On-chip stackable dielectric laser accelerator

ACCELERATOR, RAY TECHNOLOGY AND APPLICATIONS

On-chip stackable dielectric laser accelerator

Bin Sun
Yang-Fan He
Ruo-Yun Luo
Tai-Yang Zhang
Qiang Zhou
Shao-Yi Wang
Jian Zheng
Zong-Qing Zhao
Nuclear Science and TechniquesVol.34, No.2Article number 23Published in print Feb 2023Available online 22 Feb 2023
7701

In this paper, we propose a novel stacked laser dielectric acceleration structure. This structure is based on the inverse Cherenkov effect and represented by a parametric design formulation. Compared to existing dielectric laser accelerators relying on the inverse Smith–Purcell effect, the proposed structure provides an extended-duration synchronous acceleration field without requiring the pulse front tilting technique. This advantage significantly reduces the required pulse duration. In addition, the easy-to-integrate layered structure facilitates cascade acceleration, and simulations have shown that low-energy electron beams can be cascaded through high gradients over extended distances. These practical advantages demonstrate the potential of this new structure for future chip accelerators.

Dielectric laser acceleratorInverse Cherenkov effectLaser-driven particle acceleration
1

Introduction

The industrial applications of modern particle accelerators have been extended beyond scientific research [1,2]. However, radio-frequency particle accelerators are typically bulky and heavy with limited practical limitations, such as the threshold for metallic structure breakdown and restrictions in the power and wavelength of microwave sources [3-6]. In recent decades, researchers have attempted to construct miniature and even benchtop particle accelerators [7-15]. Acceleration mechanisms based on laser-dielectric interactions have recently attracted increased attention, with their promising potential to achieve the development of on-chip microscopic accelerators. The inverse Smith–Purcell-effect dielectric laser accelerator (ISP-DLA) [8-15] and the inverse Cherenkov-effect dielectric laser accelerator (ICS-DLA) [16-19] are two laser-dielectric acceleration concepts that have been investigated since the 1960s, shortly after the invention of the laser.

Most dielectric laser accelerator (DLA) schemes are based on the inverse Smith–Purcell effect. In their typical configuration, a periodic dielectric structure is adopted to spatially modulate the incident laser field and synchronize the incident-free electrons. This is achieved through the spatial harmonics of the induced waves in the electron beam channel. These previous developments have several challenges [8-15], including the following: (1) microscale fabrication of accelerator structures, which requires advanced micro-processing and nano-processing techniques; (2) selection of dielectric materials, some of which have high laser damage thresholds but cannot be fabricated as microstructures and nanostructures; (3) reduced laser damage thresholds of dielectric materials after nano-processing; (4) the pulse front tilting technique, which is required to extend the action time of the laser and dielectric to achieve extended acceleration, improve laser energy utilization efficiency, and prevent unnecessary high energy loads on the structure [20]. These challenges limit the acceleration gradient, acceleration distance, energy utilization efficiency, and flexibility of laser dielectric acceleration structures while increasing experimental difficulty.

In this study, a stacked acceleration structure based on the inverse Cherenkov effect was developed to achieve cascade acceleration of nonrelativistic low-energy electrons through a multilayer dielectric. First, the proposed structure is identical to the existing inverse Cherenkov-effect accelerator and provides effectively solves the challenges encountered in the acceleration structure of the inverse Smith–Purcell effect, including (1) elimination of the requirement for laser frontier technology and (2) not requiring multiple lasers to achieve cascade acceleration, eliminating the difficulty of coupling multiple lasers. Second, this structure considers higher acceleration gradients and energy gains and has no theoretical upper limit on acceleration energy. In addition, because this stackable structure can be assembled from modular acceleration units, it provides additional configuration flexibility to satisfy specific requirements for different applications. This flexibility of the stackable design makes the current structure a prototype for future chip accelerators in universal application scenarios.

