Low-mass constraints on WIMP effective models of inelastic scattering using the Migdal effect
Abstract
We use the Migdal effect to extend to low masses the bounds on each of the effective couplings of the non-relativistic effective field theory of a WIMP of mass and spin 1/2 that interacts inelastically with nuclei by either upscattering to a heavier state with mass splitting or by downscattering to a lighter state with . In order to do so we perform a systematic analysis of the Migdal bounds in the parameter space comparing them to those from nuclear recoil searches. The Migdal effect allows to significantly extend to low WIMP masses the nuclear recoil bounds for . In this case the bounds are driven by XENON1T, except when is vanishing or very small, when, depending on the WIMP–nucleus interaction, in the lower end of the range either DS50 or SuperCDMS are more constraining. On the other hand, when and the WIMP particle upscatters to a heavier state nuclear recoil bounds are stronger than those from the Migdal effect.
CQUeST–2024–0738
1 Introduction
The absence of a signal from Weakly Interacting Massive Particles (WIMPs), the most popular dark matter (DM) candidates, has prompted the scientific community to optimize direct detection experiments to the search of sub-GeV DM [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17]. One of the main challenges in detecting low-mass DM is the small energy deposited in the detector, which often lies below the detection thresholds. This makes it difficult to search for such DM particles, particularly through nuclear recoils. To overcome this challenge new analysis techniques such as the Migdal effect [18] have been put forward. The Migdal effect allows to detect low-mass DM through secondary atomic ionization, and occurs when the WIMP–nucleus scattering process triggers the emission of an electron from the target [15]. It has been first pointed out in [8, 9, 19, 20] that if the DM particle couples comparably to both protons and electrons, the Migdal ionization rate can dominate over electron scattering for certain DM masses and mediator types, particularly in the hundreds of MeV range [21, 22]. Even lower masses can be probed in Migdal–type scenarios where the DM particle is a scalar that is absorbed by a nucleon or an electron triggering atomic ionization [23, 24].
In direct detection experiments it is often easier to detect electromagnetic (EM) energy directly deposited in the measuring device compared to nuclear recoils, although it is more difficult to distinguish it from that generated from the background contamination. However, for very light WIMP masses the nuclear recoil process is kinematically not accessible, and the Migdal effect is the only way to detect a signal.
Recent studies [25, 26] have combined the Migdal effect with the Inelastic Dark Matter (IDM) scenario [27], where a WIMP state with mass interacts with nuclear targets by scattering to a different state with mass . The case , indicated in the literature as exothermic DM [28], implies the de-excitation of a meta-stable DM state that releases additional energy in the detector, allowing to extend further the sensitivity to very light WIMPs.
In Refs. [25] and [26] the Migdal effect for IDM was analyzed considering the cases of a spin-independent (SI) and spin-dependent (SD) WIMP–nucleus interaction, respectively. In our paper we wish to extend those analyses to a wider class of processes, by making use of the non-relativistic effective field theory (NREFT) framework [29, 30], that describes the most general WIMP–nucleus interactions for a particle of spin 1/2 by introducing a total of 14 momentum and velocity dependent operators (see Table 1). Indeed, models that imply non–standard interactions for inelastic Dark Matter exist in the literature. For instance, in its non–relativistic limit Magnetic inelastic dark matter [31, 32, 33, 34] leads to the emergence of a combination of the , , and operators, and in some kinematic conditions the operator has been shown to dominate over and to drive the scattering process (albeit this has been checked quantitatively only for the kinematics of elastic scattering [35]). So, on general grounds, one cannot rule out the possibility that, if inelastic DM exists, it interacts with nuclei in ways that are not standard.
Since previous studies discuss only specific benchmark values of the IDM parameters we conduct a thorough scan in terms of and , showing that in absence of a signal the Migdal effect allows to exclude previously unexplored regions of the NREFT parameter space. In our analysis we employ the data from the XENON1T [36], DS50 [37], and SuperCDMS [38] experiments, revealing some degree of complementarity among them. The interplay between the Migdal effect and elastic scattering has been previously explored within the NREFT framework in [10, 39].
