Abstract
The generation of nonclassical light states bears a paramount importance in quantum optics and is largely relying on the interaction between intense laser pulses and nonlinear media. Recently, electron beams, such as those used in ultrafast electron microscopy to retrieve information from a specimen, have been proposed as a tool to manipulate both bright and dark confined optical excitations, inducing semiclassical states of light that range from coherent to thermal mixtures. Here, we show that the ponderomotive contribution to the electron–cavity interaction, which we argue to be significant for low-energy electrons subject to strongly confined near-fields, can actually create a more general set of optical states, including coherent and squeezed states. The postinteraction electron spectrum further reveals signatures of the nontrivial role played by A 2 terms in the light–matter coupling Hamiltonian, particularly when the cavity is previously excited by either chaotic or coherent illumination. Our work introduces a disruptive approach to the creation of nontrivial quantum cavity states for quantum information and optics applications, while it suggests unexplored possibilities for electron beam shaping.
1 Introduction
The generation of different states of light is of fundamental interest in quantum optics and enables powerful applications such as the increase in sensitivity achieved in the interferometric detection of gravitational waves through the use of squeezed states with reduced uncertainty [1]. Likewise, the generation of approximate Gottesman–Kitaev–Preskill (GKB) states [2] is needed to implement fault-tolerant quantum computing on photonic setups [3]. In this context, quantum two-photon states are commonly produced by exploiting the nonlinear response of some materials to coherent laser illumination through processes such as four-wave mixing [4] and parametric down-conversion [5] assisted by bright modes in optical cavities. In a radically different approach, free electrons have been identified as a promising tool to generate quantum light states [6, 7]. In addition to the nanometer precision with which electron beams (e-beams) can be focused on an optical cavity, they are advantageous with respect to external light excitation in that they can interact with strongly confined dark modes because they act as broad sources of evanescent fields [8].
The interest in generating low-uncertainty states through e-beams dates back to the early stages in the development of free-electron lasers, when several theoretical works predicted this kind of statistics in the free-space radiation emission from a linear wiggler [9–11]. More recently, with the advent of ultrafast electron microscopy [12–14], the experimental ability of integrating intense laser sources and e-beams in a single setup allowed for their synchronized interaction at the specimen, giving rise to the so-called photon-induced near-field electron microscopy (PINEM) [15]. In this technique, the strong inelastic electron coupling to optical fields scattered by the specimen results in the absorption and emission of multiple light quanta, causing a substantial reshaping of the electron wave function as well as the state of the cavity modes targeted by the laser at the specimen.
From the point of view of the electron, such reshaping consists in the emergence of intense sidebands in the electron spectrum spaced by the laser photon energy [16]. Interestingly, upon further propagation over macroscopic distances, coherent electron components possessing different energies (and velocities) evolve into a train of attosecond pulses [17–19]. In addition, our ability to manipulate the electron density matrix can be extended by using squeezed light instead of coherent laser sources [20].
From the perspective of an optical mode in the specimen, the electron is known to simply act as a displacement operator on the coherent state produced by the external illumination [21]. In particular, the electron–cavity interaction leads to a density matrix with Poissonian diagonal elements when the cavity is initially prepared in the ground state [20]. Although more complicated states can be obtained from consecutive interactions with multiple electrons combined with projective measurements onto specific electron states [6, 22, 23], single electrons have so far been regarded as a semiclassical current acting on the specimen. This is indeed a robust assumption for energetic electrons whose velocity v is negligibly perturbed by the interaction (nonrecoil approximation), provided linear terms in the electron–light coupling Hamiltonian (i.e., ∝ v · A, where A is the vector potential) are dominant over ponderomotive ∝ A 2 contributions. This condition is, however, challenged for an electric field associated with a mode polarized normally to the electron propagation direction, or when sign cancellations render a vanishing linear coupling coefficient, as well as when low-energy electrons such as those used in low-energy electron microscopy (LEEM) [24, 25] (≲100 eV) are made to interact with cavities supporting resonances in the infrared range. In fact, when linear and quadratic A terms have similar strengths in the coupling Hamiltonian, we expect the effect of a single electron on a cavity mode to deviate from the classical-current description, leading, for instance, to the direct production of squeezed cavity states.
In this work, we study the interaction of a free electron with a single-mode cavity under conditions in which linear and ponderomotive interactions have commensurate strengths or when the latter dominates (Figure 1). We start by developing the theory describing the interaction of a fast electron with a single-mode cavity in the nonrecoil approximation including the effect of the quantized ponderomotive coupling, thus allowing us to study the interaction with poorly populated cavities and compute observables connected with the mode. This is in contrast to previous works that incorporated a classical description of the vector potential and, therefore, implicitly assumed that the cavity was prepared in a high-fluence coherent state [26–28]. Remarkably, the entire dynamics admits an analytical solution in terms of the consecutive action of a displacement operator and a squeezing operator. Then, we theoretically demonstrate that a cavity starting from its ground state is left in a general two-photon coherent state [29] after interaction with a single electron, including vacuum-, phase-, and amplitude-squeezed states. Finally, we study how the presence of the ponderomotive interaction can lead to asymmetric distributions in the electron spectrum both when the cavity is initially heated (i.e., prepared in a thermal state) or when it is irradiated by laser light (i.e., in an initial coherent state). Besides their relevance from a fundamental viewpoint, our results suggest a way to produce nonclassical states of light in a given dark mode that is accessible to electrons by coupling to the evanescent components of their accompanying electric fields.

