Coupled-cavity traveling wave tube (CCTWT) is a high power microwave vacuum electronic device used to amplify radio frequency signals. CCTWTs have numerous applications, including radar, radio navigation, space communication, television, radio repeaters, and charged particle accelerators. Microwave-generating interactions in CCTWTs take place mostly in coupled resonant cavities positioned periodically along the electron beam axis. Operational features of a CCTWT, particularly the amplification mechanism, are similar to those of a multicavity klystron. We advance here a Lagrangian field theory of CCTWTs with the space being represented by one-dimensional continuum. The theory integrates into it the space-charge effects, including the so-called debunching (electron-to-electron repulsion). The corresponding Euler–Lagrange field equations are ordinary differential equations with coefficients varying periodically in the space. Utilizing the system periodicity, we develop instrumental features of the Floquet theory, including the monodromy matrix and its Floquet multipliers. We use them to derive closed form expressions for a number of physically significant quantities. Those include, in particular, dispersion relations and the frequency dependent gain foundational to the RF signal amplification. Serpentine (folded, corrugated) traveling wave tubes are very similar to CCTWTs, and our theory applies to them also.
I. INTRODUCTION
We start with general principles of microwave radiation generation and amplification of RF signals (see Ref. 1, Chap. 4): “ANY generating or amplifying device converts d.c. energy into high-frequency electric field energy, and this conversion is affected by means of an electron beam. All energy exchanges between the electron beam and the alternating electric field are a result of acceleration or retardation of electrons. The kinetic energy of electrons is converted into electromagnetic energy, and vice versa. Therefore, although the mechanisms of various devices are different, in each of them, power is transferred from the constant voltage source to the alternating electromagnetic field. This is brought about in the oscillatory system by means of a density-modulated electron beam in which electrons are accelerated in the constant electric field and retarded in the alternating electric field. Density modulation of the electron beam makes it possible to retard a greater number of electrons than are accelerated by the same alternating field, thus producing the transfer of energy.”
The last sentence in the above quote underlines the critical role played by the density modulation of the electron beam (known also as “electron bunching”) in the energy transfer from the electron beam to the electromagnetic (EM) radiation.
A coupled-cavity traveling wave tube (CCTWT) shown schematically in Fig. 1 is the primary subject we pursue here. The CCTWT is a special type of traveling wave tube (TWT) that utilizes the coupled-cavity structure (CCS)8 as a slow-wave structure (SWS) (see Ref. 2, Chap. 15) (see Ref. 3, Sec. 4). The CCS commonly is a periodic linear chain of several tens of cavities coupled by coupling holes or slots and a beam tunnel (see Ref. 4, 8.7.5). The cavities can be similar to those in klystrons. As to the physical implementation, cavities are often constructed of sections of a slow-wave structure that are made resonant by suitable terminations. The quality factor of each cavity is required to be sufficiently high so that the RF field distribution in each separate cavity is substantially unaffected by the interaction with the beam.5
Schematic representation of a coupled-cavity traveling wave tube (CCTWT) composed of a periodic array of coupled cavities (often of toroidal shape) interacting with the pencil-like electron beam. The interaction causes electron bunching and consequent amplification of the RF signal.
Schematic representation of a coupled-cavity traveling wave tube (CCTWT) composed of a periodic array of coupled cavities (often of toroidal shape) interacting with the pencil-like electron beam. The interaction causes electron bunching and consequent amplification of the RF signal.
By its very design, the CCS is mechanically and thermally more robust than a helix, which is often used as the SWS, allowing for much greater average power, especially in short-wave bands of microwave range6,7 (see Ref. 4, Secs. 8.3, 8.7.5). Serpentine (folded, corrugated) TWTs are based on the corresponding waveguides that permit for electron interaction below the velocity of light (see Ref. 2, Secs. 15.1, 15.2). They are very similar to CCTWTs, and the results of our studies also apply to serpentine (folded, corrugated) TWTs.
A distinct and important feature of a CCTWT is that the interaction between the e-beam and the coupled-cavity structure (CCS), particularly the electron velocity modulation by a high-frequency EM field, is limited mostly to EM cavities positioned periodically along the e-beam axis. The cavity properties that are significant for an effective interaction with the e-beam are as follows (see Ref. 1, Sec. 2): “In order to be used in an electron tube, a cavity resonator must have a region with a relatively strong high-frequency field, which is polarized along the direction of electron flow. This region should, in the majority of cases, be so small that the electron transit time is less than the period of change of the field. Hollow toroidal resonators satisfy these conditions. Toroidal resonators consist of cylinders with a very prominent “bulge” in the middle.
The primary subject of our studies here is the construction of one-dimensional Lagrangian field theory of a coupled-cavity traveling wave tube (CCTWT), a schematic sketch of which is shown in Fig. 1 [compare it with Fig. 11 with a schematic sketch of a multicavity klystron (MCK)]. This theory integrates into it (i) our one-dimensional Lagrangian field theory for TWTs introduced and studied in Ref. 10, Chap. 4, 24 and reviewed in Sec. II; (ii) one-dimensional Lagrangian field theory for the multicavity klystron we developed in Ref. 11 and reviewed in Sec. VIII. The theory takes into account the space-charge effects, and it applies also to serpentine (folded, corrugated) traveling wave tubes.
This paper is organized as follows. In Sec. II, we review concisely the one-dimensional Lagrangian field theory for TWTs introduced and studied in Ref. 10, Chap. 4, 24. In Sec. III, we construct the Lagrangian of the CCTWT, derive the corresponding Euler–Lagrange equations, and introduce CCTWT constitutive subsystems: coupled cavity structure and the e-beam. In Sec. IV, we use the Floquet theory to study solutions to Euler–Lagrange equations. In particular, we construct the monodromy matrix. In Sec. V, we analyze Floquet multiplies, which are solutions to characteristic equations. In Sec. VI, we construct dispersion relations and study their properties. In Sec. VII, we derive expressions for the frequency dependent gain associated with CCTWT eigenmodes. In Sec. VIII, we review concisely the one-dimensional Lagrangian field theory for multicavity klystrons developed in Ref. 11 that allows us to see some of its features in the CCTWT. In Sec. IX, we study the couple-cavity structure when it is not coupled to the e-beam. That allows us to see some of its features in properties of the CCTWT. In Appendixes A–I, we review for the reader’s convenience a number of mathematical and physical subjects relevant to the analysis of the CCTWT. The kinetic and field points of view on the gap interaction are considered in Sec. X. The Lagrangian variational framework of our analytical theory is developed in Sec. XI. In Sec. XII, we consider special polynomials of the fourth degree and their root degeneracies that are useful for our studies of CCTWT exceptional points of degeneracy. In Appendixes A–I, we review some mathematical and physical subject relevant to our studies.