2 Inelastic Scattering of Dark Matter under the Migdal Effect
In this work we consider the inelastic scattering process of a DM particle off a nuclear target . Due to energy conservation,
(2.1) |
where is the DM–nucleus reduced mass, is the DM particle incoming speed relative to the nucleus, that is assumed to be at rest in the lab rest frame. In the expression above and represent the energy of the outgoing DM particle and nucleus, respectively. Moreover, is the amount of the initial kinetic energy of the DM particle lost due to the two simultaneous inelastic effects of the transition and the target ionization with the emission of one electron. In particular, it is expressed as , where accounts for the electromagnetic energy deposited in the detector by the ionization process, and represents the mass splitting between the two DM states i.e. ( corresponds to exothermic scattering and to endothermic scattering, respectively). It is clear from Eq. (2.1) that the maximum value of is equal to the initial available kinetic energy of the incoming WIMP,
(2.2) |
The nuclear recoil energy in the frame of the detector is expressed as,
(2.3) |
where is the DM–nucleon scattering angle in the center-of-mass frame. In Eq. (2.3), we made the approximation .
In our analysis we focus on the electron ionization rate in Xe, Ar, and Ge targets, and make use of the ionization probabilities from [15], obtained under the single-atom approximation 111Recently, the calculation of the ionization probabilities of [15] was improved using the Dirac-Hartree-Fock method [40]. In the ionization energy ranges relevant to our analysis both calculations lead to similar results.. In particular, the ionisation event rate in a direct detection experiment due to the Migdal effect is given by [15],
(2.4) |
with
(2.5) |
In our analysis we include the electrons up to the shells = 3 (SuperCDMS and DS50), 4, 5 (XENON1T). It is worth noticing that in the lab frame the ionization probabilities are boosted due to the recoil of the nucleus, and scale linearly with [15]. The Migdal effect is only relevant at low WIMP masses and becomes subdominant as soon as nuclear recoils are kinematically allowed. In the expression above represents the total deposited energy and is given by,
(2.6) |
where is the quenching factor, represents the energy of the outgoing electron and the atomic de-excitation energy. In particular for the electromagnetic energy produced by the recoil of the nucleus is always negligible compared to that produced by the Migdal process, leading to . The minimum and the maximum of the recoil energy are obtained from Eq. (2.3), using and respectively. Finally, the minimum DM velocity required for the recoil of the nucleus at a given energy is given by,
(2.7) |
In our analysis, we adopt a standard Maxwell–Boltzmann velocity distribution at rest in the Galactic frame with escape velocity km/s and boosted to the solar rest frame with rotation speed of km/s.
As a result of the small momentum transfer the WIMP–nucleus scattering process is non relativistic and can be described in terms of the most general NREFT allowed by symmetry under Galilean boosts [29, 30]. In this context, the expression for the interaction Hamiltonian between the target nucleus, and the WIMP particle is given by,
(2.8) |
with and the identity and third Pauli matrix in isospin space, respectively. The isoscalar and isovector (dimension -2) coupling constants and , are related to those to protons and neutrons and by and . For a spin-1/2 WIMP there are 14 operators [29, 30, 41], which depend at most linearly on the transverse velocity (defined as the component of perpendicular to the transferred momentum ), and that are listed in Table 1. In the case of the standard SI and SD interactions (corresponding to the and operators, respectively, in the convention of [29, 30]) it is common to express the couplings and (with ) in terms of the corresponding WIMP–nucleon scattering cross-sections,
(2.9) |
where is the WIMP–nucleon reduced mass.
The detailed expression for the calculation of the differential rate in Eq. (2.5) is provided in Section 2 of [42], which has been implemented in the WimPyDD code [43]. Such expression scales linearly with the DM density in the neighborhood of the Sun, for which we adopt the standard value = 0.3 GeV/cm3. Moreover, we parameterize with the fraction of provided by the particle (notice that DD experiments are only sensitive to the combination ).
In the non-relativistic limit the WIMP–nucleus differential cross section is proportional to the squared amplitude,
(2.10) |
with the nuclear mass, , the spins of the target nucleus and of the WIMP, where , and [30],
(2.11) |
In the expression above the squared amplitude is summed over initial and final spins and the ’s are WIMP response functions provided, for instance, in Eq. (38) of [30]. They depend on the couplings as well as the transferred momentum, while,
(2.12) |
and is given by Eq. (2.7). Moreover, in Eq. (2.11) the ’s are nuclear response functions and the index represents different effective nuclear operators, which, under the assumption that the nuclear ground state is an approximate eigenstate of and , can be at most eight: following the notation in [29, 30], =, , , , , , , . The ’s are function of , where is the size of the nucleus. For the target nuclei used in most direct detection experiments the functions , calculated using nuclear shell models, have been provided in Refs. [30, 44].
