx\left|eE_{\rm DM}\bra{i}\vec{r}\ket{f}/2\right|^{2}(2\,{\rm min}[\tau_{\rm coh},\tau_{\rm Ryd}])^{2}\,, (S2)

where τRyd\tau_{\rm Ryd} is the life time of the rydberg states. The transition rate, which is found by averaging the probability over the characteristic time Γif=Pif(t)τ/τ\Gamma_{i\to f}=\left<P_{i\to f}(t)\right>_{\tau}/\tau can be written as,

γnn+i2πα3|EDM|2|i|r|f|2×min[τcoh,τRyd].\displaystyle\gamma_{n\to n+i}\approx\frac{2\pi\alpha}{3}|E_{\rm DM}|^{2}\,|\left<i|r|f\right>|^{2}\times{\rm min}[\tau_{\rm coh},\tau_{\rm Ryd}]\,. (S3)

For a transition between two rydberg states nn+in\to n+i with ini\ll n, due to the large radial wave function of the rydberg states n|r|n+in2a0\left<n|r|n+i\right>\simeq n^{2}a_{0} where a0=(meα)1a_{0}=(m_{e}\alpha)^{-1} is the Bohr radius, and nn is the principle quantum number Gallagher (1994). So the rate calculation simplifies to

γif2πα3|EDM|2n4a02×min[τcoh,τRyd].\displaystyle\gamma_{i\to f}\approx\frac{2\pi\alpha}{3}|E_{\rm DM}|^{2}\,n^{4}a_{0}^{2}\,\times{\rm min}[\tau_{\rm coh},\tau_{\rm Ryd}]\,. (S4)

In our set up, we are considering a beam of rydberg atom passing through the focusing region of the BREAD experiment which is of the size of Afocus=π(mm)2A_{\rm focus}=\pi(\rm mm)^{2}. For a beam of velocity vbeamv_{\rm beam}, traversing the focusing area takes τsens.=Afocus/π/vbeam\tau_{\rm sens.}=\sqrt{A_{\rm focus}/\pi}/v_{\rm beam} amount of time. And if τsens.min[τcoh,τRyd]\tau_{\rm sens.}\lesssim{\rm min}[\tau_{\rm coh},\tau_{\rm Ryd}], then each rydberg atom only gets τsens.\tau_{\rm sens.} amount of time to interact with the photon. Taking this into account, we find

γif2πα3|EDM|2n4a02×min[τcoh,τRyd,τsens.].\displaystyle\gamma_{i\to f}\approx\frac{2\pi\alpha}{3}|E_{\rm DM}|^{2}\,n^{4}a_{0}^{2}\,\times{\rm min}[\tau_{\rm coh},\tau_{\rm Ryd},\tau_{\rm sens.}]\,. (S5)

The total photon absorption rate is given by the single-atom rate multiplied by the number of Rydberg atoms Γnn+i=NRydγnn+i\Gamma_{n\to n+i}=N_{\rm Ryd}\gamma_{n\to n+i}. For a beam of flux Φ\Phi, the total number of Rydberg atoms present at any instant in the sensing volume is NRydΦbeamAfocusτsens.N_{\rm Ryd}\simeq\Phi_{\rm beam}A_{\rm focus}\tau_{\rm sens.}.

One key advantage of a setup such as BREAD is that due to its geometry, the DM induced photon is focused in the focusing area. Thus the effective electric field EDME_{\rm DM} which enters in the rate calculation is the enhanced and can be obtained from the rate calculation of BREAD as EDM2=2RDMmDM/AfocusE_{\rm DM}^{2}=2R_{\rm DM}m_{\rm DM}/A_{\rm focus}. By plugging in everything, we find the efficiency of photon absorption η=Γnn+i/RDM\eta=\Gamma_{n\to n+i}/R_{\rm DM} as

η=4πα3Φbeamτsens.mDM(n2a0)2min[τcoh,τRyd,τsens.].\displaystyle\eta=\frac{4\pi\alpha}{3}\Phi_{\rm beam}\tau_{\rm sens.}m_{\rm DM}(n^{2}a_{0})^{2}{\rm min}[\tau_{\rm coh},\tau_{\rm Ryd},\tau_{\rm sens.}]. (S6)

An efficiency η1\eta\gtrsim 1 means the rydberg atoms efficiently absorb all the available photon produced in the experiment.

Ionization.

Now we want to calculate the rate for the ionization process due to the background electric field. Compare to the previous case, here we need to calculate the interaction matrix element between one bound state nn with energy EnE_{n} and one state in the continuum |ψcont.\ket{\psi_{\rm cont.}}. Since a free state does not have a definite energy, we consider a transition to a free state with energy in the range (EfΔE/2,Ef+ΔE/2)(E_{f}-\Delta E/2,\,E_{f}+\Delta E/2). For the continuum states, we need to calculate the density of states given an energy EE which we denote as ρ(E)\rho(E).