Free-electron interaction with an optical cavity. We present a sketch of the interaction for a single-mode cavity. Both linear (∝ v · A) and quadratic (ponderomotive ∝ A 2) terms in the quantum vector potential operator A are present in the minimal-coupling light–matter interaction Hamiltonian. Switching on and off these two terms selects the creation of either coherent or squeezed cavity mode states, respectively.
2 Results and discussion
2.1 Electron–cavity quantum dynamics
2.1.1 The evolution operator
Following the methods introduced in Ref. [21], we start by considering a structure characterized by a single photonic mode of energy ℏω
0 satisfying a bosonic statistics and interacting with an incoming electron initially prepared with a wave function
where
As explained in detail in Methods, Section 4.1.2, given the initial condition
where the phases
For the sake of simplicity, we hereinafter do not explicitly indicate the dependence on the electron position vector r. The expansion coefficients
in terms of the matrix elements
where
as well as λ = arg{μ}. These quantities are directly obtained from the solution of the Ricatti differential equation [32] ∂ z R/2 = k* − kR 2 with R = ν/μ,
and
2.1.2 Phase-matched interaction
Although the solution in Eq. (4) can be applied to any field profile, here we focus on a simple configuration that allows us to easily understand the problem, even though less symmetric scenarios could yield more efficient ponderomotive couplings. Specifically, we consider the electron to move along a path of length L while interacting with an electric field of the form
[see Eq. (3) for η i ] by disregarding the small phase θ 0. Then, the Ricatti equation admits an analytical solution that directly leads to the following compact forms of the coupling coefficients (see Methods, Section 4.2):
where
![Figure 2:
Linear and quadratic coupling in phase-matched interactions. (a) Deviation of the linear coupling coefficient
|
β
̃
0
|
$\vert {\tilde{\beta }}_{0}\vert $
from the value obtained by neglecting the ponderomotive force |β
0| as a function of |σ
0| for a phase-matched interaction with real η
i
[see Eqs. (3) and (8a)]. (b) Squeezing parameter |σ
0| (left vertical axis) and squeezing factor 20|σ
0| dB (right vertical axis) as a function of η = ω
0
a/v for different values of v
2
a under the electron–cavity configuration depicted in the inset, in which the e-beam moves with velocity v along the axis of a cylindrical hole of radius a and the cavity mode consists of a polariton made of a combination of azimuthal numbers m = ±1 and phase-matched axial wave vector q
z
= ω
0/v.](/document/doi/10.1515/nanoph-2022-0481/asset/graphic/j_nanoph-2022-0481_fig_002.jpg)
Linear and quadratic coupling in phase-matched interactions. (a) Deviation of the linear coupling coefficient
A key point in this discussion refers to the normalization of the mode field
Nonetheless, the dependence of σ 0 on the inverse of the electron velocity, the inverse mode frequency, and the field confinement (i.e., ∝ 1/A) should allow us to reach sizeable ponderomotive couplings by going to slow electrons interacting with infrared fields. Indeed, for a mode of energy ℏω 0 = 15 meV distributed over an area A = 100 nm2, a 50 eV electron produces a squeezing parameter |σ 0| ∼ 0.1.
As a potentially practical scenario, we consider the electron to be focused at the center of a circular hole (R = 0) of radius a drilled in a polaritonic material (see inset in Figure 2(b)) under the aforementioned phase-matching condition. In particular, for a mode consisting of a linear combination of two degenerate modes with azimuthal numbers m = ±1 and a relative phase ψ, the linear coupling vanishes (i.e.,
2.2 Postinteraction quantum cavity state
2.2.1 Cavity prepared in the ground state
We now discuss the photonic state after interaction with the electron for the cavity mode starting in the ground state (i.e.,
By starting from Eq. (2), we first obtain the full density matrix
Interestingly, the coefficients
where H n are Hermite polynomials and we have defined
as well as the normalization constant
Equation (9a) tells us that different postinteraction cavity states can be selected by properly tuning the factors
where the phase of the PINEM interaction has been set to zero, z
T = 4πm
e
v
3
γ
3/ℏω
0 is the so-called Talbot distance [38, 41], and J
ℓ
is the Bessel function of order ℓ. We show the resulting coherence factor in Figure 3(b). The corresponding state left in the cavity after interacting with the electron (top row of Figure 3(c)) comprises off-diagonal elements that become visible only when
![Figure 3:
Postinteraction cavity-mode population. (a) Sketch of an electron wave packet undergoing a classical PINEM interaction of strength β
PINEM, and subsequently propagating in free-space for a distance d, where it develops a sequence of probability-density pulses and interacts with a cavity with coupling coefficients σ
0 and
β
̃
0
${\tilde{\beta }}_{0}$
. (b) Coherence factor
|
M
m
ω
0
/
v
|
$\vert {M}_{m{\omega }_{0}/v}\vert $
, determining the amount of coherence left in the two-photon coherent state [see Eq. (9a)] for the scenario depicted in panel (a) with β