The ionization probabilities are calculated using the impulse approximation, that assumes that the DM–nucleus collisions happen rapidly compared to the time scale set by the potential well of the detector lattice, , with the phonon frequency [45]. In particular, the time scale of DM–nucleus collision and the emission of Migdal electrons is of the order (in natural units). As a consequence, the validity of the impulse approximation implies a lower cut on .
In the case of a xenon fluid, we utilise a time scale of s, which applies to xenon at 170K [25]. Consequently, in the energy integration of Eq. (2.4) we set meV. Using Eq. (2.3) for elastic scattering this implies GeV. However, through the parameter the lower bound on depends on the mass splitting and on the experimental threshold on . For instance, taking into account the XENON1T energy threshold (see Section 3) for keV, one gets MeV.
For an argon fluid at 90K using an inter-atomic distance of 4 Angstroms and a sound speed of 847.55 m/s [46], we estimate a time scale of s leading again to 50 meV. In the case of elastic scattering, this implies GeV. For keV, we set MeV.
As far as the germanium target is concerned, we closely follow the work of [45], where it is shown that the impulse approximation holds if the momentum transfer satisfies the condition , with estimated between 30–50 meV for a typical crystal. Taking a conservative approach, we assume meV. Using this criterion, we find GeV for elastic scattering, while for keV, we obtain MeV.
In the case of endothermic scattering () rather than from the impulse approximation cut the lower bound on is obtained from kinematics, specifically from the requirement that the argument of the square root in Eq. (2.3) is positive. For instance, for keV this leads to GeV (independent on the target as long as ).

3 Analysis
The Migdal event rate is obtained by multiplying the WIMP–nucleus scattering rate with the ionization probability (Eq. (2.5)). As shown in Eq. (2.7), the emission of an electron leads to modifications in the kinematics of the scattering rate compared to elastic scattering. In particular, Eq. (2.7) mimics that for the kinematics of IDM [27], with the replacement of the mass splitting with , defined as .
In WimPyDD [43] we implement the kinematics of the Migdal effect through the argument delta
of the wimp_dd_rate
routine. Specifically, we adopted , and integrated the scattering rate over the undetected nuclear scattering energy 50 meV, in order to comply with the impulse approximation as discussed in the previous Section. Furthermore, we used the ionization probability for Xe, Ar, and Ge atoms from [15], corresponding to the XENON1T [36], DS50 [37], and SuperCDMS [38] experiments, respectively.
In Fig. 1 we provide an explicit example of the energy dependence of the Migdal differential rate in the case of different NREFT interactions, for a Xe target and GeV. In all the plots the curves have the same normalization, to show that, especially in the energy range that drives the overall signal, the NREFT interaction has a mild effect on the energy spectrum, which is mainly determined by the nuclear ionization probabilities. The solid red, dashed blue, and dotted green lines represent keV (exothermic), keV, and keV (endothermic), respectively, for all the 14 interaction couplings of Eq. (2.8). Different peaks in Fig. 1 correspond to the contributions of the shells . It is worth mentioning that depending on the thresholds of the considered experiments, only some of the shells contribute to the Migdal event rate.


In Fig. 2 we plot the nuclear recoil spectrum for a Xe target, taking GeV. From such plot one can observe that for endothermic DM () the normalization of the spectrum decreases with while at the same time the energy interval where the spectrum is different from zero shrinks. This leads to a suppression of the Migdal signal. On the other hand for exothermic scattering () both the minimum and the maximum nuclear recoil energies are shifted to higher values for growing , while the normalization of the spectrum remains the same. This leads to an enhancement for the Migdal signal because, as previously pointed out, the ionization probabilities grow linearly with due to the boost from the recoiling atom to the lab frame. Similar conclusions hold for other NREFT interactions and targets, so that, on general grounds, compared to the elastic case the Migdal signal is enhanced for and suppressed for . It is again important to mention that the shape of the Migdal energy spectrum is determined by the ionization probabilities and remains approximately the same irrespective of the interaction operator, of and of .