2

Theory and structural design

The proposed stacked acceleration structure is depicted in Fig. 1. A case study of two dielectric layers is presented in this paper for demonstration without loss of generality. The line-shaped collimated incident laser was perpendicular to Surface II and the beveled surface of a multilayer dielectric prism with a right-angled triangular cross-section. An evanescent wave is generated at Surface I with a velocity of vp=c/(nsinα), travelling forward along the surface in the x-direction, where α is the angle between the hypotenuse and the bottom edge, c is the velocity of light in vacuum, and n is the refractive index of the prism. This wave has a longitudinal electric field component Ex, transverse electric field component Ey, and out-of-plane magnetic field component Bz.

Fig. 1
(Color online) (a) Schematic of stacked acceleration structure. The red arrow line represents the incident laser. The refractive indices of the two dielectric layers are n1 and n2. (b) Partial enlargement of Fig. 1(a) showing refraction process occurring at interface. The angles between the bottom edge of the right-angled triangle and the two traversal directions perpendicular to two laser beams (equivalently, the electric field directions) are α and α2, respectively.
pic

The electrons move in the x-direction in the vacuum below Surface I. Electrons in phase with an accelerating electric field (equivalently, negative phase Ex<0) can then be accelerated by applying the inverse Cherenkov effect. However, the electron velocity ve must be equal to that of the evanescent wave, that is, ve=vp, to achieve continuous acceleration. On Surface I, the first criterion [17-19] for achieving continuous acceleration through the inverse Cherenkov effect is sinα=c/(ven)=1/(βn). (1)

In Eq. (1), α is the angle between the hypotenuse and the bottom edge, c and ve are the velocities of the light in vacuum and electrons, respectively, and β=v/c is the relativistic velocity of the electron. As the electron accelerates and ve increases, the electrons move faster than the wave, lose synchronization, and eventually, their acceleration end. Currently, the solution to this problem is achieved by varying angle α [18,19]. In this study, the electron velocity was matched to the wave velocity by selecting a different dielectric to vary the refractive index. First, the refracted light ray A is analyzed, as shown in Fig. 1. Each time the laser passes through the interface between two adjacent layers with different dielectrics, the light is refracted owing to a change in the refractive index. Some of the laser light is reflected, which reduces the laser energy; this reduction is undesirable for the energy utilization efficiency and accelerating electric field. Thus, it is necessary to ensure that the light passes through the interface only once so that the laser experiences only one additional reflection and refraction. A limit was proposed for the dielectric length of each layer to achieve this. Taking the second layer in Fig. 1 as an example, the second-layer thickness L2 is limited by Rays B and C. The geometric relationship is expressed by L2=L1×(tanα)2. Similarly, the recurrence equation for the thickness Li of each subsequent layer can be derived, Li+1=Li/(cosα)2,i2. By combining these two equations, the relationship between the thickness Li of each layer and the thickness L1 of the first layer is expressed by Li=L1×(tanα)2/(cosα)2(i2), i2. Second, even when the laser passes through the interface only once, as shown in Fig. 1, the angle of the laser to the bottom surface (Surface Ⅰ) changes from α to α2 after refraction at the dielectric interface. Equation (1) must still be satisfied to maintain acceleration: sinα2=1/(β2n2). When refraction occurs at the interface, the law of refraction is expressed by n1sinθ1=n2sinθ2. In addition, the trigonometric relationship is 1=(sinα2)2+(cosα2)2=[1/(β2n2)]2+(sinθ2)2=[1/(β2n2)]2+(n1sinθ1/n2)2. Solving these equations yield n2, where n2=[(n1sinθ1)2+β12]1/2. A similar derivation holds at each interface, resulting in the recurrence relationship ni=[(ni1sinθi1)2+βi2]1/2,i2. In addition, because the incidence angle of the laser is the same for each layer of the dielectric, θ1=θ2==θn=π/2α, the recurrence relationship can be simplified to ni=[(ni1sinθi1)2+βi2]1/2=[(ni1cosα)2+βi2]1/2,i2. In summary, the refractive index and thickness of each layer of the dielectric, starting from the second layer, can be derived as a function of the parameters of the first layer.  {ni=[(ni1cosα)2+βi2]1/2Li=L1×(tanα)2/(cosα)2(i2),i2 (2)

