Gyromonotrons are typically believed to rely on the convective interaction between the co-propagating beam and wave, with the extended energy-exchanging process stemming from the external feedback. However, numerous studies focusing on both transverse electric (TE) mode and transverse magnetic (TM) mode gyrotrons have consistently shown that beam–wave interactions in weak-feedback systems or even in uniform tubes without any structural feedback can yield a theoretical beam efficiency of more than 30% with major forward-wave output during near-cutoff operation, which is the typical operating condition for gyromonotrons. These intriguing findings raise questions about the actual feedback mechanism of gyromonotrons. In this article, comparative studies on the linear and nonlinear behaviors of uniform-tube gyrotron are investigated. The forward and backward waves are observed to co-generate and exhibit similar characteristics of ultra-slow group velocity under near-cutoff operation. This situation allows the as-generated forward wave to modulate the fresh beam, establishing a new backward-wave-like internal feedback loop. Additionally, the quasi-degenerate nature of the bi-directional propagating waves ensures their intrinsic in-phase relationship. The consequent constructive interference enables the uniform tube to function as a high-Q resonator. These findings are found to be independent of the choices of TE or TM modes, providing valuable insights into the underlying interaction mechanism of gyrotron devices.

Gyrotrons, utilizing the electron cyclotron maser interaction to generate coherent radiation,1–3 have been experimentally demonstrated as high-power (up to sub-gigawatt level) and high-frequency (up to the terahertz range) sources.4–8 These powerful radiation sources have found numerous applications in plasma heating for fusion, communication, sensing, security, and biomedical diagnostics.9–12 Based on the feedback mechanism, gyrotron oscillators are traditionally categorized into two types the gyrotron backward-wave oscillator (gyro-BWO) and the gyromonotron. In gyro-BWO, oscillation13 arises as the forward-moving electron beam undergoes continuous modulation through the backward radiation emitted by preceding electrons. The modulation induces a stronger backward radiation tendency in the beam, leading to the gradual formation of an internal feedback loop and absolute instability.14,15 This instability allows the gyro-BWO mode to grow and the field profile contracts to the beam entrance end.16 As the magnetic field (B0) or the beam voltage (Vb) is tuned, the electron transit angle of gyro-BWO modes varies simultaneously to satisfy the beam–wave resonance condition, thereby enabling frequency tuning.14 On the other hand, the gyromonotron operates in a resonant structure.17–19 Its oscillation is believed to be formed primarily by reflective feedback due to structural nonuniformities.1,6,20–23 Previous studies have demonstrated that in a gyromonotron device, the electron beam tends to resonantly interact with the fundamental cavity axial mode,18 built up by the interference of forward and backward waves. The resonant conditions of the interaction cavity therefore restrict the frequency tunability of each gyromonotron mode as B0 or Vb is adjusted.24 However, the resonant enhancement renders the gyromonotron a high-efficiency microwave source.25,26

Almost half a century ago, Bratman et al. conducted a study on the non-fixed fields of transverse electric (TE) modes within a semi-infinite low-Q regular tube, featuring a structure cutoff at the upstream end and an abrupt drop of static magnetic field (magnetic-field cutoff) right before the downstream end27 to physically obtain a finite interaction region. The outgoing-wave (no-reflection) boundary condition was adopted in the model to analyze gyrotron devices in a self-consistent manner. They observed an exceptionally high transverse peak interaction efficiency of 0.75, at which the field profile closely resembles that of a resonator mode with weak phase variation across the entire interaction length. The feedback pathway was attributed to the downstream reflected wave generated at the cross section where resonant interaction of electrons with the radio frequency (RF) field completely ceases. As this backward reflection propagates to the upstream, it is turned into the forward wave due to the presence of cutoff, thereby forming the feedback loop. A subsequent study using a similar three-section low-Q cavity with an oversized end pointed out the involvement of both the forward and backward waves in shaping the high-Q resonant field profile.28 After these works, an increasing number of research efforts18,19,29–34 have been devoted to self-consistent investigation of low-Q gyromonotron systems. Fliflet et al.19 further showcased the persistence of strong beam–wave interaction of the TE01 mode in a low-Q RF system considering a uniform magnetic field, and the calculated results were compared with the experimental verification.

It is important to notice that those low-Q gyromonotron systems usually feature an upstream structural irregularity (e.g., cutoff) and a downstream beam irregularity (by either closing the magnetic guiding field or applying simulation boundaries to terminate the beam–wave resonant interaction). However, the extent to which these irregularities induce reflection and whether such reflection is adequate to produce a strong feedback for generating high-efficiency forward output has not been thoroughly examined. That raises a fundamental question regarding the underlying mechanism of the observed oscillations. Do the oscillations necessarily rely on the recurrent conversion between forward and backward waves formed by either tube irregularities or inhomogeneities of the electron-stream conductivity caused by the variation of the static magnetic field28 or the presence of the imposed boundary conditions?

For the sake of simplification, recent studies have intended to investigate the gyrotron oscillator in its relatively simplified configuration, i.e., a uniform tube under a constant DC magnetic bias.14,16,35–37 In such a setup, the outgoing-wave boundary conditions are adopted at both ends, as Bratman et al. did in their model.27,28,30 Among these works, Hung et al.35 have successfully attained a low starting current and achieved a high backward output efficiency in generating the TE52 mode. This work also demonstrated an oscillation frequency very close to the cutoff and a resonance mode profile, all achieved without requiring any external feedback. On the other hand, in the case of the transverse magnetic (TM) mode gyrotron implementing TM11-mode, a nonlinear and self-consistent simulation demonstrated a peak interaction efficiency of 32%, primarily contributed by the generation of the forward wave.37 

These observations suggest that a system without any external feedback can still effectively function as a gyromonotron-like oscillator, delivering a strong output toward the collector end, regardless of the operating mode being TE or TM. Several thought-provoking questions then arise: What is the underlying feedback mechanism that governs the gyrotron in the near-cutoff region, resulting in the generation of strong forward waves? Does this interaction encompass an absolute instability characterized by the internal feedback, reminiscent of the gyro-BWO,1,14–16 or does it resemble a convective instability as seen in gyro-traveling-wave tube (TWT)?1,13,38,39 Moreover, what are the corresponding inherent bunching mechanisms for TE and TM modes at the near-cutoff gyromonotron regime?

In this paper, we present comprehensive analyses of the TM11-mode and the TE01-mode gyrotrons, elucidating the physics contributing to their high output efficiencies (>30%) when implemented with a uniform tube biased by constant static magnetic field. Our investigation begins with a nonlinear and self-consistent particle-tracing simulation of the TM11-mode gyrotron (Sec. II). By examining the behavior of the forward and backward components under magnetic field tuning, we provide a qualitative explanation based on their detuning from the beam–wave resonance conditions. To further clarify the underlying feedback mechanism, the impact of introducing heavy Ohmic wall loss to both the upstream and downstream sections of the system is explored (Sec. III). In Sec. IV, we delve into the linear stage of the system and discuss the linear behavior (start-oscillation currents and field profiles) of the TM11 mode. By examining the electron transit-angle phases, we unveil that the high efficiency observed in a uniform tube results from the concurrent presence of a forward convective energy exchange process and a backward internal feedback mechanism. Furthermore, we study the TM-mode bunching mechanisms in Sec. V, revealing that the co-generated forward and backward waves can together enhance electron bunching near the cutoff region of the TM mode. In Sec. VI, the analyses in Secs. II–V are applied to the TE01-mode gyrotron. An analogous conclusion is drawn, demonstrating that the proposed fundamental mechanisms hold true under near-cutoff operation regardless of the mode choices. The identification of the dominated instability and several comparisons and the remarks of the present findings are provided in Sec. VII. Finally, in Sec. VIII, we present our conclusions, summarize the key findings and implications of our study, and provide a concluding remark.