While quoting monographs, we identify the relevant sections as follows: Reference [X,Y] refers to Section/Chapter “Y” of monograph (article) “X”, whereas [X, p. Y] refers to page “Y” of monograph (article) “X”. For instance, Ref. 12, VI.3 refers to monograph,12 Section VI.3; Ref. 12, p. 131 refers to page 131 of monograph.12
II. SKETCH OF THE ANALYTIC MODEL OF THE TRAVELING WAVE TUBE
When constructing the CCTWT Lagrangian, we use the elements of the analytic theory of TWTs developed in Refs. 10 and 13. The purpose of this section is to introduce those elements and the relevant variables of the analytic model of TWT. TWT converts the energy of the electron beam (e-beam) into the EM energy of the amplified RF signal. A schematic sketch of typical TWT is shown in Fig. 2. To facilitate the energy conversion and signal amplification, the e-beam is enclosed in the so-called slow wave structure (SWS), which supports waves that are slow enough to effectively interact with the electron flow. As a result of this interaction, the kinetic energy of electrons is converted into the EM energy stored in the field2,4 (see Ref. 14, Sec. 2.2, Ref. 15, Chap. 4). Consequently, the key operational principle of a TWT is a positive feedback interaction between the slow-wave structure and the flow of electrons. The physical mechanism of the radiation generation and its amplification is electron bunching caused by the acceleration and deceleration of electrons along the e-beam (see quotes in Sec. I).
The upper picture is a schematic representation of a traveling wave tube. The lower picture shows an RF perturbation in the form of a space-charge wave, which is amplified exponentially as it propagates through the traveling wave tube.
The upper picture is a schematic representation of a traveling wave tube. The lower picture shows an RF perturbation in the form of a space-charge wave, which is amplified exponentially as it propagates through the traveling wave tube.
A typical TWT consists of a vacuum tube containing the e-beam that passes down the middle of an SWS, such as an RF circuit. It operates as follows. The left end of the RF circuit is fed with a low-powered RF signal to be amplified. The SWS electromagnetic field acts upon the e-beam, causing electron bunching and the formation of the so-called space-space-charge wave. In turn, the electromagnetic field generated by space-charge wave induces more current back into the RF circuit with a consequent enhancement of electron bunching. As a result, the EM field is amplified as the RF signal passes down the structure until a saturation regime is reached and a large RF signal is collected at the output. The role of the SWS is to provide slow-wave modes to match up with the velocity of electrons in the e-beam. This velocity is usually a small fraction of the speed of light. Importantly, synchronism is required for effective in-phase interaction between the SWS and the e-beam with an optimal extraction of the kinetic energy of electrons. A typical simple SWS is the helix, which reduces the speed of propagation according to its pitch. The TWT is designed so that the RF signal travels along the tube at nearly the same speed as electrons in the e-beam to facilitate effective coupling. Technical details on the designs and operation of TWTs can be found in Ref. 2 (see Ref. 14, Chap. 4).4,16 As for a rich and interesting history of traveling wave tubes, we refer the reader to Ref. 3 and the references therein.
Assume the Gaussian system of units of physical dimensions of a complete set of e-beam parameters, as in Tables I and II.
Natural units relevant to the e-beam.
Frequency . | Plasma frequency . | . |
---|---|---|
Velocity | e-beam velocity | |
Wavenumber | ||
Length | Wavelength for kq | |
Time | Wave time period |
Frequency . | Plasma frequency . | . |
---|---|---|
Velocity | e-beam velocity | |
Wavenumber | ||
Length | Wavelength for kq | |
Time | Wave time period |
Physical dimensions of e-beam parameters. Abbreviations: dimensionless: dim-less; p/u: per unit.
i . | current . | . |
---|---|---|
q | Charge | |
Number of electrons p/u of volume | ||
The electron plasma wavelength | ||
The e-beam spatial scale | ||
e-beam intensity | ||
Dimensionless e-beam intensity |
i . | current . | . |
---|---|---|
q | Charge | |
Number of electrons p/u of volume | ||
The electron plasma wavelength | ||
The e-beam spatial scale | ||
e-beam intensity | ||
Dimensionless e-beam intensity |
A. TWT Lagrangian and evolution equations
B. Space-charge wave velocity and electron density fields
III. ANALYTIC MODEL OF COUPLED-CAVITY TRAVELING WAVE TUBE
When integrating into the mathematical model significant features of the CCTWT, we make a number of simplifying assumptions. In particular, we use the following basic assumptions of one-dimensional model of space-charge waves in velocity-modulated beams (Ref. 4, Sec. 7.6.1): (i) all quantities of interest only depend on a single space variable z; (ii) the electric field has only an z-component; (iii) there are no transverse velocities of electrons; (iv) ac values are small compared with dc values; (v) electrons have a constant dc velocity, which is much smaller than the speed of light; and (vi) electron beams are nondense. The list of preliminary assumptions of our ideal model for the CCTWT is as follows.
(ideal model of the e-beam and the TL interaction).
E-beam is represented by a cylinder of an infinitesimally small radius having as its axis the z axis (see Fig. 1).
Coupled-cavity structure (CCS) is represented mathematically by a periodic array of adjacent segments of a transmission line (TL) of length a > 0 connected by cavities at points aℓ, by cavities.