The results of our analysis are presented in Figs. 3–12. In all our plots we assume one non-vanishing coupling at a time in the effective Hamiltonian of Eq. (2.8) and an isoscalar interaction, i.e. . In Figs. 3–5 the exclusion limits on each effective coupling are shown as a function of for , while in Figs. 6–8 we show the analogous results for keV. The exclusion limit on the same couplings are explored in the plane in Figs. 9–11. Additionally, the constraints on the standard SI and SD cross-sections (Eq. 2.9) in the plane are presented in Fig. 12. In particular, from Figs. 9–11, it is evident that the Migdal exclusion bounds are overcome by those from nuclear recoil for keV. For this reason in Figs. 3–8 we have only shown exclusion limits for and , omitting .
In Figs. 3–8, the Migdal exclusion limits from the XENON1T, SuperCDMS, and DS50 experiments are represented by blue-dashed, cyan-dashed, and green-dashed lines, respectively. The shaded grey region in each plot represents the area excluded by nuclear recoil obtained from the data of a variety of experiments, including XENON1T [47], PICO60 (using a C3F8 [48] and a CF3I target [49]), DS50 [50], LZ [51], CDEX [52], CDMSLite [53], COSINE [54], COUPP [55], CRESST-II [56], PANDAX-II [57], PICASSO [58], SuperCDMS [59], and XENONnT [60]. From these experiments we select the most constraining bound for each combination of and . The implementation details of the considered experiments are provided in Appendices A.1-A.2.














In Figs. 3–11 we separate the interaction operators among SI–type and SD–type. The SI–type operators (, , , and ) are enhanced for heavy targets and are driven either by the or by the nuclear response functions (see Eq. (2.11)). In particular corresponds to the standard SI coupling, proportional to the square of the nuclear mass number; on the other hand, is non vanishing for all nuclei and favors heavier elements with large nuclear shell orbitals not fully occupied. The SD interactions () vanish when the spin of the nucleus is zero. Among them , , , and , couple to and which are driven by the spin of the target (the sum corresponds to the standard spin-dependent form factor). The operator couples to the , which requires a target spin . Finally, the two operators and deserve a separate discussion, since, at variance with all other cases, for them velocity-suppressed contributions (proportional to ) can be of the same order of velocity-independent ones. As a consequence they are driven by a combination of (that is velocity-independent and couples to the nuclear spin) and (velocity suppressed). In particular, in Ar, that has no spin, vanishes and the signal is driven by , so it is of the SI–type. On the other hand, in the case of Ge and Xe both and are different from zero: for Ge drives the signal, since the contribution of is negligible due to the velocity suppression, so it can be considered of the SD–type; on the other hand in Xe the enhancement of due to the large target mass compensates for the velocity suppression, so that it can be considered of the SI–type.














From Figs. 3–5 one can see that for the Migdal effect allows to probe (exclude) MeV, while nuclear recoil reaches MeV for SI–type interactions and GeV for SD ones. For similar conclusions were drawn within the framework of a relativistic effective field theory in [39], although in such scenario only a subset of the 14 NR operators considered here contributes to scattering process at low energy. So the present study complements [39] by exploring the missing operators. The use of the Migdal effect allows to significantly extend the sensitivity of direct detection searches to lower WIMP masses also thanks to the complementarity of different experiments. In particular in Figs. 3–5 for SI–type interactions DS50 [37] determines the bounds for MeV and XENON1T [36] at higher . On the other hand, since the Ar target has no spin DS50 does not put any bound on SD–type interactions, where, instead, the bounds are driven by SuperCDMS [38] at low WIMP mass due to its low 70 eV threshold, while XENON1T [36] overcomes the SuperCDMS exclusion limit for MeV.
The exclusion limits for the exothermic scattering are presented for keV in Figs. 6–8. In this scenario XENON1T drives all the bounds which, compared to the case , become flat at low WIMP masses and saturate the lower cut on the WIMP mass discussed in Section 2 required to comply with the impulse approximation.
Figs. 9–11 show the bounds on the effective couplings in the plane. The contours show the most constraining bound among XENON1T [36], DS50 [37], and SuperCDMS [38]. As in the previous plots the shaded grey regions show the areas excluded by nuclear recoil searches. The dashed green, cyan, and blue lines correspond to the impulse approximation cut meV for Ar, Ge, and Xe respectively, showing that the latter determines the upper edges of the contour plots. In this regime the signal is suppressed and eventually vanishes, so such boundaries represents an approximation of the maximal reach of Migdal searches. Their accurate determination would require to calculate the Migdal expected rate beyond the impulse approximation.