For the demonstration purpose, the TM11 mode is selected as the first testing mode in a uniform waveguide with a radius (rw) of 0.2 cm and a constant static magnetic field B0 across the entire tube. An axis-encircling beam with an electron's guiding center radius (rc) of 0 cm is chosen for the optimum beam–wave coupling strength.40,41 Throughout this work, the beam voltage Vb and the pitch factor α (electron transverse-to-axial velocity ratio) are set at 70 kV and 1.0, respectively. These parameters are consistent with those employed in the recent publication.37  Fig. 1(a) shows the TM11 dispersion (Re[kz]) and the beam–wave resonant lines (ωRe[kz]vz0Ωc0=0) at multiple magnetic fields B0, where ωc is the cutoff frequency of the mode, c is the speed of light in vacuum, vz0 is the electron's initial axial velocity, and Ωc0=eB0/γ0mec is the initial relativistic gyrating frequency of the electron. e and me are, respectively, the charge and the rest mass of electron, and γ0=(1v02/c2)1/2 is the Lorentz relativistic factor, in which v0=(v02+vz02)1/2 and v0 are, respectively, the electron's initial velocity and transverse velocity. Only the fundamental harmonic is considered to make the discussion concise. The exact formula of waveguide mode's propagation constant kz can be found in the  Appendix [Eq. (A2)]. In the present case, the corresponding grazing magnetic field Bg=(γ0mec/e)(ωc/γz0), where γz0=(1vz02/c2)1/2 is 34.96kG, at which the beam–wave resonance line is tangent to the mode dispersion curve. Beyond Bg, the intersection between the beam mode and the waveguide mode signifies the most preferred emission direction of oscillation. In the following, the terms “gyromonotron region” and “gyro-BWO region” refer to the conditions where the beam line intersects, respectively, with the forward branch and the backward branch of the waveguide mode.

FIG. 1.

(a) TM11 waveguide-mode dispersion and the beam–wave resonance lines at various magnetic fields B0. (b) Oscillation frequency ω0 and detuning factor of forward/backward wave (εfwd/bwdNL) vs B0 for L = 3.5 cm. The displayed quantities are all normalized to the TM11 cutoff ωc. The corresponding beam efficiency ηbeam, forward-wave efficiency ηfwd, and backward-wave efficiencies ηbwd of uniform tubes with various lengths L are depicted in panels (c), (d), and (e), respectively. The tube's wall radius rw is 0.2 cm, with copper resistivity ρCu of 1.72  × 10−8 Ω⋅m. The beam parameters are Vb = 70 kV, Ib = 5 A, and α = 1.0.

FIG. 1.

(a) TM11 waveguide-mode dispersion and the beam–wave resonance lines at various magnetic fields B0. (b) Oscillation frequency ω0 and detuning factor of forward/backward wave (εfwd/bwdNL) vs B0 for L = 3.5 cm. The displayed quantities are all normalized to the TM11 cutoff ωc. The corresponding beam efficiency ηbeam, forward-wave efficiency ηfwd, and backward-wave efficiencies ηbwd of uniform tubes with various lengths L are depicted in panels (c), (d), and (e), respectively. The tube's wall radius rw is 0.2 cm, with copper resistivity ρCu of 1.72  × 10−8 Ω⋅m. The beam parameters are Vb = 70 kV, Ib = 5 A, and α = 1.0.

Close modal
Under the single-mode assumption, the nonlinear and self-consistent particle-tracing method37 is employed to simulate a case with an injected current Ib of 5 A. The outgoing-wave boundary conditions,27,30 with f(0)=ikzf(0) and f(L)=ikzf(L), are adopted for solving the oscillating frequency and the overall field amplitude, in which f(z) and f(z) are the field profile and its first-order derivative. In all calculations, the boundary conditions are satisfied to the level of 10−10, implying that the incoming forward (backward) wave amplitude at upstream (downstream) is less than 10−10. The governing equations, the definitions of relevant parameters, and the method implemented for the calculation are summarized in the  Appendix. To avoid complexity, the present study only focuses on the soft-excitation (self-excitation) cases, where the oscillation grows from a noise level once Ib is beyond the start-oscillation threshold. The solved oscillation frequencies (ω0) as functions of B0 when L = 3.5 cm is demonstrated in Fig. 1(b). For the cases with different L, nearly the same oscillation frequency is obtained at a specific B0, which is thus not shown for clarity. The frequencies in Fig. 1(b) are in close proximity to the TM11 cutoff (ωc/2π = 91.42 GHz, at the W-band). In order to gain a deeper understanding of the inherent feedback loop, differentiating the roles of forward and backward waves during their interaction with an electron beam is essential. Refer to the  Appendix for the details of field decomposition as well as the calculations for the corresponding efficiencies discussed herein. Figures 1(c)–1(e), respectively, depict the total beam efficiency (ηbeam), the forward-output efficiency (ηfwd), and the backward-output efficiency (ηbwd) at various L. As L increases from 1.7 to 3.0 cm, the peak beam efficiencies drastically increase and shift toward the high-B0 region. At L = 3 cm, the maximum ηbeam achieves over 32% at B0 ∼ 36 kG, contributed majorly by the forward oscillation (ηfwd ∼ 23%) and partially by the backward oscillation (ηbwd ∼ 7.5%). Noticeably, ηfwd exhibits a hill shape under B0 tuning, while ηbwd remains relatively flat. A similar efficiency profile was observed for TE-mode gyrotron oscillators27 by Bratman et al. It should be pointed out that the simulation models utilized by Bratman et al. and the current work are similar, differing only slightly in the origins of beam terminations. The outgoing-wave boundary conditions physically match the electromagnetic fields interacting with the beam inside the tube and those propagating without being affected by the beam outside the tube. This equivalent beam termination at the boundary may introduce an artificial feedback, which was also believed to generate the strong forward wave observed in a semi-infinite cavity with an abrupt-drop magnetic field.27 However, in the present study, a similar phenomenon of high-efficiency forward wave is also observed, while the reflection resulting from artificial beam irregularity is maintained at a very low level of 10−10. This quantitative examination suggests that the previously proposed feedback mechanism may not be the major cause. The actual underlying mechanism and the reasons for similar phenomena occurring in both TE and TM modes will be further discussed in Secs. V–VII. The observations in Figs. 1(c)–1(e) can be qualitatively explained by the degree of deviation from the ideal beam–wave resonant condition,
(1)

Figure 1(b) shows εfwd and εbwd with superscripts “NL,” meaning they are calculated based on the solved oscillation frequency in the nonlinear region (ω0). Note that subscripts “0” of vz0 and Ωc0 represent their initial values at the entrance of the beam. Although Eq. (1) might introduce errors due to the omission of the beam spread caused by nonlinear beam–wave interaction, these quantities still serve as useful approaches for qualitative elucidation. When B0 is adjusted, εfwdNL displays a valley pattern as indicated by the L.H.S. arrow in Fig. 1(b), while εbwdNL remains relatively unchanged. A smaller detuning factor implies better synchronization and hence higher interaction efficiency in the soft-excitation regime. εfwdNL and εbwdNL therefore exhibit inverse responses to the output efficiencies ηfwd and ηbwd, respectively. Since εfwdNL is smaller than εbwdNL at the near-cutoff region (roughly below 37.3 kG), the forward wave appears to dominate the beam–wave interaction. At B0 ∼ 37.3 kG [marked by the R.H.S. arrow in Fig. 1(b)], εfwdNL and εbwdNL intersect with each other, indicating the crossover behavior of ηfwd and ηbwd, as can be seen by comparing Figs. 1(d) and 1(e). This manifests the gradual transition from gyromonotron to gyro-BWO, where the dominant beam–wave interaction shifts from the forward wave to the backward wave. Note that in cases of hard excitation, even higher efficiency may be achieved by large detunings.7 

Once L exceeds 3.0 cm, the fluctuation in ηbeam becomes apparent across the entire B0-tuning region. A similar response is observed for ηfwd. Regardless of the increase in L, ηfwd remains primarily within the upper bound observed as L ∼ 3.0 cm. The fluctuating ηfwd spectrum arises from the iterative beam-energy deposition and absorption. To illustrate this concept, the forward-wave power profile Pfwd at B0 = 36 kG is plotted in Fig. 2(a). For L = 1.7 to 3.0 cm, the longer tube allows more space for the beam–wave energy exchange, leading to the gradual increase in the peak forward-wave power. As L exceeds a certain threshold, the electron beam starts to reabsorb the wave energy gained at the earlier stage (upstream), leading to a decrease in the forward output power at the downstream end. The cases with L = 3.0 to 5.0 cm presented in Fig. 2(a) evidence this statement. Accordingly, the fluctuating ηfwd spectra observed at different L is attributed to the truncation of the beam–wave energy deposition and re-absorption process at the downstream port. On the contrary, ηbwd in Fig. 1(e) gradually saturates and remains flat even when L approaches 5.0 cm. At the gyro-BWO region (e.g., B0 = 38 kG), when L exceeds a relaxation length,15 the backward output efficiency would saturate due to the nonlinear field contraction. Any further increase in length beyond the relaxation length only contributes a marginal tailing field structure to the overall field distribution.16 Interestingly, a similar phenomenon is also observed in the gyromonotron case of B0 = 36 kG, as evident from the saturated and contracted backward-wave power profiles Pbwd in Fig. 2(b).