In the first row of Eq. (2), ni is the refractive index of the layer, and the label, i, follows the convention in Figs. 1 and 2. βi is the theoretical design velocity of electrons in the vacuum acceleration region on Surface I below the ith layer of the dielectric, which is also the velocity of light in the x-direction when it is in this dielectric. For this theoretical electron velocity, as the electron beam stream advances longitudinally in the acceleration channel, the electron energy and velocity increase. Electron velocity ve gradually increases when the electron moves below the surface of each dielectric layer, whereas the velocity vo of the corresponding laser pulse remains unchanged. Therefore, the phase difference gradually accumulates as the electrons move with the laser during the acceleration phase, Φ=ωt(1nβsinα). If the electron beam is injected at the maximum amplitude of the laser, it becomes necessary to ensure that the phase difference |Φ|π/2 when the electron beam leaves; otherwise, it enters the deceleration phase. This corresponds to two cases. (1) The electron velocity ve is extremely high, resulting in phase difference Φπ/2, and the electron beam reaches the front deceleration phase. (2) The electron velocity ve is extremely low, resulting in phase difference Φπ/2, and the electron beam will enter the rear deceleration phase.

Fig. 2
(Color online) Schematic of structure of 50 keV electron acceleration. The light blue, purple, and red ellipses represent the injected low-energy (50 keV) electron beam stream, accelerating electron beam stream (schematic), and output high-energy electron beam stream, respectively. The insertions “dec” and “acc,” indicate that half of the phases in a laser cycle accelerate the electrons, and the other half decelerate them.
pic

A suitable electron beam velocity, as expressed by Eq. (3), can be selected to extend the acceleration distance and duration. This is achieved such that (1) the electron beam velocity is lower than the speed of light in the dielectric at the entrance of each dielectric layer, and (2) the electron beam velocity is higher than the speed of light in the dielectric at the exit of each dielectric layer. In Eq. (3), xe is the longitudinal position of the electron, n is the entry into the nth layer of the dielectric, and i=1n1Li=0 when n=1. Thus, the relative longitudinal displacement of the electron beam stream in the phase is first backward (equivalently, the negative x-direction) and then forward (equivalently, the positive x-direction), extending the duration in the accelerated phase [9,15]. { ve<vo,xe=i=1n1Live>vo,xe=i=1nLi. (3)

In the second row of Eq. (2), Li is the thickness of the layers, known as L1 and L2 in Fig. 1. From recursive formula (2), as long as the prism base angle α, refractive index n1, and thickness L1 of the first layer of the dielectric are specified, the remaining refractive index ni and thickness Li can be obtained recursively. Moreover, Eq. (1) indicates that when the incident electron energy is determined, only one degree of freedom remains between α and n1. Generally, researchers tend to select quartz and other common materials within the range of refractive indices and make the base angle α as small as possible. A positive correlation exists between the acceleration electric field or acceleration gradient and cosα; hence, a smaller base angle α indicates a higher acceleration gradient [17-19]. In addition, the continuous or arbitrary regulation of the dielectric refractive index has led to significant progress in research on optical materials; therefore, a dielectric satisfying Eq. (2) can be fabricated [21-23].