4 Conclusion
In the present paper we have used the Migdal effect to obtain constraints on the Inelastic Dark Matter scenario at low WIMP mass within the NREFT framework. In order to do so we have included the experimental results from XENON1T, SuperCDMS, and DS50. To calculate the bounds we have assumed one non-vanishing coupling at a time in the effective Hamiltonian of Eq. (2.8) and an isoscalar interaction, i.e. .
The result of our analysis is that the use of the Migdal effect allows to significantly extend the sensitivity of direct detection searches to lower WIMP masses, especially for exothermic dark matter, . In particular, the bounds are driven by XENON1T with the exceptions of the case of a vanishing or very small mass splitting . In this case DS50 drives the bound when GeV and the interaction is of the SI–type, while SuperCDMS determines the bound for GeV and the interaction is of the SD–type. On the other hand, when the sensitivity of standard direct detection experiments that measure the recoil energy of the nucleus exceeds that of experiments searching for the Migdal effect.
















Acknowledgments
The work of S.K. and S.S. is supported by the National Research Foundation of Korea (NRF) funded by the Ministry of Education through the Center for Quantum Space Time (CQUeST) with grant number 2020R1A6A1A03047877 and by the Ministry of Science and ICT with grant number RS-2023-00241757. G.T. is forever grateful for the support and wisdom of his Babuji, whose recent passing has left a lasting impact on his life.
Appendix A Implementations of experimental bounds
A.1 Migdal effect
The main features of the experiments included in our analysis are summarized in Table 2.
Name | exposure (kg days) | energy bin (keV) | number of events |
XENON1T | 22,000 | 0.186 - 3.8 | 49 |
DS50 | 12,500 | 0.083 - 0.106 | 20 |
SuperCDMS | [18.8, 17.5] | 0.07 - 2 | [208, 193] |
A.1.1 XENON1T
A.1.2 DS50
As observed in [39], the Migdal effect analysis in DS50 [37] relies on a profile–likelihood procedure which is difficult to reproduce. However, as pointed out in Section 3, the shape of the Migdal energy spectrum measured experimentally is fixed by the ionisation probabilities and is approximately the same for all WIMP–nucleon interactions, which only determine the overall normalization. As a result we have used the normalisation of the exclusion plot in [37], valid for a standard SI interaction, to obtain the bounds for all the other interaction operators. In particular, to reproduce the exclusion Migdal limit of Fig. 3 in [37] we adopt the energy bin with 20 events, and an exposure of 12.5 tonne-day.
A.1.3 SuperCDMS
The SuperCDMS collaboration recently conducted a dedicated Migdal analysis [38]. They examined two datasets with exposures of 18.8 kg-days and 17.5 kg-days. For both datasets we consider the single energy bin with 208 and 193 DM events [61], respectively, and use the energy resolution and efficiency from [61]. In our plots we present the most constraining one between the two data sets.
A.2 Nuclear recoil
For completeness in Figs. 3–12 we have included the regions excluded by experiments searching for nuclear recoil events. We have already provided the details about their implementation in previous analyses. Specifically, for PANDAX-II see Appendix B1 of [42]; for DS50 see Appendix B2 of [42]; for PICASSO see Appendix B4 of [42]; for COUPP see Appendix B5 of [42]; for CDEX see Appendix B3 of [42]; for CRESST-II see Appendix B7 of [42]; for LZ and XENON1T see Appendix A1 of [62]; for PICO60 see Appendix A2 of [62] for and Appendix A3 of [62] for ; for SuperCDMS see Appendix B4 of [63]; for CDMSLite see Appendix B5 of [63]; for COSINE see Appendix B6 of [63]; for XENONnT see Appendix A1 of [64].
References
- [1] LUX collaboration, D. S. Akerib et al., Results of a Search for Sub-GeV Dark Matter Using 2013 LUX Data, Phys. Rev. Lett. 122 (2019) 131301, [1811.11241].
- [2] SENSEI collaboration, L. Barak et al., SENSEI: Direct-Detection Results on sub-GeV Dark Matter from a New Skipper-CCD, Phys. Rev. Lett. 125 (2020) 171802, [2004.11378].
- [3] EDELWEISS collaboration, Q. Arnaud et al., First germanium-based constraints on sub-MeV Dark Matter with the EDELWEISS experiment, Phys. Rev. Lett. 125 (2020) 141301, [2003.01046].