PINEM = 4. We plot the result as a function of the normalized propagation distance d/z
T for several harmonics m. (c) Position variance
⟨
Δ
X
̂
2
⟩
$\langle {\Delta}{\hat{X}}^{2}\rangle $
for a PINEM-compressed electron [solid curves; see panel (a)] and for a perfectly coherent electron (dashed curves). We consider different real values of σ
0 with
β
̃
=
0
$\tilde{\beta }=0$
, θ
n
= 0, and λ = 0. (d) Elements
ρ
n
n
′
ph
${\rho }_{n{n}^{\prime }}^{\text{ph}}$
of the density matrix associated with the cavity state after interaction with a compressed electron [see panel (a)]. The interaction region is taken to be placed at the propagation distances highlighted by the color-coordinated vertical dashed lines in panel (b) assuming pairs of coupling parameters
|
β
̃
0
|
=
0.1
$\vert {\tilde{\beta }}_{0}\vert =0.1$
and |σ
0| = 1 (top row); or
|
β
̃
0
|
=
2
$\vert {\tilde{\beta }}_{0}\vert =2$
, |σ
0| = 0.1 (bottom row). All plots in panel (d) are calculated for ϕ = 0.](/document/doi/10.1515/nanoph-2022-0481/asset/graphic/j_nanoph-2022-0481_fig_003.jpg)
Postinteraction cavity-mode population. (a) Sketch of an electron wave packet undergoing a classical PINEM interaction of strength β
PINEM, and subsequently propagating in free-space for a distance d, where it develops a sequence of probability-density pulses and interacts with a cavity with coupling coefficients σ
0 and
We also consider the effect of having a partially coherent electron. In particular, by compressing the electron through free-space propagation following a single PINEM interaction (setting again
In the opposite extreme (k → 0), from Eqs. (6a) and (7), we obtain |σ
0| → 0, which leads to a Poissonian distribution in
For parameters
2.2.2 Cavity under laser illumination
Although the optical cavity mode may in principle be chosen such that it constrains the phase ϕ [Eq. (9d)], in practical configurations this parameter would be hard to change in a single experimental setup. A more feasible route to gain control over the phase ϕ consists in aligning the e-beam such that
In contrast to the scenario with no external pumping, the probabilities P
n
in Eq. (10) are strongly dependent on the electron coherence factor. Specifically, Eq. (10) reduces to Eq. (9a) only if the electron bears full coherence with respect to the laser [i.e., for
2.3 Electron spectrum after interaction with an excited cavity
We now study the electron spectrum resulting from the interaction with a cavity that is exposed to a synchronized light source. We model this scenario by assuming different initial cavity states (i.e., Fock-state amplitudes
The resulting electron energy-loss distribution is
We first consider a quasi-monochromatic laser pulse with central frequency ω
0 irradiating the cavity and inducing an initial coherent state with coefficients
Reassuringly, Eq. (12) reproduces the result obtained for an electron interacting with two mutually coherent fields of frequencies ω 0 and 2ω 0 [17, 46]. Then, plugging this expression into Eq. (11), we obtain the electron probabilities
where
In Figure 4(a–c), we show electron spectra calculated from Eq. (13) as a function of σ and the normalized energy loss ω/ω
0 for selected values of

Electron energy-loss spectrum. Electron distribution as a function of the normalized energy loss ω/ω
0 and the ponderomotive coupling |σ| for
Similarly, a highly populated mode can be also obtained via cavity heating at temperatures such that k
B
T/ℏω
0 ≫ 1. In this regime, the cavity is prepared in a pure mixture with a classical Boltzmann probability distribution, which is well approximated by the expression
To compare this expression with the results obtained under coherent laser illumination, we evaluate Eq. (14) and plot Figure 4(d–f). As one could infer directly from the form of Eq. (14), the effect of the thermal population is to smear the electron spectra obtained under coherent illumination. As a consequence, the intensity oscillations observed in Figure 4(a–c) with increasing |σ| are absent in Figure 4(d–f) and just replaced by a monotonic decrease with |σ|.
3 Concluding remarks
In brief, the inclusion of A
2 terms in the interaction between e-beams and optical cavity modes causes a departure from the role of electrons as semiclassical excitation sources. In the absence of such terms, free electrons can only create coherent states by interaction with bosonic optical modes [20], unless a postselection of the electron state is performed [6]. In contrast, the presence of a ponderomotive coupling component embodied in the A
2 terms gives rise to a new set of cavity states that include vacuum-, phase-, and amplitude-squeezed states, directly created by interaction with the electron without the need of any postselection. We also demonstrate that a squeezing of
In addition, our work shows that the ponderomotive terms allow us to estimate the linear coupling strength
We understand that the present work paves the way toward the generation of nonclassical states of light with nanometric precision, even in dark modes that do not couple to light, thus revealing a new tuning knob at the intersection of quantum information and e-beam technologies.