FIG. 2.

The forward-wave (a) and the backward-wave (b) power profiles at B0 = 36 kG. The lengths of tubes considered in the calculation correspond to those in Fig. 1.

FIG. 2.

The forward-wave (a) and the backward-wave (b) power profiles at B0 = 36 kG. The lengths of tubes considered in the calculation correspond to those in Fig. 1.

Close modal

For gyro-BWOs, the backward wave plays an essential role in energy extraction, which is confirmed by negligible ηfwd at B0 = 38 kG. As for the case in the gyromonotron region (B0 = 36 kG), while the beam is coupled strongly with the forward wave, the backward wave always survives. The origin of the backward wave is understood by the intrinsic internal feedback, whereas that of the forward wave remains elusive, especially for operating in a uniform tube without any external feedback, such as the absence of structural or magnetic-field inhomogeneities. Although Fig. 2(a) appears to demonstrate the convective growth of the forward wave, the coexistence of forward and backward components operating near cutoff seems to imply the simultaneous presence of the forward-manner convective interaction and the backward-manner internal feedback. To further clarify the underlying mechanism, lossy-section tests are employed in Sec. III to assess the response of forward and backward waves.

Applying distributed wall loss to the interaction circuit is an effective way to eliminate the unwanted oscillations.38,42–44 This method is adopted here to test the response of output signals. Based on the nonlinear power profiles displayed in Fig. 2, a tube with a moderate length of L = 3.5 cm is selected. This length ensures adequate spacing between the major-field distributions of the forward and backward components, enabling their individual manipulation. Two lossy-section tests are applied, during which a lossy section of 1 cm is present at the downstream [Figs. 3(a) and 3(b)] and the upstream [Figs. 3(c) and 3(d)].

FIG. 3.

Lossy-section test. The forward-wave and backward-wave efficiencies as functions of B0 in a case with 1 cm lossy section at downstream are displayed in panels (a) and (b), respectively; (c) and (d) demonstrate the results for the 1 cm upstream lossy-section test.

FIG. 3.

Lossy-section test. The forward-wave and backward-wave efficiencies as functions of B0 in a case with 1 cm lossy section at downstream are displayed in panels (a) and (b), respectively; (c) and (d) demonstrate the results for the 1 cm upstream lossy-section test.

Close modal

Figure 3(a) demonstrates that the forward output efficiency is significantly suppressed to below 5% and becomes flat as the downstream Ohmic loss increases from 101ρCu to 104ρCu. This outcome is not surprising since the location of the lossy section coincides with the major field of the forward wave. The energy emitted by the electron beam therefore gets dissipated by the surface current on the waveguide wall directly. This test also illustrates an important fact that the generated forward wave must correspond to the cold-tube mode rather than the hot-tube mode.13 In other words, the highest achieved forward efficiency (ηfwd ∼ 23%) does not rely on the field energy residing within the gyrating electrons themselves. Instead, it relies on the energy being carried by the surface current and the guided mode, which can be subsequently extracted using appropriate couplers. Moreover, due to the nonlinear field contraction, the increased downstream Ohmic loss has a negligible impact on the backward output in the gyro-BWO region. ηbwd can even be slightly enhanced in the region of B0 = 35.5 ∼ 37.5 kG. This slight enhancement might result from the increased reflection at the interface between the two regions with high-contrast in their wall losses, reflecting the generated forward wave back to the upstream direction.

In the other case with the upstream lossy section, it is expected that the increased wall loss should effectively eliminate the backward wave, while having a lesser effect on the generation of the forward wave. Figure 3(d) confirms the first point, showing that ηbwd is reduced by half at 38 kG (gyro-BWO region) and is completely suppressed at 36 kG (gyromonotron region) when the wall resistivity reaches 104ρCu. The distributed upstream loss thus serves as a key to suppress spurious oscillations from the absolute instability, which was utilized to achieve a gyrotron traveling-wave tube (gyro-TWT) with a ultrahigh gain of ∼70 dB.13 However, the result in Fig. 3(c) contradicts the aforementioned expectation on the forward wave generated at the gyromonotron region. It is discovered that ηfwd also decreases significantly as the wall loss increases, synchronizing with the reduction in ηbwd. This observation provides valuable insight into the origin of forward oscillation, suggesting a possibility that a similar internal feedback mechanism, originally associated with the backward wave, may also play a crucial role under the near-cutoff operation adopted by most gyromonotron devices. As a result, when this internal feedback loop is disrupted by the heavy Ohmic wall loss at upstream, all subsequent nonlinear beam–wave interactions are suppressed, causing the complete absence of output. To simplify the analysis of the interplay between forward and backward waves and illustrate their origins, Sec. IV will focus on the physics during the linear growth stage, avoiding the complicated nonlinear effect.

Figure 4(a) plots the starting current Ist and the oscillation frequency ωst of the TM11 mode in a 3.5 cm uniform tube and a constant B0. These values are calculated by solving the relativistic Vlasov equation and Maxwell's equations together. Following linearization using the small-signal approximation, either the Laplace transformation method or the integral method can be employed for the solution.36 To manifest the strong beam–wave interaction at the gyromonotron region, the inverse of Ist is also displayed, which mirrors the trend observed in the nonlinear forward-wave efficiency in Fig. 1(d). Despite the nonlinear effects such as the truncation of beam-energy deposition [Fig. 2(a)] and the nonlinear field contraction [Fig. 2(b)], it becomes apparent that the lower the starting current, the higher the output efficiency. This finding emphasizes that the feedback mechanism at near-cutoff operation continues to play a significant role, even in the linear regime. The presence of such a feedback loop in a finite-length system could lead to the onset of self-oscillation during the linear stage, introducing a new energy-balance condition.16 

FIG. 4.

(a) The starting current Ist (solid line) and its inverse (dotted line) under various magnetic fields. The corresponding oscillation frequency ωst/2π is also displayed in red, referring to the R.H.S. y-axis; (b) and (c), respectively, show the field amplitude profiles |f(z)| and phase profiles f(z) of the examples at gyromonotron (36.3 kG) and gyro-BWO (38 kG). The decomposed field amplitude profiles of the forward wave |ffwd (z)| and the backward wave |fbwd (z)| are also, respectively, displayed in red and blue for discussion.

FIG. 4.

(a) The starting current Ist (solid line) and its inverse (dotted line) under various magnetic fields. The corresponding oscillation frequency ωst/2π is also displayed in red, referring to the R.H.S. y-axis; (b) and (c), respectively, show the field amplitude profiles |f(z)| and phase profiles f(z) of the examples at gyromonotron (36.3 kG) and gyro-BWO (38 kG). The decomposed field amplitude profiles of the forward wave |ffwd (z)| and the backward wave |fbwd (z)| are also, respectively, displayed in red and blue for discussion.