- Every cavity carries shunt capacitance c0. The e-beam interacts with the CCS exclusively through the cavities located at a discrete set of equidistant points, that is, the latticewhere a > 0, and we refer to this parameter as the CCS period or just period. The cavity width lg and the corresponding transit time τg are assumed to be zero; see Eq. (1.4) and comments above it.(3.1)
Functions and and their time derivatives and for j = 1, 2 are continuous for all real t and z.
Derivatives , , , and for j = 1, 2 and the mixed derivatives , exist and continuous for all real t and z except for the interaction points in the lattice .
- Let for a function and a real number b symbols and stand for its left and right limit at b assuming their existence, that is,(3.3)
The physical dimensions of quantities related to the cavities and the TL are summarized, respectively, in Tables III and IV.
Physical dimensions of cavity related quantities. Abbreviations: dimensionless: dim-less.
I . | Current . | . |
---|---|---|
Q | Charge | |
c0 | Cavity capacitance | |
l0 | Cavity inductance | |
b | Coupling parameter |
I . | Current . | . |
---|---|---|
Q | Charge | |
c0 | Cavity capacitance | |
l0 | Cavity inductance | |
b | Coupling parameter |
A. CCTWT Lagrangian and the Euler–Lagrange equations
B. Natural units and dimensionless parameters
The natural units relevant to the e-beam in CCTWT are shown in Table V.
Natural units relevant to the e-beam in CCTWT.
Velocity . | e-beam velocity . | . |
---|---|---|
Length | Period | a |
Length | Wavelength | |
Frequency | Period frequency | |
Frequency | Plasma frequency | |
Time | Time of passing the period a | = |
Time | Plasma oscillation time period |
Velocity . | e-beam velocity . | . |
---|---|---|
Length | Period | a |
Length | Wavelength | |
Frequency | Period frequency | |
Frequency | Plasma frequency | |
Time | Time of passing the period a | = |
Time | Plasma oscillation time period |
The CCTWT significant parameters. Abbreviations: dimensionless: dim-less; p/u: per unit, par.: parameter. For the sake of simplicity of the notation, we often omit “prime” super-index, indicating that the dimensionless version of the relevant parameter is involved when it is clear from the context.
a . | The MCK period . | . |
---|---|---|
The e-beam stationary velocity | ||
The period frequency | ||
The plasma frequency | ||
The electron plasma wavelength | ||
The e-beam spatial scale | ||
Normalized period in units of | ||
The number of electrons p/u of volume | ||
c0, l0 | The cavity capacitance, inductance | |
The cavity resonant frequency | ||
The e-beam intensity | ||
Dim-less e-beam intensity | ||
The first interaction par. | ||
The second interaction par. | ||
The gain coefficient | ||
The gain parameter | ||
The capacitance parameter |
a . | The MCK period . | . |
---|---|---|
The e-beam stationary velocity | ||
The period frequency | ||
The plasma frequency | ||
The electron plasma wavelength | ||
The e-beam spatial scale | ||
Normalized period in units of | ||
The number of electrons p/u of volume | ||
c0, l0 | The cavity capacitance, inductance | |
The cavity resonant frequency | ||
The e-beam intensity | ||
Dim-less e-beam intensity | ||
The first interaction par. | ||
The second interaction par. | ||
The gain coefficient | ||
The gain parameter | ||
The capacitance parameter |
C. Euler–Lagrange equations in dimensionless variables
D. CCTWT subsystems: The coupled cavity structure and the e-beam
It is instructive to take a view on the CCTWT system as a composition of its integral components, which are the coupled cavity structure (CCS) and the electron beam (e-beam). It comes as no surprise that special features of the CCS and the e-beam are manifested in fundamental properties of the CCTWT justifying their thorough analysis. This section provides the initial steps of the analysis, whereas more detailed studies of the CCS features are pursued in Sec. IX.
IV. SOLUTIONS TO THE COUPLED-CAVITY TWT EQUATIONS
The dimensionless form of the EL equations [Eqs. (3.34)–(3.36)] and their solutions can be analyzed by applying the Floquet theory reviewed in Appendix F. To use the Floquet theory, we recast first the second-order vector ODE as the first-order vector ODE following our review on the subject in Appendix E.
A. Solutions to the Euler–Lagrange equation inside the period
B. The boundary conditions and the monodromy matrix in dimensionless variables
V. CHARACTERISTIC EQUATION, THE FLOQUET MULTIPLIERS, AND THE DISPERSION RELATION
In what follows, to simplify analytical evaluations, we make the following assumption.
VI. DISPERSION RELATIONS
The CCTWT evolution is governed by spatially periodic ODEs [(3.34)–(3.37)], implying that the dispersion relations as the relations between frequency ω and wavenumber k are constructed based on the Floquet theory reviewed in Appendix F. Specifically, in view of the relation between the Floquet multiplier s and the wave number k (see Appendix F and Remark 25), the characteristic equations [Eqs. (5.1)–(5.4)] can be viewed as an expression of the dispersion relations between the frequency ω and the wavenumber k, and we will refer to it as the CCTWT dispersion relations or just the dispersion relations.
There exists a remarkability in its simplicity relations between the CCTWT dispersion function and the dispersion functions and for, respectively, the CCS and the MCK systems. These relations can be verified by tedious but elementary algebraic evaluations, and they are subjects of the following theorem.
The plots of the four complex eigenvalues (the Floquet multipliers) that solve the characteristic equation [Eq. (5.1)] for the monodromy matrix defined by Eqs. (4.21)–(4.24) in case when χ = 1, ω0 = 1, f = 1, and C0 = 4: (a) K0 = 4, ω = 0.28; (b) K0 = 4, ω = 0.29; and (c) K0 = 3.9, ω = 0.29. The horizontal and vertical axes represent, respectively, and . The eigenvalues are shown by (blue) solid dots.
The plots of the four complex eigenvalues (the Floquet multipliers) that solve the characteristic equation [Eq. (5.1)] for the monodromy matrix defined by Eqs. (4.21)–(4.24) in case when χ = 1, ω0 = 1, f = 1, and C0 = 4: (a) K0 = 4, ω = 0.28; (b) K0 = 4, ω = 0.29; and (c) K0 = 3.9, ω = 0.29. The horizontal and vertical axes represent, respectively, and . The eigenvalues are shown by (blue) solid dots.