Based on the theoretical analysis above, a two-dimensional numerical simulation study was conducted using the finite element method and the particle-in-cell algorithm. The simulated structure is shown in Fig. 2. The base angle α of the prism was set to 30°, and the refractive index and thickness of each layer were calculated using Eqs. (2) and (3), respectively. A symmetric double prism was used, in which the transverse forces on both sides of the channel centerline are of equal magnitude but in opposite directions. This configuration prevents beam deflections. The geometric parameters of the simulated structures are listed in Table 1. The laser was set as a-10.00μm far-infrared CO2 laser. Because the timescale of the simulated physical process is similar to the pulse width of widely used CO2 lasers [24-28], the laser was set as a continuous plane wave with an electric field amplitude of E0=5.0 GVm1. The selected value implies that the power density of the laser beam is approximately  6.6×1012 W/cm2, and the energy density of the laser pulse irradiating the prism is approximately 0.2 J/cm2. These values are consistent with the laser parameters used in a related study [17]. We propose that line-shaped laser pulses collimated by a cylindrical lens be utilized to drive the symmetric prism, enabling long-distance synchronization. With an energy of 50.00 keV, the injected electrons were in the low-energy nonrelative range (β1) typical for monoenergetic electron beams generated from electron guns with no initial energy dispersion and a charge of 9.36 fC. The particles in the bunch followed a KV distribution. The KV distribution uniformly places the particles in the phase space. The distance between the particles was approximately the same throughout the beam. The initially normalized emittance was 103 pmrad. The initial maximum transverse distance relative to the center of the beam was 0.1 μm, and the maximum divergence angle was approximately 23 mrad. The longitudinal beam length was 3.34 fs, and the distance between the center of the electron beam and prism Surface I was 0.5 μm. These design parameters are consistent with those of the electron beam generated using the electron gun of the DLA [29-31], and the covered parametric range has not yet been investigated for the inverse Cherenkov-effect particle accelerator [16-19].

Table 1
Simulation parameters of geometric structure
Section Length (μm) Refractive index
First 31.31 3.96
Second 10.44 3.74
Third 13.92 3.56
Fourth 18.56 3.40
Fifth 24.74 3.21
Structural Spacing 1.00 1.00
Show more
3

Simulation results and discussion

The energy spectrum of the accelerated electron beam stream is shown in Fig. 3(a). Figure 3(a) corresponds to the instant at 600.42 fs. Here, the electron beam had completely accelerated. The peak energy was 309.55 keV, and the full-width at half-maximum (FWHM) of the energy dispersion was 2.00 keV, corresponding to a relative ratio of 0.65%. Correspondingly, an electron energy gain of 259.55 keV was achieved with an average acceleration gradient of 2.62 GeVm1 and an acceleration factor Eacc/E00.52. This was obtained by dividing the total energy gain of the electrons by the entire path length of approximately 98.96 μm. The phase-space diagram of the horizontal (y-direction) space is shown in Fig. 3(b), which corresponds to the same moment as in Fig. 3(a). The maximum divergence angle of the electron beam did not exceed 21 mrad. Compared with the initial value of approximately 23 mrad, the divergence angle was reduced, and a weak focusing effect was achieved. Fig. 3(c) shows the spatial (x-direction) of the electron beam and the energy distribution with time, with the sampling points separated by 0.25 T0 or approximately 8.34 fs The electron energy increased monotonically, and continuous acceleration was achieved. The curve representing the electron beam energy increased smoothly through the vacuum channel, with only some jitter at the layer interfaces. This smooth increase indicates that the acceleration gradient does not change significantly at each stage; thus, the proposed stacked acceleration structure achieves long-range acceleration with a high acceleration gradient. The electron beam underwent diffusion and recompression in the spatial x-direction, and the energy of the electron beam underwent diffusion and recompression. The stretched and diffused of the electron beam are caused by inhomogeneous field distribution in the accelerated vacuum channel, as reported in the literature [17-19]. When the diffusion of the electron beam stream becomes very strong, the positive phase (Ex>0), that is, the decelerating phase-electric field in front and behind the electron beam, suppresses it, introducing negative feedback. Consequently, the length of the accelerated electron beam is limited to half of the laser cycle, and the extremely-fast electrons slow down, inhibiting their energy increment and minimizing the energy dispersion. This also reduces the speed of the fast-electron beams and concentrates the electrons in the movement direction. This technique facilitates the electron beam to achieve its quasi-monoenergy and compression in the direction of motion after passing through the channel. This property may be valuable in future ultrashort pulse electron beam streams such as attosecond pulses, ultrafast electron diffraction, and other fields. In addition, the electron charge for the fluctuation of the electron beam current charge was observed (Fig. 3(d)), and the simulation demonstrated that no beam loss occurred during the electron beam current acceleration. This finding is consistent with the earlier analysis that the field generated by this stacked structure does not cause transverse (y-direction) deflection of the electron beam current. Furthermore, the width of such a vacuum channel is adequate.