Close modal

Figures 4(b) and 4(c) present the amplitude and the phase of the start-oscillation field f(z) in the cases of 36.3 (gyromonotron region) and 38 kG (gyro-BWO region), respectively. In the former case, the primary field profile extends toward the downstream. The same decomposition method outlined in the  Appendix [Eq. (A10)] can be applied to obtain the forward and backward field coefficients [ffwd(z) and fbwd(z)] at the moment when the oscillation begins. The decomposed two field amplitudes (|ffwd(z)| and |fbwd(z)|) with similar maximum amplitudes reveal that both of the forward and backward waves exhibit fairly good beam–wave synchronization. The forward wave then grows convectively toward the downstream, while the backward wave is generated as it counter-propagates with the electron beam. In addition, the phase of the overall field [ f(z)] remains nearly constant along the axial direction, suggesting the presence of an intrinsic “standing-wave pattern.” Combing the distributions of |ffwd(z)| and |fbwd(z)|, we can conclude that the observed standing-wave pattern is constructed by forward and backward waves with similar amplitudes and nearly equal phases. Under this circumstance, as shown in Fig. 4(a), the oscillation frequency is almost fixed at a “resonant frequency,” however, within a uniform structure without external feedback. In the case of gyro-BWO [Fig. 4(c)], it is observed that the backward wave experiences superior beam–wave synchronization and thus fbwd(z) dominates the oscillation. Unlike the former case, the phase [ f(z)] now evolves toward the upstream during the generation and the propagation of the backward wave, indicating the traveling-wave nature of gyro-BWO. These observations align with the previous research findings.14,28

To analyze how the forward and backward waves are co-generated and interfere with each other, Fig. 5 illustrates their electron transit-angle phases, i.e., Θfwd=εfwd×L/vz0 and Θbwd=εbwd×L/vz0, where εfwd/bwd is the detuning factor defined in Eq. (1) and L/vz0 can be regarded as the electron transit time across the waveguide. At the gyro-BWO region (37 ∼ 38 kG), Θbwd is almost locked at an optimum value (∼1.05π) due to the automatic adjustment of the oscillation frequency, as shown in Fig. 5. This serves as the key mechanism for frequency-tunable gyro-BWOs.14,45 At the gyromonotron region (35.5 ∼ 36.5 kG, specifically), both Θfwd and Θbwd vary with B0. It is important noting that Θfwd and Θbwd are closely aligned with each other (ΘfwdΘbwd0), both slightly deviating from Θ0 ((ωΩc0)×L/vz0). This is associated with the fact of negligible Re[kz] near the cutoff. That is to say, the difference between forward wave and backward wave becomes insignificant from the viewpoint of the electron. Due to this very near-cutoff operation, the group velocity of the wave (vg) is largely reduced, indicating that the electrons propagate at a speed nearly two orders of magnitudes faster than the transportation of forward wave's energy (indicated by the black curve in Fig. 5). In such scenario, the newly injected electrons are afforded the opportunity to catch up the forward wave generated by the preceding electrons. The as-generated forward wave therefore can modulate the fresh electrons, gradually building up the feedback loop. This analysis confirms that the forward and backward waves generated near the cutoff exhibit very similar internal feedback mechanisms. In other words, the convectively growing forward wave now plays the same role as the backward wave, triggered by the absolute instability.

FIG. 5.

The electron transit-angle phases of forward wave Θfwd (in red) and backward wave Θbwd (in blue) under 1ρCu (solid lines) and 10−6ρCu (dotted lines) wall loss. The difference between the two phases is plotted in green. Θ0 serves as the reference phase defined by (ωΩc)×L/vz0. Notice that the dotted lines show the data of an ideal case calculated by assuming ultra-low wall loss of 10−6ρCu. The ratio of the wave group velocity vg to the initial electron's axial speed vz0 is also displayed in black, referring to the R.H.S. y-axis.

FIG. 5.

The electron transit-angle phases of forward wave Θfwd (in red) and backward wave Θbwd (in blue) under 1ρCu (solid lines) and 10−6ρCu (dotted lines) wall loss. The difference between the two phases is plotted in green. Θ0 serves as the reference phase defined by (ωΩc)×L/vz0. Notice that the dotted lines show the data of an ideal case calculated by assuming ultra-low wall loss of 10−6ρCu. The ratio of the wave group velocity vg to the initial electron's axial speed vz0 is also displayed in black, referring to the R.H.S. y-axis.

Close modal

ΘfwdΘbwd0 further implies that the forward and backward waves are almost in-phase, corresponding to the constructive interference of the bi-directional propagating signals. The coherent superposition of these waves results in the formation of a standing-wave field pattern, as illustrated in Fig. 4(b). This phenomenon enables the uniform tube to function as a resonator, even without any external feedback. Consequently, this inherent resonant characteristic guarantees high interaction efficiency as observed. The described phase relationships become clearer as the tube's wall loss is reduced to an extremely low level (10−6ρCu). As demonstrated by the dotted lines in Fig. 5, ΘfwdΘbwd is perfectly locked at zero within the range of 34.5–36.5 kG, corresponding to a perfect on-resonance state. Despite the in-phase relationship and the mutual constructive interreference between the forward and backward waves, it is essential to address whether their corresponding electron bunching effects can effectively cooperate or compete. This issue will be examined in Sec. V.

Without loss of generality, the behaviors of electron bunching in the phase space can be examined by applying the temporal variation on the effective electron cyclotron frequency Ωeff=kzvz+Ωc.1,46 We then obtain
(2)
where
(3)
in which v is the transverse velocity of the electron, B is the transverse magnetic field, and E and Ez are the transverse and the axial electric fields, respectively. Previous research has investigated the bunching mechanism of TM-mode gyrotrons under the assumption of weakly relativistic electron beam.46 An important finding has been obtained for gyro-BWO, indicating that the azimuthal bunching (through Δγ) and the axial bunching (through Δvz) induced by Ez always cooperate. The cooperation becomes prominent during the far from cutoff operation. On the contrary, the band of current interest is operated near the cutoff, characterized by small kz values. This condition leads to the following special features. First, the axial bunching resulting from the modulation of the electron axial speed Δvz are automatically suppressed due to negligible kz, as evident in Eq. (2). Second, the azimuthal bunching triggered by the transverse electric field E is also suppressed, because E of the TM modes is also proportional to kz. Among all bunching processes, only the longitudinal electric field (Ez) has the capability to significantly modulate the beam dynamics by performing substantial work (vzEz) in Δγ. The azimuthal bunching then emerges as a result of the modulation of the electron's relativistic cyclotron frequency, giving rise to a collective energy exchange between beam and wave. Based on the analyses presented in the previous work46 [referring to its Figs. 6(a) and 7(a)], the center of the azimuthal bunching caused by forward Ez and backward Ez coincide with each other. This result implies that the nearly degenerate forward and backward waves would work cooperatively to modulate the incoming electron beam. The enhanced bunching therefore facilitates the efficient beam–wave energy transfer.
FIG. 6.

Δγ of totally 21 electron ensembles obtained by particle-tracing simulation with Ib = 5 A; (a) for B0 = 36 kG and (b) for B0 = 38 kG.

FIG. 6.

Δγ of totally 21 electron ensembles obtained by particle-tracing simulation with Ib = 5 A; (a) for B0 = 36 kG and (b) for B0 = 38 kG.

Close modal
FIG. 7.

(a) Nonlinear efficiencies vs B0 for the TE01-mode gyrotron. Uniform waveguide geometry: rw = 0.2 cm and L = 3.5 cm. Beam parameters: Vb = 70 kV, Ib = 1 A, and α = 1.0. (b) ηfwd (upper panel) and ηfwd (bottom panel) vs B0 under the 1 cm upstream lossy-section test for the TE01 mode. (c) TE01 starting current Ist (solid line) and frequency ωst (red line, referring to the R.H.S. y-axis) under B0 tuning. 1/Ist is also displayed (blue dotted line). (d) Electron transit-angle phases of the forward wave (Θfwd) and the backward wave (Θbwd) vs B0. Their difference (ΘfwdΘbwd) is plotted in green. The velocity ratio vg/vz0 is shown by the black line, referring to the R.H.S y-axis. The wall loss used for calculation is fixed at 1ρCu.

FIG. 7.

(a) Nonlinear efficiencies vs B0 for the TE01-mode gyrotron. Uniform waveguide geometry: rw = 0.2 cm and L = 3.5 cm. Beam parameters: Vb = 70 kV, Ib = 1 A, and α = 1.0. (b) ηfwd (upper panel) and ηfwd (bottom panel) vs B0 under the 1 cm upstream lossy-section test for the TE01 mode. (c) TE01 starting current Ist (solid line) and frequency ωst (red line, referring to the R.H.S. y-axis) under B0 tuning. 1/Ist is also displayed (blue dotted line). (d) Electron transit-angle phases of the forward wave (Θfwd) and the backward wave (Θbwd) vs B0. Their difference (ΘfwdΘbwd) is plotted in green. The velocity ratio vg/vz0 is shown by the black line, referring to the R.H.S y-axis. The wall loss used for calculation is fixed at 1ρCu.

Close modal

The beam–wave energy transfer can be further examined by the change in Lorentz factors Δγ of all electron ensembles. Figure 6(a) demonstrates Δγ of 21 representative electrons calculated in particle-tracing simulation at near-cutoff operation (B0 = 36 kG), while the similar plot at a relative far from cutoff region (B0 = 38 kG) is shown in Fig. 6(b). In the former case [Fig. 6(a)], it is evident that the electron bunching modulated by the forward and backward waves cooperate with each other, resulting in the enhanced energy extraction from electrons to wave (most Δγ becomes negative when z = 2 ∼ 3.5 cm). As for the far-cutoff gyro-BWOs [Fig. 6(b)], since only the backward wave is present, less energy change in electrons is observed.