- [4] PandaX collaboration, D. Zhang et al., Search for Light Fermionic Dark Matter Absorption on Electrons in PandaX-4T, Phys. Rev. Lett. 129 (2022) 161804, [2206.02339].
- [5] DAMIC collaboration, A. Aguilar-Arevalo et al., Constraints on Light Dark Matter Particles Interacting with Electrons from DAMIC at SNOLAB, Phys. Rev. Lett. 123 (2019) 181802, [1907.12628].
- [6] R. Essig, J. Mardon and T. Volansky, Direct Detection of Sub-GeV Dark Matter, Phys. Rev. D 85 (2012) 076007, [1108.5383].
- [7] M. J. Dolan, F. Kahlhoefer and C. McCabe, Directly detecting sub-GeV dark matter with electrons from nuclear scattering, Phys. Rev. Lett. 121 (2018) 101801, [1711.09906].
- [8] J. D. Vergados and H. Ejiri, The role of ionization electrons in direct neutralino detection, Phys. Lett. B 606 (2005) 313–322, [hep-ph/0401151].
- [9] C. C. Moustakidis, J. D. Vergados and H. Ejiri, Direct dark matter detection by observing electrons produced in neutralino-nucleus collisions, Nucl. Phys. B 727 (2005) 406–420, [hep-ph/0507123].
- [10] N. F. Bell, J. B. Dent, J. L. Newstead, S. Sabharwal and T. J. Weiler, Migdal effect and photon bremsstrahlung in effective field theories of dark matter direct detection and coherent elastic neutrino-nucleus scattering, Phys. Rev. D 101 (2020) 015012, [1905.00046].
- [11] G. Grilli di Cortona, A. Messina and S. Piacentini, Migdal effect and photon Bremsstrahlung: improving the sensitivity to light dark matter of liquid argon experiments, JHEP 11 (2020) 034, [2006.02453].
- [12] W. Wang, K.-Y. Wu, L. Wu and B. Zhu, Direct detection of spin-dependent sub-GeV dark matter via Migdal effect, Nucl. Phys. B 983 (2022) 115907, [2112.06492].
- [13] N. F. Bell, J. B. Dent, R. F. Lang, J. L. Newstead and A. C. Ritter, Observing the Migdal effect from nuclear recoils of neutral particles with liquid xenon and argon detectors, Phys. Rev. D 105 (2022) 096015, [2112.08514].
- [14] CDEX collaboration, Z. Y. Zhang et al., Constraints on Sub-GeV Dark Matter–Electron Scattering from the CDEX-10 Experiment, Phys. Rev. Lett. 129 (2022) 221301, [2206.04128].
- [15] M. Ibe, W. Nakano, Y. Shoji and K. Suzuki, Migdal Effect in Dark Matter Direct Detection Experiments, JHEP 03 (2018) 194, [1707.07258].
- [16] K. V. Berghaus, A. Esposito, R. Essig and M. Sholapurkar, The Migdal effect in semiconductors for dark matter with masses below 100 MeV, JHEP 01 (2023) 023, [2210.06490].
- [17] D. Adams, D. Baxter, H. Day, R. Essig and Y. Kahn, Measuring the Migdal effect in semiconductors for dark matter detection, Phys. Rev. D 107 (2023) L041303, [2210.04917].
- [18] A. Migdal, Ionization of atoms accompanying - and - decay, J. Phys. USSR 4 (1941) 449.
- [19] H. Ejiri, C. C. Moustakidis and J. D. Vergados, Dark matter search by exclusive studies of X-rays following WIMPs nuclear interactions, Phys. Lett. B 639 (2006) 218–222, [hep-ph/0510042].
- [20] R. Bernabei et al., On electromagnetic contributions in WIMP quests, Int. J. Mod. Phys. A 22 (2007) 3155–3168, [0706.1421].
- [21] R. Essig, J. Pradler, M. Sholapurkar and T.-T. Yu, Relation between the Migdal Effect and Dark Matter-Electron Scattering in Isolated Atoms and Semiconductors, Phys. Rev. Lett. 124 (2020) 021801, [1908.10881].
- [22] D. Baxter, Y. Kahn and G. Krnjaic, Electron Ionization via Dark Matter-Electron Scattering and the Migdal Effect, Phys. Rev. D 101 (2020) 076014, [1908.00012].