Fig. 3
(Color online) (a) Electron energy spectrum at end of acceleration, normalized to peak. (b) Phase-space diagram of horizontal (y-direction) space. The vertical coordinate is the divergence angle, and the horizontal coordinate is the lateral position of the electron. (c) Spatial (x-direction) and energy distribution of electron beam at each moment (difference of 0.25 T0 for each data sampling point). (d) Total power variation in electron beam at each moment (difference of 0.25 T0 for each data sampling point).
pic

The spatial and temporal variations in the longitudinal electric field Ex along the central axis of the vacuum channel were investigated to further examine the acceleration process of the electron. The path of an electron carrying average energy was adopted to approximate the mean trajectory of the electron beam, as shown in Fig. 4. This demonstrates how an electric field induces electron acceleration or deceleration. The blue-violet region represents the deceleration field, whereas the orange-red zone represents the acceleration field. The net energy gain of a particle is closely related to the phase in which the electron beam is placed in the electric field, Ex. In the phase of the laser electric field, the solid green line represents the global line of the particle beam with average velocity. The electron beam is synchronized with the laser pulse and is always in the negative phase of the electric field (Ex<0), known as the acceleration phase. This is why a stacked structure can achieve a high acceleration gradient for long-distance acceleration. Accelerating a single beam does not fully utilize all the acceleration phases. Therefore, in the future, a multibeam cluster acceleration mode similar to modern conventional accelerators, with each pair of neighboring electron beam clusters spaced by a deceleration phase, can be adopted to improve the energy conversion efficiency and electron beam power.

Fig. 4
(Color online) Electric field Ex (magnitude normalized with respect to E0) as functions of time and space (x-position). The vector of this field is in the lateral direction of the vacuum channel. The green line indicates the electron trajectory, and the sampling time step is 0.25T0. The color bar on the right represents the normalized field strength, Ex/E0 ratio.
pic

Moreover, the performance of each layer of the structure was assessed to adapt it to future chip accelerator applications and construct a structure that can be dismantled to alter the electron beam energy to satisfy the needs of various situations. The energy spectrum of the electron beam stream produced by the acceleration of each layer is shown in Fig. 5. The moments indicated by different colors in Fig. 5 correspond to the electron energy spectra obtained from the first layer to the sixth layer of the dielectric acceleration. The values are listed in Table 2. The energy dispersion of the intermediate electron beam flow was relatively high. However, none of the energy dispersions exceeded 4.00 keV of the FWHM, and none of the corresponding relative ratio values exceeded 2.00%. The peak energy of the electron beam obtained from the final acceleration was 309.55 keV, and the FWHM of the energy dispersion did not exceed 2.00 keV. This indicates that the intermediate electron beam flow can also be used as the injection step, as discussed in the next section.

Table 2
Electron beam profile and acceleration performance for each layer
Section Electron energy (keV) Half-height-width (keV) Ratio (%) Electron gain energy (keV) Average acceleration gradient (GeVm1)
First 125.55 0.34 0.27 75.55 2.41
Second 137.15 2.40 1.75 11.60 1.11
Third 178.75 3.53 1.97 41.60 2.99
Fourth 225.85 1.67 0.74 47.10 2.54
Fifth 309.55 2.00 0.65 83.70 3.38
Total 309.55 2.00 0.65 259.55 2.62
Show more
Fig. 5
(Color online) Electron beam energy spectrum corresponding to end of acceleration of each layer of dielectric.
pic

Table 2 presents the results of an investigation of the acceleration efficiency, in which the energy gain in the electron beam stream in each accelerated stage is determined, together with the corresponding average acceleration gradient. This coincided with the case presented in Fig. 3(c).