The preceding discussions primarily revolve around TM-mode gyrotron oscillators. However, a question still lingers: Does the proposed intrinsic resonance apply to all types of operating modes? To explore this, Sec. VI shifts the focus toward TE-mode gyrotrons.

In order to compare distinctions and similarities between TE-mode and TM-mode gyrotrons, the TE01 mode is chosen due to its degeneracy with the previously discussed TM11 mode.36 The same uniform waveguide with rw = 0.2 cm and L = 3.5 cm with a constant B0 throughout the tube are chosen. The optimum electron guiding center radius rc = 0.48rw is selected for efficiently interacting with the TE01 mode.40,44 The beam current used in the TE01 nonlinear particle-tracing simulation13 is 1 A, while the other beam parameters hold the same (Vb = 70 kV and α = 1). The obtained nonlinear efficiencies are demonstrated in Fig. 7(a). Similar to the TM11 case, a high beam efficiency of 30% is achieved by a 3.5 cm uniform tube without external feedback. Note that similar hill-shaped efficiency spectrum for the TE-mode gyromonotron has also been observed,27 in which a semi-infinite regular tube with a upstream structure cutoff and a downstream magnetic-field cutoff was employed. This high similarity suggests that the two cutoffs adopted27 may not be the primary reasons for feedback.

Based on the spectra of ηfwd and ηbwd in Fig. 7(a), we can deduce that the interaction in the TE-mode gyrotron under near-cutoff operation involves the simultaneous generation of both forward and backward waves. This statement aligns with the conclusion drawn in Sec. II regarding the TM11-mode gyrotron. On the other hand, similar to the approach in Sec. III, two lossy-section tests are conducted on the TE01-mode gyrotron. The impact of introducing a 1 cm upstream lossy section is illustrated in Fig. 7(b). At relatively low-B0 region, the heavy wall loss at upstream terminate the forward and backward oscillations at the same time. This indicates the interconnected nature of these two waves, and their reliance on the internal feedback loop closely resemble the trend observed in Figs. 3(c) and 3(d). The results of the downstream lossy section are also quite similar to those in Fig. 3(a) and thus are not displayed here for conciseness.

Figure 7(c) shows the starting current Ist and oscillation frequency ωst of the TE01 mode, calculated by the method proposed in the literature.47, Ist of the TE01 mode is three times smaller than that of the TM11 mode [Fig. 4(a)] owing to the larger optimum beam–wave coupling strength.36,40 Notice that 1/Ist exhibits a striking resemblance to the trend of ηfwd shown in Fig. 7(a), corresponding to the strong beam–wave interaction in the gyromonotron regime. A similar valley-shaped Ist spectrum has also been demonstrated before,19,30 however, within the structures of nonuniformities. We can therefore conclude that the observed strong beam–wave interaction under near-cutoff operation of gyromonotrons should originate from the proposed “intrinsic resonance” rather than the traditional cavity resonance induced by external feedbacks. Figure 7(d) demonstrates the electron transit angles of the TE01 mode at the gyromonotron region. Once again, Θfwd and Θbwd align with each other, with a minor deviation from Θ0. The nearly in-phase condition promises the constructive interferences of the forward and backward waves, giving rise to the intrinsic resonance nature of the TE01 mode when B0 ∼ 36 kG.

As for the bunching mechanism at the near-cutoff, again, due to negligible kz, the azimuthal bunching induced by Δγ should dominate as seen from Eq. (2). In other words, E of the TE01 mode governs the energy-exchange process, which shows no difference when the wave's propagation direction is reverse. This implies that the dominated azimuthal bunching effects caused by the forward E and the backward E always cooperate with each other,1 responsible for the high interaction efficiency observed at the near-cutoff region for TE modes.

As the last part of the discussion, we attempt to identify the type of instability dominating the forward oscillation in the near cutoff region [Figs. 1(d) and 7(a)]. The general mapping analyses39 on the complex-kz plane, also known as Briggs–Bers criterion (BBC),48,49 for both the TM11 and TE01 modes are conducted. The result reveals that the current phenomena, i.e., the obtained strong forward wave, correspond to the convective instability rather than the absolute instability. While this finding provides an explanation for the origin of the “convectively growing” forward wave observed in Figs. 2(a) and 4(b), it appears to be in contradiction with the arguments of the “effective internal feedback loop” proposed in Secs. IV and VI. This apparent contradiction may arise from the differing frames of reference. From the waveguide frame's perspective, the oscillation is indeed dominated by the convective instability, manifesting in the growing forward wave. However, the condition with negligible group velocity at the cutoff promises that the forward wave would approach the electrons backwardly, building up the aforementioned internal feedback loop and the “effective absolute instability” for electrons. In addition, the BBC39 for absolute instability is established based on the assumption of infinitely long interaction length, while the classification of the absolute/convective instabilities for a finite-length system is not well-defined.16 Rigorous definition of instability criteria is, however, beyond the scope of the present work. These questions should be carefully addressed in the follow-up studies. Noteworthily, the absolute instability near the zero group velocity points, similar to the condition in the present study, has recently been carefully re-examined using the BBC, for the conventional traveling-wave tube cases.50,51

On the other hand, it is important to clarify similarities and differences between the findings presented in this paper and the predictions of classical theories linked to the backward-wave oscillator (BWO) and the traveling-wave tube (TWT). During the linear and start-oscillation stages, as illustrated in Figs. 4(b) and 4(c), the forward-wave component undergoes exponential growth as it propagates, eventually forming a slowly varying spatial plateaus. When operating within the nonlinear regime, if the tube length exceeds a specific threshold, the RF field may redistribute some of its power back to the beam in the steady state, as evident in Fig. 2(a)—a phenomenon akin to the well-known behavior in TWT.52 

Owing to the imposed outgoing-wave boundary conditions, the backward-wave component, in both gyromonotron and gyro-BWO regimes, maintains a zero value at the downstream end (z = L) and grows backwardly to a substantial value at the upstream end (z = 0). These patterns can be observed in both the linear start-oscillation case [Fig. 4(c)] and the nonlinear steady state [Fig. 2(b)], perfectly aligning with predictions from the classical BWO theory of Johnson.53 Additionally, the dip in the starting-current curve [Fig. 4(a) for the TM11 mode and Fig. 7(c) for the TE01 mode] can also be rationalized through the classical BWO theory. While the BBC theory for an infinitely long tube predicts a starting current of zero at the cutoff point, in a finite-length system, there must be a finite, non-zero starting current beyond which oscillations commence. Hence, a minimum starting current should occur near the cutoff for a finite-length system.

Furthermore, in the realm of a gyrotron traveling-wave tube (gyro-TWT), it typically operates below the grazing condition to achieve a steady output with high gain by inducing convective instability.1,13,38,39,42 However, absolute instability may still be triggered as the beam current surpasses a certain threshold.43 Lau et al. conducted a comprehensive study on the oscillation mechanism in a gyro-TWT system.54 The overdrive of the beam current is believed to induce self-excitation of the “backward waveguide mode” due to the internal feedback mechanism by extending the unstable spectrum to negative values of kz. This statement aligns with BBC established in the infinitely long system, suggesting that only the backward wave may result from absolute instability, while the forward wave relies on convective instability consistent with our findings using BBC. However, the discoveries in this work reveal that forward oscillation can also occur based on “effective” absolute instability, manifesting in the intrinsic resonance constructed by strongly coupled forward and backward waves in the finite-length system. This clarifies the question raised by Lau et al. regarding the existence of absolute instability when both beam and waveguide modes propagate in the forward direction. Even when the current is not large enough to strongly couple the beam mode with the backward waveguide mode, a steady and strong forward wave oscillation can still be observed. This must be carefully avoided when optimizing gyro-TWT devices because the self-excited forward and backward oscillations might have the potential to significantly spoil the desired functionality.