- [23] H. B. T. Tan, A. Derevianko, V. A. Dzuba and V. V. Flambaum, Atomic Ionization by Scalar Dark Matter and Solar Scalars, Phys. Rev. Lett. 127 (2021) 081301, [2105.08296].
- [24] V. A. Dzuba, V. V. Flambaum and I. B. Samsonov, Migdal-type effect in the dark matter absorption process, Phys. Rev. D 109 (2024) 115032, [2312.10566].
- [25] N. F. Bell, J. B. Dent, B. Dutta, S. Ghosh, J. Kumar and J. L. Newstead, Low-mass inelastic dark matter direct detection via the Migdal effect, Phys. Rev. D 104 (2021) 076013, [2103.05890].
- [26] J. Li, L. Su, L. Wu and B. Zhu, Spin-dependent sub-GeV inelastic dark matter-electron scattering and Migdal effect. Part I. Velocity independent operator, JCAP 04 (2023) 020, [2210.15474].
- [27] D. Tucker-Smith and N. Weiner, Inelastic dark matter, Phys. Rev. D64 (2001) 043502, [hep-ph/0101138].
- [28] P. W. Graham, R. Harnik, S. Rajendran and P. Saraswat, Exothermic Dark Matter, Phys. Rev. D 82 (2010) 063512, [1004.0937].
- [29] A. L. Fitzpatrick, W. Haxton, E. Katz, N. Lubbers and Y. Xu, The Effective Field Theory of Dark Matter Direct Detection, JCAP 1302 (2013) 004, [1203.3542].
- [30] N. Anand, A. L. Fitzpatrick and W. C. Haxton, Weakly interacting massive particle-nucleus elastic scattering response, Phys. Rev. C89 (2014) 065501, [1308.6288].
- [31] S. Chang, N. Weiner and I. Yavin, Magnetic Inelastic Dark Matter, Phys. Rev. D 82 (2010) 125011, [1007.4200].
- [32] N. Weiner and I. Yavin, UV completions of magnetic inelastic and Rayleigh dark matter for the Fermi Line(s), Phys. Rev. D 87 (2013) 023523, [1209.1093].
- [33] E. Izaguirre, G. Krnjaic and B. Shuve, Discovering Inelastic Thermal-Relic Dark Matter at Colliders, Phys. Rev. D 93 (2016) 063523, [1508.03050].
- [34] S. Chatterjee and R. Laha, Explorations of pseudo-Dirac dark matter having keV splittings and interacting via transition electric and magnetic dipole moments, Phys. Rev. D 107 (2023) 083036, [2202.13339].
- [35] F. Bishara, J. Brod, B. Grinstein and J. Zupan, From quarks to nucleons in dark matter direct detection, JHEP 11 (2017) 059, [1707.06998].
- [36] XENON collaboration, E. Aprile et al., Search for Light Dark Matter Interactions Enhanced by the Migdal Effect or Bremsstrahlung in XENON1T, Phys. Rev. Lett. 123 (2019) 241803, [1907.12771].
- [37] DarkSide collaboration, P. Agnes et al., Search for Dark-Matter–Nucleon Interactions via Migdal Effect with DarkSide-50, Phys. Rev. Lett. 130 (2023) 101001, [2207.11967].
- [38] SuperCDMS collaboration, M. F. Albakry et al., Search for low-mass dark matter via bremsstrahlung radiation and the Migdal effect in SuperCDMS, Phys. Rev. D 107 (2023) 112013, [2302.09115].
- [39] G. Tomar, S. Kang and S. Scopel, Low-mass extension of direct detection bounds on WIMP-quark and WIMP-gluon effective interactions using the Migdal effect, Astropart. Phys. 150 (2023) 102851, [2210.00199].
- [40] P. Cox, M. J. Dolan, C. McCabe and H. M. Quiney, Precise predictions and new insights for atomic ionization from the Migdal effect, Phys. Rev. D 107 (2023) 035032, [2208.12222].
- [41] P. Gondolo, S. Kang, S. Scopel and G. Tomar, Effective theory of nuclear scattering for a WIMP of arbitrary spin, Phys. Rev. D 104 (2021) 063017, [2008.05120].
- [42] S. Kang, S. Scopel, G. Tomar and J.-H. Yoon, Present and projected sensitivities of Dark Matter direct detection experiments to effective WIMP-nucleus couplings, Astropart. Phys. 109 (2019) 50–68, [1805.06113].