This section presents the expandability of the proposed structure. First, the acceleration distance required to accelerate a 50 keV electron beam to 1 MeV was approximately 362.60 μm using an average acceleration gradient of 2.62 GeVm1 for the entire structure. Second, the thickness and refractive index of the dielectric to be added could be obtained using Eqs. (2) and (3). The length (x-direction or Surface I direction) and height (y-direction or Surface III direction, the highest point of the prism) of each subsequent layer are listed in Table 3. In this case, the acceleration structure of the first five segments used the above structure. Furthermore, preliminary estimates were obtained for longer acceleration distances and higher energy gains. The electron relativistic velocity β and dielectric refractive index n changed with increasing electron energy (Fig. 6(a)). The pinch angle was based on the same initial angle as used in this study. The refractive index of the dielectric rapidly decreased as the electron energy increased, and the refractive index eventually reached 2.00 and remained constant. Fig. 6(b) shows the variation in the thickness of each section of the dielectric structure. The thickness of each stage increased exponentially as the number increased. Dielectrics of this size are easy to prepare. In addition, the laser could be infinitely extended in one direction (with the focal length or extendable distance significantly longer than the acceleration distance) and tightly focused in the other direction (laser wavelength magnitude) using beam spreading and line-shaped collimation [32]. This implies that on Surfaces I and II (Fig. 1), the passing distance of the laser in the dielectric is significantly longer than the required acceleration distance. In general, the proposed structure can be extended to higher energy targets by adding a new dielectric layer behind the structure.

Table 3
Size of subsequent structure
Section  Length  (μm) Height (μm)
Sixth 32.99 76.18
Seventh 43.98 101.58
Eighth 58.64 135.44
Ninth 78.19 180.58
Tenth 104.26 240.78
Show more
Fig. 6
(Color online) (a) Electron relativistic velocity (blue line) and dielectric refractive index (red line) as functions of electron energy; (b) Length of each section of dielectric structure (The green and brown lines indicate measurements in microns and meters, respectively, while logarithmic coordinates are used.).
pic
4

Conclusion

A laser dielectric accelerator is proposed based on the inverse Cherenkov phenomenon using a multilayer structure. A parametric design approach was applied to this structure and is demonstrated with a design example for accelerating electron beam currents ranging from 50.00 to 309.55 keV. This suggests that the stackable structure requires a high-energy-gain cascade acceleration technique featuring a high acceleration gradient and an extended acceleration distance, as demonstrated via the simulation. Furthermore, this stackable accelerating structure has a flexible configuration. Theoretically, the structure can achieve arbitrarily high acceleration under design assumptions by extending the accelerator length following the proposed design procedure. The proposed structure is a promising prototype for use in future chip accelerators.

References
[1] Y. Zhang, W.-C. Fang, X.-X. Huang et al.,

Design, fabrication, and cold test of an S-band high-gradient accelerating structure for compact proton therapy facility

. Nucl. Sci. Tech. 32, 38 (2021). doi: 10.1007/s41365-021-00869-z
Baidu ScholarGoogle Scholar
[2] W. Nan, B. Guo, C.-J. Lin et al.,

First proof-of-principle experiment with the post-accelerated isotope separator on-line beam at BRIF: measurement of the angular distribution of 23Na + 40Ca elastic scattering

. Nucl. Sci. Tech. 32, 53 (2021). doi: 10.1007/s41365-021-00889-9
Baidu ScholarGoogle Scholar
[3] Z.-Y. Ma, S.-J. Zhao, X.-M. Liu et al.,

High RF power tests of the first 1.3 GHz fundamental power coupler prototypes for the SHINE project

. Nucl. Sci. Tech. 33, 10 (2022). doi: 10.1007/s41365-022-00984-5
Baidu ScholarGoogle Scholar
[4] B. Naranjo, A. Valloni, S. Putterman et al.,

Stable charged-particle acceleration and focusing in a laser accelerator using spatial harmonics

. Phys. Rev. Lett. 109, 164803 (2012). doi: 10.1103/PhysRevLett.109.164803
Baidu ScholarGoogle Scholar
[5] E.A. Peralta, K. Soong, R.J. England et al.,