4 Methods
4.1 Theoretical description of the quantum electron–cavity interaction in the nonrecoil approximation
4.1.1 The effective Hamiltonian
We start by considering a fast electron interacting with a classical electromagnetic field described in the temporal gauge (i.e., with zero scalar potential) in the presence of a material structure. In our analysis, the initial electron wave function is assumed to be in a superposition of momenta concentrated around a value ℏ
k
0, which corresponds to a relativistic energy E
0 and velocity v = ℏ
k
0
c
2/E
0. Under these conditions, and for optical excitations in the infrared-visible range, the system dynamics can be modeled by means of the effective Schrödinger equation
where A(r, t) is the time-dependent vector potential associated with the external illumination in the presence of the cavity. The coefficient vector g = (e
2/2m
e
c
2
γ)(1, 1, γ
−2) in the ponderomotive term approximately incorporates relativistic corrections through the Lorentz factor
In this work, we are interested in a quantum description of the cavity and, thus, adopt a quantum-optics approach [34] with the vector potential treated as an operator:
where we have added the noninteracting Hamiltonian of the cavity mode ℏω
0
a
†
a and included the photonic degrees of freedom in the combined state
4.1.2 Exact analytical solution
The interaction Hamiltonian in Eq. (15b) presents a quadratic term in the mode ladder operators; therefore, generating a dynamics that substantially differs from the one addressed in previous works [16, 35] where only the linear coupling with the electromagnetic field was taken into account. Despite this difference, by following similar steps as those used in Ref. [21], we find an analytical solution for the combined wave function.
We start by inserting the ansatz
with
Importantly, Eq. (16) conserves the total number of excitations, and thus, the set of coefficients defined by ℓ + n = s for each choice of an integer s evolves separately. This property allows us to map our problem onto a quantum harmonic oscillator under nonlinear coupling. Indeed, given the Hamiltonian
which connects to Eq. (16) by means of the transformation
To find the scattering operator, we first notice that an analytical solution to the equation
where we have defined the operators
At this point, we invoke the relation
Finally, once the aforementioned transformation is applied to Eq. (18) and the t
0 → −∞ limit is taken, the coefficients
4.2 Phase-matched interaction
For an electron interacting with a structure that is translationally invariant along the e-beam propagation direction z, the electric field of the mode assumes the general functional form
4.2.1 Hole in a metallic slab
In order to estimate the squeezing coupling parameter σ
0 in a scenario of practical interest, we consider a circular hole of radius a and length L extending along the z direction and drilled in a homogeneous metallic slab of permittivity ϵ(ω), with the electron traveling in vacuum parallel to the hole axis. For ω
0
a/c ≪ 1, we can neglect retardation in the description of the mode, which for an azimuthal number m, has an associated electric potential inside the hole (R < a) of the form
As an example, we take the electron to be focused at the hole center (R = 0, approaching with an azimuthal angle φ). Since excitations of symmetries ±m are degenerate, we can choose the mode as any linear combination of these two. In particular, we take m = ±1 waves combined with a relative phase ψ, such that the mode field becomes
4.3 Proof of Eq. (9)
We start by considering the coefficient
where the prime in the summation symbol indicates that the indices are restricted by the inequalities n + ℓ − i′ + i − 2j ≥ 0, n + ℓ − i′ ≥ 0, and j + (i′ − i − ℓ)/2 ≥ 0, as well as by the condition that i′ − ℓ − i is an even number. In virtue of assumption (ii), the first two inequalities are automatically satisfied. Also, this assumption allows us to follow a procedure similar to the one described in the Methods section of Ref. [20], consisting in using the Striling formula to approximate the factorials and neglecting all the indices in front of n when they are not exponentiated. Then, the entire factor in the fourth line of Eq. (19) reduces to (n + ℓ) i′+2j−ℓ/2, and by making the substitutions m = (i′ − i − ℓ)/2 and s = j + m, we obtain
where now the prime indicates that the leftmost sum is restricted to even values of i + ℓ. Finally, by pushing the lower limits of the m and s sums to m = −∞ and s = 0, and further using the series expansion of the Bessel functions [55]
Funding source: Directorate-General XII, Science, Research, and Development
Award Identifier / Grant number: Horizon 2020 Grant 101017720 FET-Proactive EBEAM
Award Identifier / Grant number: Horizon 2020 Grant 964591-SMART-electron
Funding source: H2020 European Research Council
Award Identifier / Grant number: Advanced Grant 789104-eNANO
Funding source: Ministerio de Ciencia e Innovación
Award Identifier / Grant number: PID2020-112625GB-I00
Award Identifier / Grant number: Severo Ochoa CEX2019-000910-S
-
Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: This work has been supported in part by the European Research Council (Advanced Grant 789104-eNANO), European Commission (Horizon 2020 Grants 101017720 FET-Proactive EBEAM and 964591-SMART-electron), Spanish MICINN (PID2020-112625GB-I00 and Severo Ochoa CEX2019-000910-S), Catalan CERCA Program, and Fundaciós Cellex and Mir-Puig.
-
Conflict of interest statement: The authors declare no conflicts of interest regarding this article.