In this work, we conduct a comprehensive investigation of the TM11-mode and TE01-mode gyrotrons, with a particular focus on a uniform tube without external structural feedbacks. The nonlinear particle-tracing simulation yields an interesting finding: the uniform tube can function as gyromonotron, in the absence of structural and magnetic-field nonuniformities, with a beam efficiency exceeding 30%. By decomposing the fields, we demonstrate that the forward and backward waves are co-generated at this region. A series of lossy-section tests are further employed to test the statement. The upstream wall loss is found to terminate both the backward and forward oscillation, despite the significant deviation of the heavy-loss location from the main steady-state forward field profile. Additionally, a linear analysis conducted during the start-oscillation stage reveals that the oscillation frequency is closely locked to the cutoff frequency of the mode. Operating near the cutoff opens up a possibility: the forward and backward waves generated by preceding electrons may both modulate the newly injected beam, and the internal feedback loop of the forward wave is thus established. From the electron's perspective, the forward and backward waves are nearly degenerate. These in-phase waves coherently form a standing-wave pattern in the steady state, effectively transforming the uniform tube into a resonant cavity, which explains the high interaction efficiency observed at the near-cutoff region.

The findings presented in this study reshape the previous understanding of the open-cavity-based gyromonotrons. Contrary to previous beliefs, it is now evident that the commonly used upstream cutoff end, taper reflection ends, or downstream magnetic-field cutoff only serves as a secondary enhancement rather than the primary driving force of oscillations. Despite the potential challenges in wave coupling that may arise due to near-cutoff operation, the insights provided here are believed to shed light on the underlying mechanism for existing gyrotron devices. The current trend in gyrotron design is shifting toward the development of reflective-type gyro-backward wave oscillators (R-gyro-BWOs),20,55–57 which combine the high efficiency of the gyromonotron with the advantage of continuous frequency tunability found in gyro-BWOs. However, the underlying mechanism behind the operation of R-gyro-BWOs remains unclear. The interplay between the forward and backward waves discussed in the present work is believed to offer more clues into its working principle.

Last but not least, it is worth noting that while most gyrotron devices have predominantly utilized the TE mode,1,2,7,8 there has been limited exploration36,37,40,41 into the TM-mode gyrotrons due to the significantly higher start-oscillation current required, stemming from their weaker mode-coupling strength in a classical open-cavity-type gyrotron adopting weakly relativistic electron beams.58 This might lead some to regard the comparative studies between TE and TM modes in the current study as purely theoretical endeavors. However, a recent study by Yang et al. has demonstrated the feasibility of a TM12-mode gyrotron oscillator operating in the sub-terahertz band (∼220 GHz).59 The importance of considering a TM mode has been emphasized, especially in the context of harmonic TE-mode gyrotrons using the axis-encircling electron beam, because the mode-coupling strength of a s + 1-harmonic TE mode and that of a s-harmonic TM mode are comparable. Since the competitiveness of TM modes has been confirmed by the multi-mode simulations using particle-in-cell simulator (PIC) solver of computer simulation technology (CST) Studio, the potential of TM-mode gyrotrons warrants further theoretical exploration. This article should help to partially elucidate the similarities and differences between TM-mode and TE-mode gyrotrons, particularly within the near-cutoff gyromonotron regions.

The authors would like to thank Professor Kwo-Ray Chu for valuable comments and suggestions. This work was supported by the National Science and Technology Council, Taiwan, under Contract Nos. NSTC 111-2112-M-194-009-MY3, NSTC 112-2112-M-194-006-MY3, and NSTC 110-2112-M-007-013-MY3.

The authors have no conflicts to disclose.

Tien-Fu Yang: Conceptualization (equal); Formal analysis (equal); Investigation (lead); Methodology (equal); Software (equal); Validation (supporting); Writing—original draft (lead); Writing—review & editing (equal). Hsin-Yu Yao: Conceptualization (equal); Formal analysis (equal); Methodology (equal); Project administration (equal); Software (equal); Supervision (equal); Validation (lead); Writing—review & editing (equal). Shih-Hung Chen: Conceptualization (supporting); Formal analysis (supporting); Validation (supporting); Writing—review & editing (supporting). Tsun-Hsu Chang: Conceptualization (equal); Funding acquisition (lead); Project administration (equal); Supervision (equal); Writing—review & editing (supporting).

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

The particle-tracing process is a self-consistent, frequency-domain, and single-mode method designed to evaluate the frequency and field profile of self-oscillation in a steady state caused by the beam–wave interaction. In this framework, the generating field Ez of a right-hand circularly polarized TMmn mode is
(A1)
where f(z) is the complex field profile, ω is the angular frequency, c is the speed of light in vacuum, and Jm is the Bessel function of the order m. kmn=xmn/rw, where xmn is the n-th root of Jm, satisfying Jm(xmn)=0. kz satisfies the formula of cold-tube TM mode, which can be expressed as
(A2)
where δ=2/μcωσ denotes the skin depth, μc is the permeability of the metal, and σ is the conductivity. The governing nonlinear wave equation is derived from Maxwell's equations, yielding
(A3)
where Ib represents the electron beam current, Kmn=Jm2(xmn), Csm1/2=Jsm(kmnrc,j)Js(kmnrL,j) is the TM-mode field coupling strength, Asm,j=ωtjsϕj(sm)ψj+(ms/2)π, and s is the harmonic order. The prime and asterisk notations, respectively, represent the first-order derivative of the variable with respect to z and the complex conjugate. The subscript j denotes the quantities of the j-th electron, and the corresponding variables Wj (momentum weighting factor), p,j (transverse momentum), pz,j (axial momentum), rc,j (guiding center radius), rL,j (Larmor radius), ϕj, and ψj are graphically defined in Fig. 1(a) of the work.37 
On the other hand, the dynamic differential equations for electrons are derived from the equation of motion dp/dt=eEtote(p×Btot)/(γmec) as follows:
(A4)
(A5)
(A6)
(A7)
(A8)
where Ωe=eB0/mec (electron cyclotron frequency), B0 is the applied magnetic field for gyromotion, k0=ω/c, Hsm1/2=Jsm(kmnrc)Js(kmnrL), and Gsm1/2=Jsm(kmnrc)Js(kmnrL).

Equations (A1)–(A8) can be solved together. Beginning from the beam entrance (z=0), the first outgoing-wave wave boundary condition f(z=0)=ikzf(z=0) is specified for ensuring purely backward wave, where f(z=0) is still an unknown factor. The coupled differential equations are then integrated axially from z=0 to z=L. At the end of the system (z=L), the second outgoing-wave boundary condition f(z=L)=ikzf(z=L) is imposed for ensuring a purely forward wave. This complex equation can be solved by an iterative procedure for obtaining the purely real oscillation frequency ω and initial field amplitude f(z=0). Based on these two parameters, the complex field amplitude f(z) and all other electron's dynamic parameters are obtained. The iterative root searching processes are completed if the satisfaction of boundary conditions achieves the level of 10−10, guaranteeing negligible reflections from the boundaries. It is worth noting that a similar set of equations for TE modes can be found in Ref. 13, which were utilized to compute the results presented in Figs. 7(a) and 7(b).