Demonstration of electron acceleration in a laser-driven dielectric microstructure

. Nature 503, 91-94 (2013). doi: 10.1038/nature12664
Baidu ScholarGoogle Scholar
[6] R.J. England, R.J. Noble, K. Bane et al.,

Dielectric laser accelerators

. Rev. Modern Phys. 86, 1337-1389 (2014). doi: 10.1103/RevModPhys.86.1337
Baidu ScholarGoogle Scholar
[7] Y.-F. He, B. Sun, M.-J. Ma et al.,

Topology optimization of on-chip integrated laser-driven particle accelerator

. Nucl. Sci. Tech. 33, 120 (2022). doi: 10.1007/s41365-022-01101-2
Baidu ScholarGoogle Scholar
[8] K.J. Leedle, R.F. Pease, R.L. Byer et al.,

Laser acceleration and deflection of 96.3 keV electrons with a silicon dielectric structure

. Optica 2, 158-161 (2015) doi: 10.1364/OPTICA.2.000158
Baidu ScholarGoogle Scholar
[9] K.P. Wootton, Z.R. Wu, B.M. Cowan et al.,

Demonstration of acceleration of relativistic electrons at a dielectric microstructure using femtosecond laser pulses

. Opt. Lett. 41, 2696-2699 (2016) doi: 10.1364/OL.41.002696
Baidu ScholarGoogle Scholar
[10] J. McNeur, M. Kozák, N. Schönenberger et al.,

Elements of a dielectric laser accelerator

. Optica 5, 687-690 (2018) doi: 10.1364/OPTICA.5.000687
Baidu ScholarGoogle Scholar
[11] P. Yousefi, N. Schönenberger, J. Mcneur et al.,

Dielectric laser electron acceleration in a dual pillar grating with a distributed Bragg reflector

. Opt. Lett. 44, 1520-1523 (2019) doi: 10.1364/OL.44.001520
Baidu ScholarGoogle Scholar
[12] N.V. Sapra, K.Y. Yang, D. Vercruysse et al.,

On-chip integrated laser-driven particle accelerator

. Science 367, 79-83 (2020). doi: 10.1126/science.aay5734
Baidu ScholarGoogle Scholar
[13] R. Shiloh, J. Illmer, T. Chlouba et al.

Electron phase-space control in photonic chip-based particle acceleration

. Nature 597, 498-502 (2021). doi: 10.1038/s41586-021-03812-9
Baidu ScholarGoogle Scholar
[14] Y. Meng, Y. Chen, L. Lu et al.,

Optical meta-waveguides for integrated photonics and beyond

. Light Sci. Appl. 10, 235 (2021). doi: 10.1038/s41377-021-00655-x
Baidu ScholarGoogle Scholar
[15] T. Chlouba, R. Shiloh, P. Forsberg et al.,

Diamond-based dielectric laser acceleration

. Opt. Express 30, 505-510 (2022) doi: 10.1364/OE.442752
Baidu ScholarGoogle Scholar
[16] M. Kozák, P. Beck, H. Deng et al.,

Acceleration of sub-relativistic electrons with an evanescent optical wave at a planar interface

. Opt. Express 25, 19195-19204 (2017) doi: 10.1364/OE.25.019195
Baidu ScholarGoogle Scholar
[17] W. Liu, Z. Yu, L. Sun et al.,

Microscale laser-driven particle accelerator using the inverse Cherenkov effect

. Phys. Rev. Appl. 14, 014018 (2020). doi: 10.1103/PhysRevApplied.14.014018
Baidu ScholarGoogle Scholar
[18] L. Sun, W. Liu, J. Zhou et al.,

GV m−1 on-chip particle accelerator driven by few-cycle femtosecond laser pulse

. New J. Phys. 23, 063031 (2021). doi: 10.1088/1367-2630/ac03cf
Baidu ScholarGoogle Scholar
[19] W.H. Liu, L. Sun, Z.J. Yu et al.,