References
[1] D. F. Walls, “Squeezed states of light,” Nature, vol. 306, p. 141, 1983. https://doi.org/10.1038/306141a0.Search in Google Scholar
[2] D. Gottesman, A. Kitaev, and J. Preskill, “Encoding a qubit in an oscillator,” Phys. Rev. A, vol. 64, p. 012310, 2001. https://doi.org/10.1103/physreva.64.012310.Search in Google Scholar
[3] N. C. Menicucci, “Fault-tolerant measurement-based quantum computing with continuous-variable cluster states,” Phys. Rev. Lett., vol. 112, p. 120504, 2014. https://doi.org/10.1103/physrevlett.112.120504.Search in Google Scholar
[4] R. E. Slusher, L. W. Hollberg, B. Yurke, J. C. Mertz, and J. F. Valley, “Observation of squeezed states generated by four-wave mixing in an optical cavity,” Phys. Rev. Lett., vol. 55, p. 2409, 1985. https://doi.org/10.1103/physrevlett.55.2409.Search in Google Scholar PubMed
[5] L. A. Wu, H. J. Kimble, J. L. Hall, and H. Wu, “Generation of squeezed states by parametric down conversion,” Phys. Rev. Lett., vol. 57, p. 2520, 1986. https://doi.org/10.1103/physrevlett.57.2520.Search in Google Scholar
[6] X. M. Bendaña, A. Polman, and F. J. García de Abajo, “Single-photon generation by electron beams,” Nano Lett., vol. 11, p. 5099, 2011. https://doi.org/10.1021/nl1034732.Search in Google Scholar PubMed
[7] A. Feist, G. Huang, G. Arend, et al.., “Cavity-mediated electron-photon pairs,” arXiv, p. 2202.12821, 2022.Search in Google Scholar
[8] F. J. García de Abajo, “Optical excitations in electron microscopy,” Rev. Mod. Phys., vol. 82, p. 209, 2010. https://doi.org/10.1103/revmodphys.82.209.Search in Google Scholar
[9] W. Becker, M. O. Scully, and M. S. Zubairy, “Generation of squeezed coherent states via a free-electron laser,” Phys. Rev. Lett., vol. 48, p. 475, 1982. https://doi.org/10.1103/physrevlett.48.475.Search in Google Scholar
[10] I. Gjaja and A. Bhattacharjee, “Generation of squeezed radiation from a free-electron laser,” Phys. Rev. A, vol. 36, p. 5486, 1987. https://doi.org/10.1103/physreva.36.5486.Search in Google Scholar PubMed
[11] I. Gjaja and A. Bhattacharjee, “Effective density matrix for free-electron-laser radiation,” Phys. Rev. A, vol. 43, p. 3206, 1991. https://doi.org/10.1103/physreva.43.3206.Search in Google Scholar PubMed
[12] V. A. Lobastov, R. Srinivasan, and A. H. Zewail, “Four-dimensional ultrafast electron microscopy,” Proc. Natl. Acad. Sci., vol. 102, p. 7069, 2005. https://doi.org/10.1073/pnas.0502607102.Search in Google Scholar PubMed PubMed Central
[13] B. Barwick and A. H. Zewail, “Photonics and plasmonics in 4D ultrafast electron microscopy,” ACS Photonics, vol. 2, p. 1391, 2015. https://doi.org/10.1021/acsphotonics.5b00427.Search in Google Scholar
[14] S. A. Aseyev, E. A. Ryabov, B. N. Mironov, and A. A. Ischenko, “The development of ultrafast electron microscopy,” Crystals, vol. 10, p. 452, 2020. https://doi.org/10.3390/cryst10060452.Search in Google Scholar
[15] B. Barwick, D. J. Flannigan, and A. H. Zewail, “Photon-induced near-field electron microscopy,” Nature, vol. 462, p. 902, 2009. https://doi.org/10.1038/nature08662.Search in Google Scholar PubMed
[16] A. Feist, K. E. Echternkamp, J. Schauss, S. V. Yalunin, S. Schäfer, and C. Ropers, “Quantum coherent optical phase modulation in an ultrafast transmission electron microscope,” Nature, vol. 521, p. 200, 2015. https://doi.org/10.1038/nature14463.Search in Google Scholar PubMed
[17] K. E. Priebe, C. Rathje, S. V. Yalunin, et al.., “Attosecond electron pulse trains and quantum state reconstruction in ultrafast transmission electron microscopy,” Nat. Photon., vol. 11, p. 793, 2017. https://doi.org/10.1038/s41566-017-0045-8.Search in Google Scholar
[18] Y. Morimoto and P. Baum, “Diffraction and microscopy with attosecond electron pulse trains,” Nat. Phys., vol. 14, p. 252, 2018. https://doi.org/10.1038/s41567-017-0007-6.Search in Google Scholar
[19] P. Baum, “Quantum dynamics of attosecond electron pulse compression,” J. Appl. Phys., vol. 122, p. 223105, 2017. https://doi.org/10.1063/1.5006864.Search in Google Scholar
[20] V. Di Giulio and F. J. García de Abajo, “Free-electron shaping using quantum light,” Optica, vol. 7, p. 1820, 2020. https://doi.org/10.1364/optica.404598.Search in Google Scholar
[21] V. Di Giulio, M. Kociak, and F. J. García de Abajo, “Probing quantum optical excitations with fast electrons,” Optica, vol. 6, p. 1524, 2019. https://doi.org/10.1364/optica.6.001524.Search in Google Scholar
[22] A. B. Hayun, O. Reinhardt, J. Nemirovsky, A. Karnieli, N. Rivera, and I. Kaminer, “Shaping quantum photonic states using free electrons,” Sci. Adv., vol. 7, p. eabe4270, 2021.10.1126/sciadv.abe4270Search in Google Scholar PubMed PubMed Central