Again using the Fermi’s golden rule, we find the transition probability as

Pion=EfΔE/2Ef+ΔE/2\displaystyle P_{\rm ion}=\int_{E_{f}-\Delta E/2}^{E_{f}+\Delta E/2} dEρ(E)|eE(t)i|r|ψf|2(t/2)2×\displaystyle dE\,\rho(E)\,\left|e\vec{E}(t)\bra{i}\vec{r}\ket{\psi_{f}}\right|^{2}(t/2)^{2}\times (S7)
(sin[(EEnmDM)t/2](EEnmDM)t/2)2.\displaystyle\left(\frac{\sin[(E-E_{n}-m_{\rm DM})t/2]}{(E-E_{n}-m_{\rm DM})t/2}\right)^{2}\,.

The sinc function is peaked for E=En+mDM=EfE=E_{n}+m_{\rm DM}=E_{f}, and we get

Pionρ(Ef)|eE(t)n|r|ψcont|2t2ΔEt/4ΔEt/4𝑑x(sinxx)2,\displaystyle P_{\rm ion}\approx\rho(E_{f})\,\left|e\vec{E}(t)\bra{n}\vec{r}\ket{\psi_{\rm cont}}\right|^{2}\frac{t}{2}\int_{-\Delta Et/4}^{\Delta Et/4}dx\,\left(\frac{\sin x}{x}\right)^{2}\!,

where we define x=(EEf)t/2=(EEnmDM)t/2x=(E-E_{f})t/2=(E-E_{n}-m_{\rm DM})t/2. The integral evaluates to π\pi in the large time limit and like before, by averaging the probability over the characteristic time, we get the ionization rate as,

γncont=π2×ρ(Ef)×e2|E(t)|2|nm|r|ψcont|2,\displaystyle\gamma_{n\to{\rm cont}}=\frac{\pi}{2}\times\rho(E_{f})\times e^{2}|\vec{E}(t)|^{2}\left|\bra{n\ell m}\vec{r}\ket{\psi_{\rm cont}}\right|^{2}\,, (S8)

where we take the bound state has |i=|nlm\ket{i}=\ket{nlm} quantum numbers respectively.

Calculating both ρ(Ef)\rho(E_{f}) and nm|r|ψcont\bra{n\ell m}\vec{r}\ket{\psi_{\rm cont}} involves the normalization of the continuum states. Away from the ionization threshold InI_{n}, i.e. for mDM|In|m_{\rm DM}\gg|I_{n}|, one can approximate the free states as plane waves. In that case the absorbed photon energy mostly get converted to the momentum of the free electron, k\vec{k}, as mDM=|In|+|k|2/(2me)m_{\rm DM}=|I_{n}|+|\vec{k}|^{2}/(2m_{e}) from the energy conservation. Using ψcont(r)=1/L3exp(ikr)\psi_{\rm cont}(\vec{r})=1/\sqrt{L^{3}}\exp(-i\vec{k}\cdot\vec{r}) and ρ(Ef)=(L/2π)3mekdΩ\rho(E_{f})=(L/2\pi)^{3}m_{e}k\,d\Omega (density of states in a given solid angle), we get

γncont=π2×mekdΩ8π3×e2|E(t)|2×cos2θ|0𝑑rr302π𝑑ϕ11d(cosθ)cosθei|k||r|cosθψnm(r,θ,ϕ)|2,\displaystyle\gamma_{n\to{\rm cont}}=\frac{\pi}{2}\times\frac{m_{e}k\,d\Omega}{8\pi^{3}}\times e^{2}|\vec{E}(t)|^{2}\times\cos^{2}\theta\left|\int_{0}^{\infty}dr\,r^{3}\int_{0}^{2\pi}d\phi^{\prime}\int_{-1}^{1}d(\cos\theta^{\prime})\cos\theta^{\prime}e^{-i|\vec{k}||\vec{r}|\cos\theta^{\prime}}\psi_{n\ell m}(r,\theta^{\prime},\phi^{\prime})\right|^{2}\,, (S9)

where cosθ\cos\theta is the effective angle between the electric field and k\vec{k}, and LL is the size of the box that is introduced to regulate the integral. Note that, LL drops out of the equation and we can safely take the LL\to\infty limit to represent the continuum states.