THz-driven dielectric particle accelerator on chip

. Opt. Lett. 46, 4398-4401 (2021) doi: 10.1364/OL.430451
Baidu ScholarGoogle Scholar
[20] Y. Wei, M. Ibison, G. Xia et al.,

Dual-grating dielectric accelerators driven by a pulse-front-tilted laser

. Appl. Opt. 56, 8201-8206 (2017). doi: 10.1364/AO.56.008201
Baidu ScholarGoogle Scholar
[21] G. Pilania, E. Weis, E. M. Walker et al.,

Computational screening of organic polymer dielectrics for novel accelerator technologies

. Scientific Reports 8, 9258 (2018). doi: 10.1038/s41598-018-27572-1
Baidu ScholarGoogle Scholar
[22] D.H. Lippman, N.S. Kochan, T. Yang et al.,

Freeform gradient-index media: a new frontier in freeform optics

. Opt. Express 29, 36997-37012 (2021). doi: 10.1364/OE.443427
Baidu ScholarGoogle Scholar
[23] R. Dylla-Spears, D. Yee Timothy, K. Sasan et al.,

3D printed gradient index glass optics

. Science Advances 6, eabc7429. doi: 10.1126/sciadv.abc7429
Baidu ScholarGoogle Scholar
[24] W.D. Kimura, I.V. Poaorelsky, L. Schächter,

CO2-laser-driven dielectric laser accelerator

. In 2018 IEEE Advanced Accelerator Concepts Workshop (AAC), (2018), 1-5. doi: 10.1109/AAC.2018.8659403
Baidu ScholarGoogle Scholar
[25] M.N. Polyanskiy, I.V. Pogorelsky, V. Yakimenko,

Picosecond pulse amplification in isotopic CO2 active medium

. Opt. Express 19, 7717-7725 (2011) doi: 10.1364/OE.19.007717
Baidu ScholarGoogle Scholar
[26] M.N. Polyanskiy, M. Babzien, I.V. Pogorelsky,

Chirped-pulse amplification in a CO2 laser

. Optica 2, 675-681 (2015) doi: 10.1364/OPTICA.2.000675
Baidu ScholarGoogle Scholar
[27] M.N. Polyanskiy, I.V. Pogorelsky, M. Babzien et al.,

Demonstration of a 2 ps, 5 TW peak power, long-wave infrared laser based on chirped-pulse amplification with mixed-isotope CO2 amplifiers

. OSA Continuum 3, 459-472 (2020) doi: 10.1364/OSAC.381467
Baidu ScholarGoogle Scholar
[28] P. Panagiotopoulos, M.G. Hastings, M. Kolesik et al.,

Multi-terawatt femtosecond 10 µm laser pulses by self-compression in a CO2 cell

. OSA Continuum 3, 3040-3047 (2020) doi: 10.1364/OSAC.399992
Baidu ScholarGoogle Scholar
[29] J. Hoffrogge, J. Paul Stein, M. Krüger et al.,

Tip-based source of femtosecond electron pulses at 30 keV

. J. Appl. Phys. 115, 094506 (2014). doi: 10.1063/1.4867185
Baidu ScholarGoogle Scholar
[30] M. Krüger, C. Lemell, G. Wachter et al.,

Attosecond physics phenomena at nanometric tips

. J. Phys. B Atomic, Molecular and Optical Physics 51, 172001 (2018). doi: 10.1088/1361-6455/aac6ac
Baidu ScholarGoogle Scholar
[31] T. Hirano, K.E. Urbanek, A.C. Ceballos et al.,

A compact electron source for the dielectric laser accelerator

. Appl. Phys. Lett. 116, 161106 (2020). doi: 10.1063/5.0003575
Baidu ScholarGoogle Scholar
[32] P.A. Naik, S.R. Kumbhare, V. Arora et al.,

A new technique for obtaining uniform intensity line focus of Gaussian laser beams

. Optics Communications 223, 137-142 (2003). doi: 10.1016/S0030-4018(03)01544-X
Baidu ScholarGoogle Scholar