[23] R. Dahan, G. Baranes, A. Gorlach, R. Ruimy, N. Rivera, and I. Kaminer, “Creation of optical cat and GKP states using shaped free electrons,” arXiv, p. 2206.08828, 2022.10.1364/CLEO_QELS.2022.FTu5A.4Search in Google Scholar
[24] M. Rocca, “Low-energy EELS investigation of surface electronic excitations on metals,” Surf. Sci. Rep., vol. 22, p. 1, 1995. https://doi.org/10.1016/0167-5729(95)00004-6.Search in Google Scholar
[25] T. Nagao, S. Yaginuma, T. Inaoka, and T. Sakurai, “One-dimensional plasmon in an atomic-scale metal wire,” Phys. Rev. Lett., vol. 97, p. 116802, 2006. https://doi.org/10.1103/physrevlett.97.116802.Search in Google Scholar PubMed
[26] N. Talebi, “Strong interaction of slow electrons with near-field light visited from first principles,” Phys. Rev. Lett., vol. 125, p. 080401, 2020. https://doi.org/10.1103/physrevlett.125.080401.Search in Google Scholar
[27] M. Kozák, “Electron vortex beam generation via chiral light-induced inelastic ponderomotive scattering,” ACS Photonics, vol. 8, p. 431, 2021. https://doi.org/10.1021/acsphotonics.0c01650.Search in Google Scholar
[28] N. Talebi and I. Bezinová, “Exchange-mediated mutual correlations and dephasing in free-electrons and light interactions,” New J. Phys., vol. 23, p. 063066, 2021. https://doi.org/10.1088/1367-2630/ac06e7.Search in Google Scholar
[29] H. P. Yuen, “Two-photon coherent states of the radiation field,” Phys. Rev. A, vol. 13, p. 2226, 1976. https://doi.org/10.1103/physreva.13.2226.Search in Google Scholar
[30] F. J. García de Abajo, E. J. C. Dias, and V. Di Giulio, “Complete excitation of discrete quantum systems by single free electrons,” Phys. Rev. Lett., vol. 129, p. 093401, 2022. https://doi.org/10.1103/physrevlett.129.093401.Search in Google Scholar
[31] V. Di Giulio and F. J. García de Abajo, “Electron diffraction by vacuum fluctuations,” New J. Phys., vol. 22, p. 103057, 2020. https://doi.org/10.1088/1367-2630/abbddf.Search in Google Scholar
[32] W. T. Reid, Ricatti Differential Equations, New York, 111 5th Avenue, Academic Press, 1972.Search in Google Scholar
[33] R. F. Bishop and A. Vourdas, “Generalised coherent states and Bogoliubov transformations,” J. Phys. A: Math. Gen., vol. 19, p. 2525, 1986. https://doi.org/10.1088/0305-4470/19/13/018.Search in Google Scholar
[34] R. J. Glauber and M. Lewenstein, “Quantum optics of dielectric media,” Phys. Rev. A, vol. 43, p. 467, 1991. https://doi.org/10.1103/physreva.43.467.Search in Google Scholar PubMed
[35] O. Kfir, “Entanglements of electrons and cavity photons in the strong-coupling regime,” Phys. Rev. Lett., vol. 123, p. 103602, 2019. https://doi.org/10.1103/physrevlett.123.103602.Search in Google Scholar PubMed
[36] J. H. Yoon, F. Selbach, L. Schumacher, J. Jose, and S. Schlücker, “Surface plasmon coupling in dimers of gold nanoparticles: experiment and theory for ideal (spherical) and nonideal (faceted) building blocks,” ACS Photonics, vol. 6, p. 642, 2019. https://doi.org/10.1021/acsphotonics.8b01424.Search in Google Scholar
[37] Y. Huang, L. Ma, J. Li, and Z. Zhang, “Nanoparticle-on-mirror cavity modes for huge and/or tunable plasmonic field enhancement,” Nanotechnology, vol. 28, p. 105203, 2017. https://doi.org/10.1088/1361-6528/aa5b27.Search in Google Scholar PubMed
[38] V. Di Giulio, O. Kfir, C. Ropers, and F. J. García de Abajo, “Modulation of cathodoluminescence emission by interference with external light,” ACS Nano, vol. 15, p. 7290, 2021. https://doi.org/10.1021/acsnano.1c00549.Search in Google Scholar PubMed PubMed Central
[39] O. Kfir, V. Di Giulio, F. J. García de Abajo, and C. Ropers, “Optical coherence transfer mediated by free electrons,” Sci. Adv., vol. 7, p. eabf6380, 2021. https://doi.org/10.1126/sciadv.abf6380.Search in Google Scholar PubMed PubMed Central
[40] M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information (Cambridge Series on Information and the Natural Sciences), 1st ed. Cambridge, Cambridge University Press, 2004.Search in Google Scholar
[41] Z. Zhao, X. Q. Sun, and S. Fan, “Quantum entanglement and modulation enhancement of free-electron--bound-electron interaction,” Phys. Rev. Lett., vol. 126, p. 233402, 2021.10.1103/PhysRevLett.126.233402Search in Google Scholar PubMed
[42] S. V. Yalunin, A. Feist, and C. Ropers, “Tailored high-contrast attosecond electron pulses for coherent excitation and scattering,” Phys. Rev. Res., vol. 3, p. L032036, 2021. https://doi.org/10.1103/physrevresearch.3.l032036.Search in Google Scholar