(graphical confirmation of the dispersion factorization). The statements of Theorem 2 are well illustrated in Figs. 4(f), 5(f), 6(f), and 7 when compared with Fig. 16 for the CCS and Fig. 15 for the MCK. One can confidently identify in the CCTWT dispersion-instability graphs the patterns of the dispersion-instability graphs of its integral components—the CCS and the MCK.
The dispersion-instability graphs for the CCTWT as the gain coefficient K0 varies. In all plots, the horizontal and vertical axes represent, respectively, and ω. Each of the plots shows three bands of the dispersion of the CCTWT described by Eqs. (5.1)–(5.4) over three Brillouin zones for χ = 1, ω0 = 1, f = 1, and C0 = 1: (a) K0 = 2, fcr ≅ 0.927 292 180, and fmax ≅ 2.034; (b) K0 = 2.25, fcr ≅ 0.837, and fmax ≅ 1.99; (c) K0 = 2.3, fcr ≅ 0.820, and fmax ≅ 1.981; (d) K0 = 2.5, fcr ≅ 0.761, and fmax ≅ 1.951; (e) K0 = 2.7, fcr ≅ 0.709, and fmax ≅ 1.926; and (f) K0 = 5, fcr ≅ 0.395, and fmax ≅ 1.768 (see Theorem 2). When , which is the case of oscillatory modes, and , the corresponding branches are shown as (blue) solid curves. When , that is, there is an instability, and , then the corresponding branches overlap, they are shown as bold solid curves in brown color, and each point of these branches represents exactly two modes with complex-conjugate wave numbers k±.
The dispersion-instability graphs for the CCTWT as the gain coefficient K0 varies. In all plots, the horizontal and vertical axes represent, respectively, and ω. Each of the plots shows three bands of the dispersion of the CCTWT described by Eqs. (5.1)–(5.4) over three Brillouin zones for χ = 1, ω0 = 1, f = 1, and C0 = 1: (a) K0 = 2, fcr ≅ 0.927 292 180, and fmax ≅ 2.034; (b) K0 = 2.25, fcr ≅ 0.837, and fmax ≅ 1.99; (c) K0 = 2.3, fcr ≅ 0.820, and fmax ≅ 1.981; (d) K0 = 2.5, fcr ≅ 0.761, and fmax ≅ 1.951; (e) K0 = 2.7, fcr ≅ 0.709, and fmax ≅ 1.926; and (f) K0 = 5, fcr ≅ 0.395, and fmax ≅ 1.768 (see Theorem 2). When , which is the case of oscillatory modes, and , the corresponding branches are shown as (blue) solid curves. When , that is, there is an instability, and , then the corresponding branches overlap, they are shown as bold solid curves in brown color, and each point of these branches represents exactly two modes with complex-conjugate wave numbers k±.
The dispersion-instability graphs for the CCTWT as the capacitance parameter C0 varies. In all plots, the horizontal and vertical axes represent, respectively, and ω. Each of the plots shows three bands of the dispersion of the CCTWT described by Eqs. (5.1)–(5.4) over three Brillouin zones for χ = 1, ω0 = 1, f = 1, and K0 = 2, fcr ≅ 0.927 292 180, and fmax ≅ 2.034: (a) C0 = 0.7, (b) C0 = 0.3, (c) C0 = 0.15, (d) C0 = 0.1, (e) C0 = 0.05, and (f) C0 = 0.03 (see Theorem 2). When , that is, the case of oscillatory modes, and , the corresponding branches are shown as (blue) solid curves. When , that is, there is an instability, and , then the corresponding branches overlap, they are shown as bold solid curves in brown color, and each point of these branches represents exactly two modes with complex-conjugate wave numbers k±.
The dispersion-instability graphs for the CCTWT as the capacitance parameter C0 varies. In all plots, the horizontal and vertical axes represent, respectively, and ω. Each of the plots shows three bands of the dispersion of the CCTWT described by Eqs. (5.1)–(5.4) over three Brillouin zones for χ = 1, ω0 = 1, f = 1, and K0 = 2, fcr ≅ 0.927 292 180, and fmax ≅ 2.034: (a) C0 = 0.7, (b) C0 = 0.3, (c) C0 = 0.15, (d) C0 = 0.1, (e) C0 = 0.05, and (f) C0 = 0.03 (see Theorem 2). When , that is, the case of oscillatory modes, and , the corresponding branches are shown as (blue) solid curves. When , that is, there is an instability, and , then the corresponding branches overlap, they are shown as bold solid curves in brown color, and each point of these branches represents exactly two modes with complex-conjugate wave numbers k±.
The dispersion-instability graphs for the CCTWT as the normalized period f varies. In all plots, the horizontal and vertical axes represent, respectively, and ω. Each of the plots shows three bands of the dispersion of the CCTWT described by Eqs. (5.1)–(5.4) over three Brillouin zones for χ = 1, ω0 = 1, C0 = 1, K0 = 2, fcr ≅ 0.927 292 180, and fmax ≅ 2.034: (a) f = 0.2, (b) f = 0.8, (c) f = 1.1, (d) f = 1.3, (e) f = 1.7, and (f) f = 2 (see Theorem 2). When , that is, the case of oscillatory modes, and , the corresponding branches are shown as (blue) solid curves. When , that is, there is an instability, and , then the corresponding branches overlap, they are shown as bold solid curves in brown color, and each point of these branches represents exactly two modes with complex-conjugate wave numbers k±.
The dispersion-instability graphs for the CCTWT as the normalized period f varies. In all plots, the horizontal and vertical axes represent, respectively, and ω. Each of the plots shows three bands of the dispersion of the CCTWT described by Eqs. (5.1)–(5.4) over three Brillouin zones for χ = 1, ω0 = 1, C0 = 1, K0 = 2, fcr ≅ 0.927 292 180, and fmax ≅ 2.034: (a) f = 0.2, (b) f = 0.8, (c) f = 1.1, (d) f = 1.3, (e) f = 1.7, and (f) f = 2 (see Theorem 2). When , that is, the case of oscillatory modes, and , the corresponding branches are shown as (blue) solid curves. When , that is, there is an instability, and , then the corresponding branches overlap, they are shown as bold solid curves in brown color, and each point of these branches represents exactly two modes with complex-conjugate wave numbers k±.