- [43] I. Jeong, S. Kang, S. Scopel and G. Tomar, WimPyDD: An object–oriented Python code for the calculation of WIMP direct detection signals, Comput. Phys. Commun. 276 (2022) 108342, [2106.06207].
- [44] R. Catena and B. Schwabe, Form factors for dark matter capture by the Sun in effective theories, JCAP 1504 (2015) 042, [1501.03729].
- [45] S. Knapen, J. Kozaczuk and T. Lin, Migdal Effect in Semiconductors, Phys. Rev. Lett. 127 (2021) 081805, [2011.09496].
- [46] D. Bowman, C. Lim and R. Aziz, Velocity of sound in liquid argon at high pressures and temperatures, Canadian Journal of Chemistry 46 (1968) 1175–1180.
- [47] XENON collaboration, E. Aprile et al., Dark Matter Search Results from a One Ton-Year Exposure of XENON1T, Phys. Rev. Lett. 121 (2018) 111302, [1805.12562].
- [48] PICO collaboration, C. Amole et al., Dark Matter Search Results from the PICO-60 C3F8 Bubble Chamber, Phys. Rev. Lett. 118 (2017) 251301, [1702.07666].
- [49] PICO collaboration, C. Amole et al., Dark Matter Search Results from the PICO-60 CF3I Bubble Chamber, Submitted to: Phys. Rev. D (2015) , [1510.07754].
- [50] DarkSide collaboration, P. Agnes et al., Low-Mass Dark Matter Search with the DarkSide-50 Experiment, Phys. Rev. Lett. 121 (2018) 081307, [1802.06994].
- [51] LZ collaboration, J. Aalbers et al., First Dark Matter Search Results from the LUX-ZEPLIN (LZ) Experiment, Phys. Rev. Lett. 131 (2023) 041002, [2207.03764].
- [52] CDEX collaboration, L. T. Yang et al., Limits on light WIMPs with a 1 kg-scale germanium detector at 160 eVee physics threshold at the China Jinping Underground Laboratory, Chin. Phys. C42 (2018) 023002, [1710.06650].
- [53] SuperCDMS collaboration, R. Agnese et al., Low-mass dark matter search with CDMSlite, Phys. Rev. D97 (2018) 022002, [1707.01632].
- [54] G. Adhikari et al., An experiment to search for dark-matter interactions using sodium iodide detectors, Nature 564 (2018) 83–86, [1906.01791].
- [55] COUPP collaboration, E. Behnke et al., First Dark Matter Search Results from a 4-kg CF3I Bubble Chamber Operated in a Deep Underground Site, Phys. Rev. D86 (2012) 052001, [1204.3094].
- [56] CRESST collaboration, G. Angloher et al., Results on light dark matter particles with a low-threshold CRESST-II detector, Eur. Phys. J. C76 (2016) 25, [1509.01515].
- [57] PandaX-II collaboration, X. Cui et al., Dark Matter Results From 54-Ton-Day Exposure of PandaX-II Experiment, Phys. Rev. Lett. 119 (2017) 181302, [1708.06917].
- [58] E. Behnke et al., Final Results of the PICASSO Dark Matter Search Experiment, Astropart. Phys. 90 (2017) 85–92, [1611.01499].
- [59] SuperCDMS collaboration, R. Agnese et al., Results from the Super Cryogenic Dark Matter Search Experiment at Soudan, Phys. Rev. Lett. 120 (2018) 061802, [1708.08869].
- [60] XENON collaboration, E. Aprile et al., First Dark Matter Search with Nuclear Recoils from the XENONnT Experiment, Phys. Rev. Lett. 131 (2023) 041003, [2303.14729].
- [61] SuperCDMS collaboration, R. Agnese et al., Search for Low-Mass Dark Matter with CDMSlite Using a Profile Likelihood Fit, Phys. Rev. D 99 (2019) 062001, [1808.09098].
- [62] S. Kang, A. Kar and S. Scopel, Halo-independent bounds on inelastic dark matter, Journal of Cosmology and Astroparticle Physics 2023 (nov, 2023) 077.
- [63] S. Kang, I. Jeong and S. Scopel, Bracketing the direct detection exclusion plot for a WIMP of spin one half in non-relativistic effective theory, Astropart. Phys. 151 (2023) 102854, [2209.03646].
- [64] K. Choi, I. Jeong, S. Kang, A. Kar and S. Scopel, Sensitivity of WIMP bounds on the velocity distribution in the limit of a massless mediator, 2408.09658.