[43] I. T. J. E. Bourassa, N. C. Menicucci, and K. K. Sabapathy, “Progress towards practical qubit computation using approximate Gottesman-Kitaev-Preskill codes,” Phys. Rev. A, vol. 101, p. 032315, 2020.10.1103/PhysRevA.101.032315Search in Google Scholar
[44] X. Guo, N. Li, X. Yang, et al.., “Approaching the Purcell factor limit with whispering-gallery hyperbolic phonon polaritons in hBN nanotubes,” arXiv, p. 2112.11820, 2022.Search in Google Scholar
[45] K. E. Echternkamp, A. Feist, S. Schäfer, and C. Ropers, “Ramsey-type phase control of free-electron beams,” Nat. Phys., vol. 12, p. 1000, 2016. https://doi.org/10.1038/nphys3844.Search in Google Scholar
[46] A. Konečná, V. Di Giulio, V. Mkhitaryan, C. Ropers, and F. J. García de Abajo, “Nanoscale nonlinear spectroscopy with electron beams,” ACS Photonics, vol. 7, p. 1290, 2020. https://doi.org/10.1021/acsphotonics.0c00326.Search in Google Scholar
[47] O. Kfir, H. Lourenço-Martins, G. Storeck, et al.., “Controlling free electrons with optical whispering-gallery modes,” Nature, vol. 582, p. 46, 2020. https://doi.org/10.1038/s41586-020-2320-y.Search in Google Scholar PubMed
[48] R. Dahan, S. Nehemia, M. Shentcis, et al.., “Resonant phase-matching between a light wave and a free-electron wavefunction,” Nature Phys., vol. 16, p. 1123, 2020. https://doi.org/10.1038/s41567-020-01042-w.Search in Google Scholar
[49] N. Müller, V. Hock, H. Koch, N. Bach, C. Rathje, and S. Schäfer, “Broadband coupling of fast electrons to high-Q whispering-gallery mode resonators,” ACS Photonics, vol. 8, p. 1569, 2021. https://doi.org/10.1021/acsphotonics.1c00456.Search in Google Scholar
[50] G. Huang, N. J. Engelsen, O. Kfir, C. Ropers, and T. J. Kippenberg, “Coupling ideality of free electrons with photonic integrated waveguides,” arXiv, p. 2206.08098, 2022.Search in Google Scholar
[51] F. J. García de Abajo and A. Konečná, “Optical modulation of electron beams in free space,” Phys. Rev. Lett., vol. 126, p. 123901, 2021. https://doi.org/10.1103/physrevlett.126.123901.Search in Google Scholar PubMed
[52] J. J. Sakurai, Modern Quantum Mechanics, Boston, Addison-Wesley, 1994.Search in Google Scholar
[53] S. Chaturvedi, M. S. Sriram, and V. Srinivasan, “Berry’s phase for coherent states,” J. Phys. A: Math. Gen., vol. 20, p. L1071, 1987. https://doi.org/10.1088/0305-4470/20/16/007.Search in Google Scholar
[54] S. M. Barnett and P. M. Radmore, Methods in Theoretical Quantum Optics, Oxford, Oxford University Press, 1997.Search in Google Scholar
[55] I. S. Gradshteyn and I. M. Ryzhik, Table of Integrals, Series, and Products, London, Academic Press, 2007.Search in Google Scholar
© 2022 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Obituary
- A tribute to the memory of professor Alexander K. Popov
- Frontmatter
- Research Articles
- Novel fiber-tip micro flowmeter based on optofluidic microcavity filled with silver nanoparticles solutions
- Multifunctional on-chip directional coupler for spectral and polarimetric routing of Bloch surface wave
- Multifunctional croconaine nanoparticles for efficient optoacoustic imaging of deep tumors and photothermal therapy
- A large-size and polarization-independent two dimensional grating fabricated by scanned reactive-ion-beam etching
- Optical-cavity mode squeezing by free electrons
- Controlled optical near-field growth of individual free-standing well-oriented carbon nanotubes, application for scattering SNOM/AFM probes
- Integrated metasurfaces on silicon photonics for emission shaping and holographic projection
- High-efficiency SOI-based metalenses at telecommunication wavelengths
- 3D Dirac semimetals supported tunable terahertz BIC metamaterials
- Turning a polystyrene microsphere into a multimode light source by laser irradiation
- Hologram imaging quality improvement by ionization controlling based on the self-trapped excitons with double-pulse femtosecond laser
- Graphene plasmons-enhanced terahertz response assisted by metallic gratings
- Low-loss, geometry-invariant optical waveguides with near-zero-index materials
- Manipulating light scattering and optical confinement in vertically stacked Mie resonators
- An operator-based approach to topological photonics
- Ultrasmall SnS2 quantum dot−based photodetectors with high responsivity and detectivity
- Suppression of (0001) plane emission in GaInN/GaN multi-quantum nanowires for efficient micro-LEDs
- Super-resolved three-dimensional near-field mapping by defocused imaging and tracking of fluorescent emitters
- Quantitative and sensitive detection of alpha fetoprotein in serum by a plasmonic sensor