The dispersion-instability graphs for the CCTWT as the normalized period f varies (see Theorem 2 and Remark 3). The horizontal and vertical axes represent, respectively, and ω. The plot shows three bands of the dispersion of the CCTWT described by Eqs. (5.1)–(5.4) over three Brillouin zones for χ = 1, ω0 = 1, C0 = 1, f ≅ 1.77, K0 = K0T = 4.95 (typical value) and, consequently, fcr ≅ 0.397, fmax ≅ 1.770. When , which is the case of oscillatory modes, and , the corresponding branches are shown as (blue) solid curves. When , that is, there is an instability, and , then the corresponding branches overlap, they are shown as bold solid curves in brown color, and each point of these branches represents exactly two modes with complex-conjugate wave numbers k± (see Theorem 2 and Remark 3).
The dispersion-instability graphs for the CCTWT as the normalized period f varies (see Theorem 2 and Remark 3). The horizontal and vertical axes represent, respectively, and ω. The plot shows three bands of the dispersion of the CCTWT described by Eqs. (5.1)–(5.4) over three Brillouin zones for χ = 1, ω0 = 1, C0 = 1, f ≅ 1.77, K0 = K0T = 4.95 (typical value) and, consequently, fcr ≅ 0.397, fmax ≅ 1.770. When , which is the case of oscillatory modes, and , the corresponding branches are shown as (blue) solid curves. When , that is, there is an instability, and , then the corresponding branches overlap, they are shown as bold solid curves in brown color, and each point of these branches represents exactly two modes with complex-conjugate wave numbers k± (see Theorem 2 and Remark 3).
A. Graphical representation of the dispersion relations
As to the graphical representation of the dispersion relation, recall that conventional dispersion relations are defined as the relations between real-valued frequency ω and real-valued wavenumber k associated with the relevant eigenmodes. In the case of interest, k can be complex-valued, and to represent all system modes geometrically, we follow Ref. 10, Chap. 7. First, we parameterize every mode of the system uniquely by the pair , where ω is its frequency and is its wavenumber. If is degenerate, it is counted a number of times according to its multiplicity. In view of the importance to us of the mode instability, that is, when , we partition all the system modes represented by pairs into two distinct classes—oscillatory modes and unstable ones—based on whether the wavenumber is real- or complex-valued with . We refer to a mode (eigenmode) of the system as an oscillatory mode if its wavenumber is real-valued. We associate with such an oscillatory mode point in the kω-plane, with k being the horizontal axis and ω being the vertical one. Similarly, we refer to a mode (eigenmode) of the system as a (convective) unstable mode if its wavenumber k is complex-valued with a nonzero imaginary part, that is, . We associate with such an unstable mode point in the kω-plane. Since we consider here only convective unstable modes, we refer to them shortly as unstable modes. Note that every point is, in fact, associated with two complex conjugate system modes with .
Based on the above discussion, we represent the set of all oscillatory and unstable modes of the system geometrically by the set of the corresponding modal points and in the kω-plane. We name this set the dispersion-instability graph. To distinguish graphically points associated oscillatory modes when is real-valued from points associated unstable modes when is complex-valued with , we show points in blue color, whereas points with are shown in brown color. We remind once again that every point with represents exactly two complex conjugate unstable modes associated with .
When and , consequently, the corresponding branches overlap with each point on the segments representing two modes with complex-conjugate wave numbers k±. These branches represent exponential growth or decay in space modes and are shown in plot (c) in brown color.
We generated three sets of dispersion-instability graphs for the CCTWT shown in Figs. 4–6 to demonstrate their dependence on the gain coefficient K0, the capacitance parameter C0, and the normalized period f as they vary in indicated ranges. Figures 4(f), 5(f), 6(f), and 7 when compared with Fig. 16 for the CCS and Fig. 15 for the MCK clearly indicate that the CCTWT dispersion-instability graph is composed of dispersion-instability graphs of its integral components—the CCS and the MCK. The latter is important since the CCS and the MCK are significantly simpler systems compared to the original CCTWT.
B. Exceptional points of degeneracy
Jordan eigenvector degeneracy, which is a degeneracy of the system evolution matrix when not only some eigenvalues coincide but also the corresponding eigenvectors coincide, is sometimes referred to as exceptional point of degeneracy (EPD) (Ref. 22, Sec. II.1). Our prior studies of traveling wave tubes (TWTs) in Ref. 10, Chap. 4, 7, 13, 14, 54, 55 demonstrate that TWTs always have EPDs. A particularly important class of applications of EPDs is sensing.23–27 For applications of EPDs for traveling wave tubes, see Refs. 28–32.
Figure 8 shows examples of the dispersion-instability graphs with EPDs as points, which are the points of the transition to instability. In particular, Fig. 8(c) when compared with Fig. 16 for the CCS and Fig. 15 for the MCK indicates convincingly that the components of the CCTWT dispersion-instability graph can be attributed to the dispersion-instability graphs of its integral components—the CCS and the MCK (see Theorem 2 and Remark 3).
The dispersion-instability graphs for the CCTWT showing the degeneracy (transition to instability) points as (green) diamond dots for C0 = 1, K0 = 2 and, consequently, fcr ≅ 0.927, fmax ≅ 1.770: (a) f = 0.9 < fcr, (b) f = 1.1 > fcr, and (c) f = 1.5 > fcr (see Theorem 2 and Remark 3). In all plots, the horizontal and vertical axes represent, respectively, and ω. Solid (grin) diamond dots identify points of the transition from the instability to the stability, which are also EPD points.
The dispersion-instability graphs for the CCTWT showing the degeneracy (transition to instability) points as (green) diamond dots for C0 = 1, K0 = 2 and, consequently, fcr ≅ 0.927, fmax ≅ 1.770: (a) f = 0.9 < fcr, (b) f = 1.1 > fcr, and (c) f = 1.5 > fcr (see Theorem 2 and Remark 3). In all plots, the horizontal and vertical axes represent, respectively, and ω. Solid (grin) diamond dots identify points of the transition from the instability to the stability, which are also EPD points.