We can further simplify the above expression if we take the nsns state of the rydberg atom as the initial state. In that case we have

ψn00(r)=1πn5a03er/(na0)Ln11[2r/(na0)],\displaystyle\psi_{n00}(r)=\frac{1}{\sqrt{\pi n^{5}a_{0}^{3}}}\,e^{-r/(na_{0})}L^{1}_{n-1}[2r/(na_{0})]\,, (S10)

where, Ln11[2r/(na0)]L^{1}_{n-1}[2r/(na_{0})] is a generalized Laguerre polynomial of degree n1n-1. Using this simplification we obtain

11d(cosθ)cosθei|k||r|cosθ=2ik2r2×[krcos(kr)+sin(kr)],\displaystyle\int_{-1}^{1}d(\cos\theta^{\prime})\cos\theta^{\prime}e^{-i|\vec{k}||\vec{r}|\cos\theta^{\prime}}=\frac{2i}{k^{2}r^{2}}\times\left[-kr\cos(kr)+\sin(kr)\right]\,, (S11)

and plugging the above expression we further get,

γncont=mee2|E(t)|2πn5a03×cos2θdΩ×1k3×|0𝑑rr[krcos(kr)+sin(kr)]er/(na0)Ln11[2r/(na0)]|2.\displaystyle\gamma_{n\to{\rm cont}}=\frac{m_{e}e^{2}|E(t)|^{2}}{\pi n^{5}a_{0}^{3}}\times\cos^{2}\theta d\Omega\times\frac{1}{k^{3}}\times\left|\int_{0}^{\infty}dr\,r\left[-kr\cos(kr)+\sin(kr)\right]e^{-r/(na_{0})}L^{1}_{n-1}[2r/(na_{0})]\right|^{2}\,. (S12)

By using the fact that cos2θdΩ=4π/3\cos^{2}\theta d\Omega=4\pi/3, and definign r~=r/(na0)\tilde{r}=r/(na_{0}) and k~=kna0\tilde{k}=kna_{0}, we can further simplify the above expression as

γncont=16π|E(t)|2a03×(na0)2(kna0)3×|0𝑑r~r~[k~r~cos(k~r~)+sin(k~r~)]er~Ln11(2r~)|2.\displaystyle\gamma_{n\to{\rm cont}}=\frac{16\pi|E(t)|^{2}\,a_{0}}{3}\times\frac{(na_{0})^{2}}{(kna_{0})^{3}}\times\left|\int_{0}^{\infty}d\tilde{r}\,\tilde{r}\left[-\tilde{k}\tilde{r}\cos(\tilde{k}\tilde{r})+\sin(\tilde{k}\tilde{r})\right]e^{-\tilde{r}}L^{1}_{n-1}(2\tilde{r})\right|^{2}\,. (S13)

Note that the integral is given in terms of dimensionless quantities. To evaluating the above integral for a generic nn, we find that integral is proportional to

0𝑑r~r~[k~r~cos(k~r~)+sin(k~r~)]er~Ln11(2r~)8nk~3(1+k~2)2+n=8n(ka0n)3(1+k2a02n2)2+n,\displaystyle\int_{0}^{\infty}d\tilde{r}\,\tilde{r}\left[-\tilde{k}\tilde{r}\cos(\tilde{k}\tilde{r})+\sin(\tilde{k}\tilde{r})\right]e^{-\tilde{r}}L^{1}_{n-1}(2\tilde{r})\propto\frac{8\,n\,\tilde{k}^{3}}{(1+\tilde{k}^{2})^{2+n}}=\frac{8n(ka_{0}n)^{3}}{(1+k^{2}a_{0}^{2}n^{2})^{2+n}}\,, (S14)

with the proportionality constant fixed to n(k2a02n2)n1n(k^{2}a_{0}^{2}n^{2})^{n-1} for large kk, i.e. for ka0n1ka_{0}n\gg 1 and to (1)n1n(n2+2)/3(-1)^{n-1}n(n^{2}+2)/3 for k0k\to 0. However, very close to the threshold, the electron wavefunction can not be approximated as plane-waves as one expects corrections to this coming from the columbic part of the potential. In that case, the continuum wave function of the Columb potential should be used as discussed in Ovsiannikov et al. (2011); Glukhov et al. (2010). Thus, in what follows we will only consider the case of ka0n1ka_{0}n\gg 1. By solving Eq. (S13) for some values of nn, we find the above simplification is a conservative estimate and the actual ionization cross-section is much larger than our approximated value close to the threshold.

Plugging everything together, we obtain the ionization rate as

γncont\displaystyle\gamma_{n\to{\rm cont}} =1024π3EDM2a03n4(kna0)3(1+k2a02n2)4+2n×n2(k2a02n2)2n2\displaystyle=\frac{1024\pi}{3}E_{\rm DM}^{2}a^{3}_{0}n^{4}\frac{(kna_{0})^{3}}{(1+k^{2}a_{0}^{2}n^{2})^{4+2n}}\times n^{2}(k^{2}a_{0}^{2}n^{2})^{2n-2}
1024π3EDM2a03n4(|In|mDM)4+2n(mDM|In|1)3/2×n2(mDM|In|)2n2,\displaystyle\approx\frac{1024\pi}{3}E_{\rm DM}^{2}a^{3}_{0}n^{4}\left(\frac{|I_{n}|}{m_{\rm DM}}\right)^{4+2n}\left(\frac{m_{\rm DM}}{|I_{n}|}-1\right)^{3/2}\times n^{2}\left(\frac{m_{\rm DM}}{|I_{n}|}\right)^{2n-2}\,, (S15)