- Abundant dynamics of group velocity locked vector solitons from Er-doped fiber laser based on GO/PVA film
- Dual-band bound states in the continuum based on hybridization of surface lattice resonances
- To realize a variety of structural color adjustments via lossy-dielectric-based Fabry–Perot cavity structure
- Topology-optimized silicon-based dual-mode 4 × 4 electro-optic switch
- Tunable narrowband excitonic Optical Tamm states enabled by a metal-free all-organic structure
- Mode manipulation in a ring–core fiber for OAM monitoring and conversion
- Ultrafast terahertz transparency boosting in graphene meta-cavities
- Exceptional points at bound states in the continuum in photonic integrated circuits
- NV-plasmonics: modifying optical emission of an NV− center via plasmonic metal nanoparticles
- Directional dependence of the plasmonic gain and nonreciprocity in drift-current biased graphene
- Demonstration of conventional soliton, bound-state soliton, and noise-like pulse based on chromium sulfide as saturable absorber
- Errata
- Erratum to: High-Q asymmetrically cladded silicon nitride 1D photonic crystals cavities and hybrid external cavity lasers for sensing in air and liquids
- Erratum to: NIR-II light-activated two-photon squaric acid dye with type I photodynamics for antitumor therapy
Articles in the same Issue
- Obituary
- A tribute to the memory of professor Alexander K. Popov
- Frontmatter
- Research Articles
- Novel fiber-tip micro flowmeter based on optofluidic microcavity filled with silver nanoparticles solutions
- Multifunctional on-chip directional coupler for spectral and polarimetric routing of Bloch surface wave
- Multifunctional croconaine nanoparticles for efficient optoacoustic imaging of deep tumors and photothermal therapy
- A large-size and polarization-independent two dimensional grating fabricated by scanned reactive-ion-beam etching
- Optical-cavity mode squeezing by free electrons
- Controlled optical near-field growth of individual free-standing well-oriented carbon nanotubes, application for scattering SNOM/AFM probes
- Integrated metasurfaces on silicon photonics for emission shaping and holographic projection
- High-efficiency SOI-based metalenses at telecommunication wavelengths
- 3D Dirac semimetals supported tunable terahertz BIC metamaterials
- Turning a polystyrene microsphere into a multimode light source by laser irradiation
- Hologram imaging quality improvement by ionization controlling based on the self-trapped excitons with double-pulse femtosecond laser
- Graphene plasmons-enhanced terahertz response assisted by metallic gratings
- Low-loss, geometry-invariant optical waveguides with near-zero-index materials
- Manipulating light scattering and optical confinement in vertically stacked Mie resonators
- An operator-based approach to topological photonics
- Ultrasmall SnS2 quantum dot−based photodetectors with high responsivity and detectivity
- Suppression of (0001) plane emission in GaInN/GaN multi-quantum nanowires for efficient micro-LEDs
- Super-resolved three-dimensional near-field mapping by defocused imaging and tracking of fluorescent emitters
- Quantitative and sensitive detection of alpha fetoprotein in serum by a plasmonic sensor
- Abundant dynamics of group velocity locked vector solitons from Er-doped fiber laser based on GO/PVA film
- Dual-band bound states in the continuum based on hybridization of surface lattice resonances
- To realize a variety of structural color adjustments via lossy-dielectric-based Fabry–Perot cavity structure
- Topology-optimized silicon-based dual-mode 4 × 4 electro-optic switch
- Tunable narrowband excitonic Optical Tamm states enabled by a metal-free all-organic structure
- Mode manipulation in a ring–core fiber for OAM monitoring and conversion
- Ultrafast terahertz transparency boosting in graphene meta-cavities
- Exceptional points at bound states in the continuum in photonic integrated circuits
- NV-plasmonics: modifying optical emission of an NV− center via plasmonic metal nanoparticles
- Directional dependence of the plasmonic gain and nonreciprocity in drift-current biased graphene
- Demonstration of conventional soliton, bound-state soliton, and noise-like pulse based on chromium sulfide as saturable absorber
- Errata
- Erratum to: High-Q asymmetrically cladded silicon nitride 1D photonic crystals cavities and hybrid external cavity lasers for sensing in air and liquids
- Erratum to: NIR-II light-activated two-photon squaric acid dye with type I photodynamics for antitumor therapy