VII. GAIN EXPRESSION IN TERMS OF THE FLOQUET MULTIPLIERS
Just as in the case of the dispersion relations that we analyzed in Sec. VI, there are simple relations between the CCTWT characteristic function and the characteristic polynomials and for, respectively, the CCS and MCK systems. These relations can be verified by tedious but elementary algebraic evaluations, and they are subjects of the following theorem that relates the characteristic polynomials for CCTTX, CCS, and MCK systems.
(graphical confirmation of the characteristic polynomial factorization). The statements of Theorem 4 are well illustrated in Figs. 9 and 10 when compared with Fig. 17 for the CCS and Fig. 12 for the MCK. One can confidently recognize in components of the graph of the gain CCTWT the patents of the graphs for the gain of the CCS and the MCK.
Plots of gain G per one period in dB as a function of frequency ω defined by Eq. (7.1) for ω0 = 1, K0 = 2.5 and, consequently, fcr ≅ 0.761 012 7542, fmax ≅ 1.951 302 704, Gmax = 14.307 667 94 (see Sec. VIII B for the definition of the MCK quantities fcr, fmax, and Gmax): (a) f = 1.2 > fcr; (b) f = 1.95 ≈ fmax. In all plots, the horizontal and vertical axes represent, respectively, frequency ω and gain G in dB. The (brown) solid curves represent gain G as a function of frequency ω; the (blue) dashed horizontal line G = Gmax represents the maximal Gmax value of G in the high frequency limit (see Sec. VIII B); and the (green) dashed horizontal line represents the value of G in the high frequency limit for given value of f (see Sec. VIII B). The envelope of the local maxima of the gain for large values of frequency ω behaves as (see captions of Fig. 17), and it is shown as a (blue) dashed curve.
Plots of gain G per one period in dB as a function of frequency ω defined by Eq. (7.1) for ω0 = 1, K0 = 2.5 and, consequently, fcr ≅ 0.761 012 7542, fmax ≅ 1.951 302 704, Gmax = 14.307 667 94 (see Sec. VIII B for the definition of the MCK quantities fcr, fmax, and Gmax): (a) f = 1.2 > fcr; (b) f = 1.95 ≈ fmax. In all plots, the horizontal and vertical axes represent, respectively, frequency ω and gain G in dB. The (brown) solid curves represent gain G as a function of frequency ω; the (blue) dashed horizontal line G = Gmax represents the maximal Gmax value of G in the high frequency limit (see Sec. VIII B); and the (green) dashed horizontal line represents the value of G in the high frequency limit for given value of f (see Sec. VIII B). The envelope of the local maxima of the gain for large values of frequency ω behaves as (see captions of Fig. 17), and it is shown as a (blue) dashed curve.
Plot of gain G per one period in dB as a function of frequency ω defined by Eq. (9.24) for ω0 = 1, K0 = K0T = 4.95 and, consequently, fcr ≅ 0.398 674 6100, fmax ≅ 1.770 133 632, Gmax = 20 (see Sec. VIII B for the definition of the MCK quantities fcr, fmax, and Gmax), and f = 1.2 > fcr. The horizontal and vertical axes represent, respectively, frequency ω and gain G in dB. The (brown) solid curves represent gain G as a function of frequency ω; the (blue) dashed horizontal line G = Gmax represents the maximal Gmax value of G in the high frequency limit (see Sec. VIII B); and the (green) dashed horizontal line represents the value of G in the high frequency limit for f = fcr (see Sec. VIII B). The envelope of the local maxima of the gain for large values of frequency ω behaves as (see captions of Fig. 17), and it is shown as a (blue) dashed curve.
Plot of gain G per one period in dB as a function of frequency ω defined by Eq. (9.24) for ω0 = 1, K0 = K0T = 4.95 and, consequently, fcr ≅ 0.398 674 6100, fmax ≅ 1.770 133 632, Gmax = 20 (see Sec. VIII B for the definition of the MCK quantities fcr, fmax, and Gmax), and f = 1.2 > fcr. The horizontal and vertical axes represent, respectively, frequency ω and gain G in dB. The (brown) solid curves represent gain G as a function of frequency ω; the (blue) dashed horizontal line G = Gmax represents the maximal Gmax value of G in the high frequency limit (see Sec. VIII B); and the (green) dashed horizontal line represents the value of G in the high frequency limit for f = fcr (see Sec. VIII B). The envelope of the local maxima of the gain for large values of frequency ω behaves as (see captions of Fig. 17), and it is shown as a (blue) dashed curve.
(amplification in stopbands). E-beam interactions in periodic slow-wave structures were studied by Solntsev in Ref. 33. Under the condition of exact synchronism as in our assumption 3, the amplification was observed in stopbands, known also as spectral gaps in the system (oscillatory) spectrum. Our theory accounts for this general spectral phenomenon too as indicated by growing in magnitude “bumps” in Figs. 9 and 10. One can also see similar bumps in Fig. 17 for the CCS.
VIII. SKETCH OF THE MULTICAVITY KLYSTRON ANALYTICAL MODEL
Usage of cavity resonators in the klystron was a revolutionary idea of Hansen and the Varians (Ref. 4, Sec. 7.1). In the pursuit of higher power and efficiency, the original design of Vairan klystrons evolves significantly over years featuring today multiple cavities and multiple electron beam (Ref. 4, Sec. 7.7). The advantages of klystrons are their high power and efficiency, potentially wide bandwidth, phase, and amplitude stability (Ref. 34, Sec. 9.1).
The construction of an analytic model for the multicavity klystron (MCK) in Ref. 11 utilizes elements of the analytic model of the traveling wave tube (TWT) introduced and studied in our monograph (Ref. 10, Chap. 4, 24); see Sec. II. Multicavity klystron, known also as cascade amplifier (Ref. 9, Sec. IIb), is composed of the e-beam interacting with a periodic array of electromagnetic cavities; see Fig. 11. Consequently, the MCK can be naturally viewed as a subsystem of the CCTWT that contributes to the properties of CCTWT.