Based on the obtained f(z) and f(z), the net local power Pnet(z) can be determined by
(A9)
Since all solutions are required to strictly satisfy the outgoing-wave boundary conditions, the ratio of Pnet(z=0) to input beam power (Pin=Ib×Vb) directly determines the backward-wave efficiency ηbwd, whereas the ratio of Pnet(z=L) to Pin determines the forward-wave efficiency ηfwd. Further decomposition of the complex field profile by f(z)=ffwdeikzz+fbwdeikzz helps to extract the detailed forward and backward field coefficients, i.e., ffwd and fbwd, respectively. By assuming that the ffwd and fbwd are weekly dependent on z compared to the phasor e±ikzz, they can be simply expressed as
(A10)
Note that Eq. (A10) is applied not only in the nonlinear and self-consistent model but also in the linear start-oscillation analysis. The above-mentioned assumption holds perfectly for a uniform tube with a uniform distribution of wall loss, which is the primary case under consideration in this study. However, it is important to emphasize that this approximated formula may not be suitable for cases involving sharp boundaries of wall loss, as considered in Figs. 3 and 7(b).
1.
K. R.
Chu
, “
The electron cyclotron maser
,”
Rev. Mod. Phys.
76
(
2
),
489
(
2004
).
2.
G. S.
Nusinovich
,
Introduction to the Physics of Gyrotrons
(
JHU Press
,
2004
).
3.
G. S.
Nusinovich
,
M. K.
Thumm
, and
M. I.
Petelin
, “
The gyrotron at 50: Historical overview
,”
J. Infrared, Millimeter, Terahertz Waves
35
,
325
(
2014
).
4.
V. L.
Bratman
,
Y. K.
Kalynov
, and
V. N.
Manuilov
, “
Large-orbit gyrotron operation in the terahertz frequency range
,”
Phys. Rev. Lett.
102
(
24
),
245101
(
2009
).
5.
M. Y.
Glyavin
,
A. G.
Luchinin
, and
G. Y.
Golubiatnikov
, “
Generation of 1.5-kW, 1-THz coherent radiation from a gyrotron with a pulsed magnetic field
,”
Phys. Rev. Lett.
100
(
1
),
015101
(
2008
).
6.
Y. K.
Kalynov
,
V. N.
Manuilov
,
A. S.
Fiks
, and
N. A.
Zavolskiy
, “
Powerful continuous-wave sub-terahertz electron maser operating at the 3rd cyclotron harmonic
,”
Appl. Phys. Lett
114
(
21
),
213502
(
2019
).
7.
S. P.
Sabchevski
,
M. Y.
Glyavin
, and
G. S.
Nusinovich
, “
The progress in the studies of mode interaction in gyrotrons
,”
J. Infrared, Millimeter, Terahertz Waves
43
(
1–2
),
1
(
2022
).
8.
M.
Thumm
, “
State-of-the-art of high-power gyro-devices and free electron masers
,”
J. Infrared, Millimeter, Terahertz Waves
41
(
1
),
1
(
2020
).
9.
T.
Idehara
,
S. P.
Sabchevski
,
M.
Glyavin
, and
S.
Mitsudo
, “
The gyrotrons as promising radiation sources for THz sensing and imaging
,”
Appl. Sci
10
(
3
),
980
(
2020
).
10.
S.
Sabchevski
,
M.
Glyavin
,
S.
Mitsudo
,
Y.
Tatematsu
, and
T.
Idehara
, “
Novel and emerging applications of the gyrotrons worldwide: Current status and prospects
,”
J. Infrared, Millimeter, Terahertz Waves
42
(
7
),
715
(
2021
).
11.
M. K. A.
Thumm
,
G. G.
Denisov
,
K.
Sakamoto
, and
M. Q.
Tran
, “
High-power gyrotrons for electron cyclotron heating and current drive
,”
Nucl. Fusion
59
(
7
),
073001
(
2019
).
12.
M. Q.
Tran
,
P.
Agostinetti
,
G.
Aiello
,
K.
Avramidis
,
B.
Baiocchi
,
M.
Barbisan
,
V.
Bobkov
,
S.
Briefi
,
A.
Bruschi
,
R.
Chavan
et al, “
Status and future development of heating and current drive for the EU DEMO
,”
Fusion Eng. Des.
180
,
113159
(
2022
).
13.
K. R.
Chu
,
H.-Y.
Chen
,
C.-L.
Hung
,
T.-H.
Chang
,
L. R.
Barnett
,
S.-H.
Chen
,
T.-T.
Yang
, and
D. J.
Dialetis
, “
Theory and experiment of ultrahigh-gain gyrotron traveling wave amplifier
,”
IEEE Trans. Plasma Sci.
27
(
2
),
391
(
1999
).
14.
S. H.
Chen
,
T. H.
Chang
,
K. F.
Pao
,
C. T.
Fan
, and
K. R.
Chu
, “
Linear and time-dependent behavior of the gyrotron backward-wave oscillator
,”
Phys. Rev. Lett.
89
(
26
),
268303
(
2002
).
15.
S. H.
Chen
,
K. R.
Chu
, and
T. H.
Chang
, “
Saturated behavior of the gyrotron backward-wave oscillator
,”
Phys. Rev. Lett.
85
(
12
),
2633
(
2000
).
16.
S.-H.
Chen
and
L.
Chen
, “
Linear and nonlinear behaviors of gyrotron backward wave oscillators
,”
Phys. Plasmas
19
(
2
),
023116
(
2012
).
17.
T. H.
Chang
,
S. H.
Chen
,
L. R.
Barnett
, and
K. R.
Chu
, “
Characterization of stationary and nonstationary behavior in gyrotron oscillators
,”
Phys. Rev. Lett.
87
(
6
),
064802
(
2001
).
18.
T. H.
Chang
,
K. F.
Pao
,
S. H.
Chen
, and
K. R.
Chu
, “
Self-consistent effects on the starting current of gyrotron oscillators
,”
Int. J. Infrared Millimeter Waves
24
,
1415
(
2003
).
19.
A. W.
Fliflet
,
M. E.
Read
,
K. R.
Chu
, and
R.
Seeley
, “
A self-consistent field theory for gyrotron oscillators: Application to a low Q gyromonotron
,”
Int. J. Electron.
53
(
6
),
505
(
1982
).
20.
T. H.
Chang
,
T.
Idehara
,
I.
Ogawa
,
L.
Agusu
, and
S.
Kobayashi
, “
Frequency tunable gyrotron using backward-wave components
,”
J. Appl. Phys.
105
(
6
),
063304
(
2009
).
21.
J.
Huang
,
T.
Song
,
C.
Zhang
,
W.
Wang
,
K.
Zhang
,
M.
Hu
,
Z.
Wu
,
R.
Zhong
,
J.
Zhou
,
T.
Zhao
et al, “
Theoretical investigations on forward-wave and backward-wave operation of a frequency-tunable gyrotron
,”
IEEE Trans. Electron Devices
67
(
9
),
3809
(
2020
).
22.
S.
Sabchevski
and
T.
Idehara
, “
Resonant cavities for frequency tunable gyrotrons
,”
Int. J. Infrared, Millimeter Waves
29
(
1
),
1
22
(
2008
).
23.
V. I.
Shcherbinin
,
T. I.
Tkachova
, and
V. I.
Tkachenko
, “
Improved cavity for broadband frequency-tunable gyrotron
,”
IEEE Trans. Electron Devices
65
(
1
),
257
(
2018
).
24.
S. H.
Kao
,
C. C.
Chiu
,
P. C.
Chang
,
K. L.
Wu
, and
K. R.
Chu
, “
Harmonic mode competition in a terahertz gyrotron backward-wave oscillator
,”
Phys. Plasmas
19
(
10
),
103103
(
2012
).
25.
K. A.
Avramidis
,
A.
Marek
,
I.
Chelis
,
Z. C.
Ioannidis
,
L.
Feuerstein
,
J.
Jelonnek
,
M.
Thumm
, and
I.
Tigelis
, “
Simulation of parasitic backward-wave excitation in high-power gyrotron cavities
,”
IEEE Trans. Electron Devices
70
(
4
),
1898
(
2023
).
26.
Z. C.
Ioannidis
,
K. A.
Avramidis
,
T.
Rzesnicki
,
I.
Chelis
,
G.
Gantenbein
,
S.
Illy
,
J.
Jin
,
I. G.
Pagonakis
,
M.
Thumm
, and
J.
Jelonnek
, “
Generation of 1.5 MW–140 GHz pulses with the modular pre-prototype gyrotron for W7-X
,”
IEEE Electron Device Lett.
42
(
6
),
939
(
2021