where we have used k2/2me=mDM|In|k^{2}/2m_{e}=m_{\rm DM}-|I_{n}| with |In|=meα2/(2n2)|I_{n}|=m_{e}\alpha^{2}/(2n^{2}) and have approximated (k2a02n2)2n2(mDM/|In|)2n2(k^{2}a_{0}^{2}n^{2})^{2n-2}\simeq(m_{\rm DM}/|I_{n}|)^{2n-2}. As expected we find that the Ionization rate is small close to the threshold and reaches a maximum value for mDM𝒪(|In|)m_{\rm DM}\sim\mathcal{O}(|I_{n}|) before falling off as a polynomial in incident photon energy away from the threshold. By expressing the ionization rate in term of cross-section (σ)(\sigma) with γncont=(EDM2/2mDM)×σ\gamma_{n\to{\rm cont}}=(E_{\rm DM}^{2}/2m_{\rm DM})\times\sigma, we find

σ\displaystyle\sigma \displaystyle\propto a03n6mDM(|In|mDM)4+2n(mDM|In|1)3/2×(mDM|In|)2n2\displaystyle a_{0}^{3}n^{6}m_{\rm DM}\left(\frac{|I_{n}|}{m_{\rm DM}}\right)^{4+2n}\left(\frac{m_{\rm DM}}{|I_{n}|}-1\right)^{3/2}\times\left(\frac{m_{\rm DM}}{|I_{n}|}\right)^{2n-2} (S16)
\displaystyle\propto a03n6mDM(|In|mDM)6(mDM|In|1)3/2\displaystyle a_{0}^{3}n^{6}m_{\rm DM}\left(\frac{|I_{n}|}{m_{\rm DM}}\right)^{6}\left(\frac{m_{\rm DM}}{|I_{n}|}-1\right)^{3/2}
\displaystyle\propto a0n4me(|In|mDM)5(mDM|In|1)3/21n3e2me(meα2/2)5/2mDM7/2,\displaystyle\frac{a_{0}n^{4}}{m_{e}}\left(\frac{|I_{n}|}{m_{\rm DM}}\right)^{5}\left(\frac{m_{\rm DM}}{|I_{n}|}-1\right)^{3/2}\sim\frac{1}{n^{3}}\frac{e^{2}}{m_{e}}\frac{(m_{e}\alpha^{2}/2)^{5/2}}{m_{\rm DM}^{7/2}}\,,

in the large momentum limit i.e. for nka01mDM|In|nka_{0}\gg 1\Rightarrow m_{\rm DM}\gg|I_{n}|. Thus away from the threshold we obtain the result given in Bethe and Salpeter (1957). Even for =1\ell=1 states, we obtain the same result as Bethe and Salpeter (1957) in the large-momentum (away from the threshold).

As discussed previously, n2(mDM/|In|)(2n2)n^{2}(m_{\rm DM}/|I_{n}|)^{(2n-2)} term in the rate calculation is valid away from threshold. It turns out that extending this estimate to the threshold region overestimates the rate Ovsiannikov et al. (2011). We conservatively drop n2n^{2} and use γncont.(1024π/3)EDM2n4a03(In/mDM)6(mDM/In1)3/2\gamma_{n\rightarrow{\rm cont.}}\simeq(1024\pi/3)E_{\rm DM}^{2}n^{4}a^{3}_{0}\,(I_{n}/m_{\rm DM})^{6}\left(m_{\rm DM}/I_{n}-1\right)^{3/2} in the main text to obtain the experimental reach. Note that close to the threshold, the rate obtained from our estimate is consistent with the numerical results as well as the nn-scalings in Refs. Ovsiannikov et al. (2011); Glukhov et al. (2010). We leave a detailed computation of near-threshold ionization rate for future work. Close to the threshold, the plane wave approximation of the electron wave-function breaks down as the corrections due to the columbic potential start to become important. In this case one should use the continuum (positive-energy) Coulomb wavefunctions for the final electron and evaluate the dipole matrix element between the bound Rydberg state and those continuum states.

We computed the rates for =0\ell=0, m=0m=0 states. These results agree with the result of Bethe and Salpeter (1957) for =0\ell=0 states. For low-\ell states, the power-law fall-off away from threshold has only a mild dependence on the \ell Bethe and Salpeter (1957); Ovsiannikov et al. (2011). Also, the threshold ionization rate and its nn scaling does not depend on \ell at the leading order Ovsiannikov et al. (2011). Thus, we justify the use of the conservative order-of-magnitude estimate mentioned above.