A schematic representation of a multicavity klystron (MCK) that exploits constructive interaction between the pencil-like electron beam and an array of electromagnetic cavities (often of toroidal shape). The interaction causes electron bunching and consequent amplification of the RF signal.
A schematic representation of a multicavity klystron (MCK) that exploits constructive interaction between the pencil-like electron beam and an array of electromagnetic cavities (often of toroidal shape). The interaction causes electron bunching and consequent amplification of the RF signal.
A. The Euler–Lagrange equations in dimensionless variables
B. The monodromy matrix, the dispersion-instability relations, and the gain
Plots of gain G as a function of frequency ω defined by Eq. (8.19) for ω0 = 1 and (a) K0 = 2, f = 2 > fcr with fcr ≅ 0.927 295 2180, fmax ≅ 2.034 443 936, and Gmax ≅ 12.539 258 41; (b) K0 = 1, f = 2.4 > fcr with fcr ≅ 1.570 796 327, fmax ≅ 2.356 194 491, and Gmax ≅ 7.655 513 706. In all plots, the horizontal and vertical axes represent, respectively, the frequency ω and gain G in dB. The (brown) solid curves represent gain G as a function of frequency ω; the (blue) dashed line G = Gmax represents the maximal Gmax value of G in the high frequency limit. The (green) diamond solid dots mark the values of , which is the lower frequency boundary of the instability interval.
Plots of gain G as a function of frequency ω defined by Eq. (8.19) for ω0 = 1 and (a) K0 = 2, f = 2 > fcr with fcr ≅ 0.927 295 2180, fmax ≅ 2.034 443 936, and Gmax ≅ 12.539 258 41; (b) K0 = 1, f = 2.4 > fcr with fcr ≅ 1.570 796 327, fmax ≅ 2.356 194 491, and Gmax ≅ 7.655 513 706. In all plots, the horizontal and vertical axes represent, respectively, the frequency ω and gain G in dB. The (brown) solid curves represent gain G as a function of frequency ω; the (blue) dashed line G = Gmax represents the maximal Gmax value of G in the high frequency limit. The (green) diamond solid dots mark the values of , which is the lower frequency boundary of the instability interval.
Plots of as the (brown) solid curve and as the (blue) dashed curve. The horizontal and vertical axes represent, respectively, K0 and f.
Plots of as the (brown) solid curve and as the (blue) dashed curve. The horizontal and vertical axes represent, respectively, K0 and f.
Equations (8.30) in turn can be recast into an even more explicit form as stated in the following theorem.11
Equations (8.32) for the complex-valued wave numbers represent the dispersion relations of the MCK.
Theorem 2 shows how the MCK dispersion function and its high-frequency approximation are integrated into the relevant dispersion functions associated with the CCTWT.
Figures 14 and 15 illustrate graphically the dispersion relations described by Eq. (8.32). The pairs of nearly straight lines above the shadowed instability zone depicted in Fig. 14 are consistent with the high-frequency approximation (8.39) to the MCK dispersion relation.
The MCK dispersion-instability plot (brown solid curves and lines) over three Brillouin zones for K0 = 3, ω0 = 1 for which fcr ≅ 0.643 5011 088, fmax ≅ 1.892 546 882, and f = 0.5 < fcr ≅ 0.643 501 1088. The horizontal and vertical axes represent, respectively, and . Two (green) solid diamond dots identify the values of and , which are the frequency boundaries of the instability. The (brown) solid disk dots identify points of the transition from the instability to the stability, which are also EPD points. Two (brown) horizontal dashed-dotted lines identify the frequency boundaries of the instability, and the (light blue) shaded region between the lines identifies points of instability. The (green) dashed horizontal line ω = ω0 identifies the resonance frequency ω0. Note that the plot has a jump-discontinuity along the (green) dashed line, namely, jumps by π according to Eq. (8.32) as the frequency ω passes through the resonance frequency ω0 and the sign of changes. The shadowed area marks points associated with the instability. The (blue) dashed straight lines correspond to the high frequency approximation defined by Eq. (8.39).
The MCK dispersion-instability plot (brown solid curves and lines) over three Brillouin zones for K0 = 3, ω0 = 1 for which fcr ≅ 0.643 5011 088, fmax ≅ 1.892 546 882, and f = 0.5 < fcr ≅ 0.643 501 1088. The horizontal and vertical axes represent, respectively, and . Two (green) solid diamond dots identify the values of and , which are the frequency boundaries of the instability. The (brown) solid disk dots identify points of the transition from the instability to the stability, which are also EPD points. Two (brown) horizontal dashed-dotted lines identify the frequency boundaries of the instability, and the (light blue) shaded region between the lines identifies points of instability. The (green) dashed horizontal line ω = ω0 identifies the resonance frequency ω0. Note that the plot has a jump-discontinuity along the (green) dashed line, namely, jumps by π according to Eq. (8.32) as the frequency ω passes through the resonance frequency ω0 and the sign of changes. The shadowed area marks points associated with the instability. The (blue) dashed straight lines correspond to the high frequency approximation defined by Eq. (8.39).
The MCK dispersion-instability plot (solid brown curves and lines) over three Brillouin zones for K0 = 1, ω0 = 1 for which fcr ≅ 1.570 796 327, fmax ≅ 1.892 546 882, and f = 1.569 ≅ fcr ≅ 1.570 796 327. The horizontal and vertical axes represent, respectively, and . Two (green) solid diamond dots identify the values of and , which are the frequency boundaries of the instability. The (brown) solid disk dots identify points of the transition from the instability to the stability, which are also EPD points. Two (brown) horizontal dashed-dotted lines identify the frequency boundaries of the instability, and the (light blue) shaded region between the lines identifies points of instability. The (green) dashed horizontal line ω = ω0 identifies the resonance frequency ω0. Note that the plot has a jump-discontinuity along the (green) dashed line, namely, jumps by π according to Eq. (8.32) as the frequency ω passes through the resonance frequency ω0 and the sign of changes. The shadowed area marks points associated with the instability.