).
27.
V. L.
Bratman
,
M. A.
Moiseev
,
M. I.
Petelin
, and
R. É.
Érm
, “
Theory of gyrotrons with a nonfixed structure of the high-frequency field
,”
Radiophys. Quantum Electron.
16
(
4
),
474
(
1973
).
28.
V. L.
Bratman
, “
The starting regime for an MCR-monotron with a cavity having a low diffraction Q
,”
Radiophys. Quantum Electron.
17
(
10
),
1181
(
1974
).
29.
E.
Borie
, “
Self consistent code for a 150 GHz gyrotron
,”
Int. J. Infrared Millimeter Waves
7
(
12
),
1863
(
1986
).
30.
V. L.
Bratman
and
M. A.
Moiseev
, “
Conditions for self-excitation of a cyclotron resonance maser with a nonresonant electrodynamic system
,”
Radiophys. Quantum Electron.
18
(
7
),
772
(
1975
).
31.
V. L.
Bratman
,
A. V.
Savilov
, and
T. H.
Chang
, “
Possibilities for continuous frequency tuning in terahertz gyrotrons with nontunable electrodynamic systems
,”
Radiophys. Quantum Electron.
58
(
9
),
660
(
2016
).
32.
L.
Hongfu
,
D.
Pinzhong
, and
X.
Zhonglin
, “
Self-consistent field theory and calculation for gyromonotron
,”
Int. J. Infrared Millimeter Waves
9
(
2
),
135
(
1988
).
33.
H.
Saito
,
K.
Kreischer
,
B. G.
Danly
,
T. M.
Tran
, and
R. J.
Temkin
, “
A gyrotron with a minimum Q cavity
,”
Int. J. Electron.
61
(
6
),
757
(
1986
).
34.
V. E.
Zapevalov
,
Y. K.
Kalynov
,
A. N.
Kuftin
,
S. A.
Malygin
, and
E. M.
Tai
, “
Low-Q cavities for high-power gyrotrons
,”
Radiophys. Quantum Electron.
37
(
3
),
233
(
1994
).
35.
C.-L.
Hung
,
Y.-S.
Yeh
, and
T.-H.
Chang
, “
Linear and nonlinear characteristics of a low-voltage gyrotron backward-wave oscillator in the subterahertz range
,”
IEEE Trans. Plasma Sci.
49
(
5
),
1557
(
2021
).
36.
H.-Y.
Yao
,
C.-C.
Chen
, and
T.-H.
Chang
, “
Starting behaviors of the TM-mode gyrotrons
,”
Phys. Plasmas
27
(
2
),
022113
(
2020
).
37.
H.-Y.
Yao
,
C.-H.
Wei
, and
T.-H.
Chang
, “
Nonlinear and self-consistent single-mode formulation for TM-mode gyrotrons
,”
Phys. Rev. E
104
(
6
),
065205
(
2021
).
38.
K. R.
Chu
,
H. Y.
Chen
,
C. L.
Hung
,
T. H.
Chang
,
L. R.
Barnett
,
S. H.
Chen
, and
T. T.
Yang
, “
Ultrahigh gain gyrotron traveling wave amplifier
,”
Phys. Rev. Lett.
81
(
21
),
4760
(
1998
).
39.
K. R.
Chu
and
A. T.
Lin
, “
Gain and bandwidth of the gyro-TWT and CARM amplifiers
,”
IEEE Trans. Plasma Sci.
16
(
2
),
90
(
1988
).
40.
T.-H.
Chang
,
H.-Y.
Yao
,
B.-Y.
Su
,
W.-C.
Huang
, and
B.-Y.
Wei
, “
Nonlinear oscillations of TM-mode gyrotrons
,”
Phys. Plasmas
24
(
12
),
122109
(
2017
).
41.
T.-H.
Chang
and
K.-J.
Xu
, “
Gain and bandwidth of the TM-mode gyrotron amplifiers
,”
Phys. Plasmas
25
(
11
),
112109
(
2018
).
42.
K. R.
Chu
,
L. R.
Barnett
,
H. Y.
Chen
,
S. H.
Chen
,
C.
Wang
,
Y. S.
Yeh
,
Y. C.
Tsai
,
T. T.
Yang
, and
T. Y.
Dawn
, “
Stabilization of absolute instabilities in the gyrotron traveling wave amplifier
,”
Phys. Rev. Lett.
74
(
7
),
1103
(
1995
).
43.
T. H.
Chang
and
N. C.
Chen
, “
Transition of absolute instability from global to local modes in a gyrotron traveling-wave amplifier
,”
Phys. Rev. E
74
(
1
),
016402
(
2006
).
44.
T. H.
Chang
,
C. F.
Yu
,
C. L.
Hung
,
Y. S.
Yeh
,
M. C.
Hsiao
, and
Y. Y.
Shin
, “
W-band TE01 gyrotron backward-wave oscillator with distributed loss
,”
Phys. Plasmas
15
(
7
),
073105
(
2008
).
45.
S. H.
Kao
,
C. C.
Chiu
, and
K. R.
Chu
, “
A study of sub-terahertz and terahertz gyrotron oscillators
,”
Phys. Plasmas
19
(
2
),
023112
(
2012
).
46.
T. H.
Chang
,
W. C.
Huang
,
H. Y.
Yao
,
C. L.
Hung
,
W. C.
Chen
, and
B. Y.
Su
, “
Asymmetric linear efficiency and bunching mechanisms of TM modes for electron cyclotron maser
,”
Phys. Plasmas
24
(
2
),
023302
(
2017
).
47.
C. S.
Kou
,
Q. S.
Wang
,
D. B.
McDermott
,
A. T.
Lin
,
K. R.
Chin
, and
N. C.
Luhmann
, “
High-power harmonic gyro-TWT's. I. Linear theory and oscillation study
,”
IEEE Trans. Plasma Sci.
20
(
3
),
155
(
1992
).
48.
R. J.
Briggs
,
Electron-Stream Interaction with Plasmas
(
MIT Press
,
Cambridge, MA
,
1964
), Vol.
187
.
49.
A.
Bers
,
M.
Rosenbluth
, and
R.
Sagdeev
,
Handbook of Plasma Physics
, edited by
M. N.
Rosenbluth
and
R. Z.
Sagdeev
(
North-Holland
,
Amsterdam
,
1983
), Vol.
1
, chap. 3.2.
50.
D.
Hung
,
I.
Rittersdorf
,
P.
Zhang
,
D.
Chernin
,
Y.
Lau
,
T.
Antonsen
, Jr.
,
J.
Luginsland
,
D.
Simon
, and
R.
Gilgenbach
, “
Absolute instability near the band edge of traveling-wave amplifiers
,”
Phys. Rev. Lett.
115
(
12
),
124801
(
2015
).
51.
F.
Antoulinakis
,
P.
Wong
,
A.
Jassem
, and
Y. Y.
Lau
, “
Absolute instability and transient growth near the band edges of a traveling wave tube
,”
Phys. Plasmas
25
(
7
),
072102
(
2018
).
52.
A. P.
Kuznetsov
,
S. P.
Kuznetsov
,
A. G.
Rozhnev
,
E. V.
Blokhina
, and
L. V.
Bulgakova
, “
Wave theory of a traveling-wave tube operated near the cutoff
,”
Radiophys. Quantum Electron.
47
(
5
),
356
(
2004
).
53.
H. R.
Johnson
, “
Backward-wave oscillators
,”
Proc. IRE
43
(
6
),
684
(
1955
).
54.
Y. Y.
Lau
,
K. R.
Chu
,
L. R.
Barnett
, and
V. L.
Granatstein
, “
Gyrotron travelling wave amplifier: I. Analysis of oscillations
,”
Int. J. Infrared Millimeter Waves
2
(
3
),
373
(
1981
).
55.
N.-C.
Chen
,
T.-H.
Chang
,
C.-P.
Yuan
,
T.
Idehara
, and
I.
Ogawa
, “
Theoretical investigation of a high efficiency and broadband subterahertz gyrotron
,”
Appl. Phys. Lett.
96
(
16
),
161501
(
2010
).
56.
C. H.
Tsai
,
T. H.
Chang
,
Y.
Tatematsu
,
Y.
Yamaguchi
,
M.
Fukunari
,
T.
Saito
, and
T.
Idehara
, “
Reflective gyrotron backward-wave oscillator with piecewise frequency tunability
,”
IEEE Trans. Electron Devices
68
(
1
),
324
(
2021
).
57.
F. H.
Li
,
C. H.
Du
,
Z. W.
Zhang
,
S. Q.
Li
,
Z. C.
Gao
,
P. K.
Liu
,
G. W.
Ma
,
Q. L.
Huang
,
H. G.
Ma
,
L.
Zhang
et al, “
Design of a 1-THz fourth-harmonic gyrotron driven by axis-encircling electron beam
,”
IEEE Trans. Electron Devices
69
(
6
),
3393
(
2022
).
58.
É. B.
Abubakirov
, “
Excitation of transverse magnetic modes and mode selection in relativistic cyclotron-resonance masers
,”
Radiophys. Quantum Electron.
26
(
4
),
379
(
1983
).
59.
T. F.
Yang
,
H. Y.
Yao
,
C. C.
Chang
,
C. H.
Du
,
F.
Zhang
, and
T. H.
Chang
, “
TE- and TM-Mode competition in subterahertz gyrotron using axis-encircling electron beam
,”
IEEE Trans. Electron Devices
71
(
1
),
815
(
2024
).