The MCK dispersion-instability plot (solid brown curves and lines) over three Brillouin zones for K0 = 1, ω0 = 1 for which fcr ≅ 1.570 796 327, fmax ≅ 1.892 546 882, and f = 1.569 ≅ fcr ≅ 1.570 796 327. The horizontal and vertical axes represent, respectively, and . Two (green) solid diamond dots identify the values of and , which are the frequency boundaries of the instability. The (brown) solid disk dots identify points of the transition from the instability to the stability, which are also EPD points. Two (brown) horizontal dashed-dotted lines identify the frequency boundaries of the instability, and the (light blue) shaded region between the lines identifies points of instability. The (green) dashed horizontal line ω = ω0 identifies the resonance frequency ω0. Note that the plot has a jump-discontinuity along the (green) dashed line, namely, jumps by π according to Eq. (8.32) as the frequency ω passes through the resonance frequency ω0 and the sign of changes. The shadowed area marks points associated with the instability.
IX. COUPLED CAVITY STRUCTURE
We introduce and study here basic properties of the coupled cavity structure (CCS). Since CCS is naturally an integral part of CCTWT, the knowledge of its properties would allow us to find out its contribution to the properties of CCTWT. For particular designs of coupled cavities and the way they interact with TWTs, see Ref. 34, Sec. 9.1, 9.3.3.
A. Monodromy matrix and the dispersion-instability relations
The CCS for C0 = 1. (a) Dispersion-instability graph: horizontal axis is and the vertical axis is ω. (b) The plot of the instability parameter defined by the second equation in (9.13). The horizontal axis is , and the vertical axis is W.
The CCS for C0 = 1. (a) Dispersion-instability graph: horizontal axis is and the vertical axis is ω. (b) The plot of the instability parameter defined by the second equation in (9.13). The horizontal axis is , and the vertical axis is W.
Based on the prior analysis, we introduce the CCS gain GC in dB per one period as a the rate of the exponential growth of the CCS eigenmodes associated with Floquet multipliers s± defined by Eq. (9.16). More precisely, the definition is as follows.
Figure 17 shows the frequency dependence of the gain GC per one period. Growing in magnitude “bumps” in Fig. 17 indicates the presence of gain/amplification inside of stopbands, known also as spectral gaps in the system (oscillatory) spectrum, of the CCS; see Remark 6.
The plot of the CCS gain per one period for C0 = 0.5. The horizontal and vertical axes represent, respectively, the frequency ω and gain G in dB. The instability frequencies ω are identified by condition . The envelope of the local maxima of the gain behaves asymptotically for large values of frequency ω as as it follows from Eq. (9.24). It is shown as a (blue) dashed curve.
The plot of the CCS gain per one period for C0 = 0.5. The horizontal and vertical axes represent, respectively, the frequency ω and gain G in dB. The instability frequencies ω are identified by condition . The envelope of the local maxima of the gain behaves asymptotically for large values of frequency ω as as it follows from Eq. (9.24). It is shown as a (blue) dashed curve.
B. Exceptional points of degeneracy
X. THE KINETIC AND FIELD POINTS OF VIEW ON THE GAP INTERACTION
We compare here some of the features of our field theory with the relevant features of the kinematic/ballistic theory of the CCTWT operation. Before going into technical details, we would like to point out that from the outset, our Lagrangian field theory takes into account the space-charge forces, that is, the electron-to-electron repulsion, whereas the standard hydrokinetic analysis completely neglects them.
A. Some points from the kinetic theory
We briefly review here some points of the kinetic/ballistic theory. Kinematic analysis of the CCTWT operation involves (i) the electron velocity modulation in gaps of the klystron cavities, (ii) consequent electron bunching, (iii) the energy exchange between the e-beam to the EM field, and (iv) the energy transfer from the e-beam to the EM field under proper conditions and consequent RF signal amplification. The listed subjects were thoroughly studied by many scholars; see, for instance, Refs. 5, 21, and 37 (see Ref. 2, Chap. 15, Ref. 35, Sec. 7.2, Ref. 4, Secs. 6.1–6.3; 7.1–7.7, and Ref. 1, Chap. II] and the references therein. When presenting relevant to us conclusions of the studies, we mostly follow the hydrokinetic (ballistic) approach that utilizes the Eulerian (spatial) and the Lagrangian (material) descriptions (points of view) as in Ref. 4, Sec. 7.1–7.7 and Ref. 1, Chap. II. As to general aspects of the hydrokinetic approach in continua, which includes, in particular, the Eulerian and the Lagrangian descriptions, we refer the reader to Ref. 38, Secs. I.4–I.8, Ref. 39, Secs. 3.1–3.2, and Ref. 40, Sec. 1.7.
In other words, according the above scenario, electron bunching takes place. More precisely, the velocity-modulated, uniformly dense electron beam becomes a density-modulated beam with nearly constant dc velocity .
B. Field theory point of view on the kinetic properties of the electron flow
According the CCTWT design, all the interactions between the electron flow and the EM field occur in cavity gaps. In what follows, we use notations and results from Sec. II B. Let us consider first the action of the cavity ac EM field on the electron flow. The cavity ac EM field acts upon the e-beam by accelerating and decelerating its electrons and effectively modulating their velocities by the relatively small compared to electron velocity field . Therefore, as to this part of the interaction, we may view the electron density to be essentially constant , whereas its ac velocity field is modulated by ac EM field. Consider now the action of the e-beam on the cavity ac EM field. The space charge acts upon the cavity ac EM field essentially quasi-electrostatically through relatively small ac electron number density field . Hence, for this part of the interaction, we may view the electron flow to be of nearly constant velocity perturbed by relatively small ac electron number density . Following the results of Sec. III, let us take a look at the variation of the ac electron velocity and ac electron number density in the vicinity of centers aℓ of the cavity gaps.
Note that according to Eq. (10.4), the jumps in the velocity the number density are in antiphase.
C. Relation between the kinetic and the field points of view on the gap interaction
An insightful comparative analysis of “electron-wave theory” and the kinetic/ballistic theory of bunching is provided in Ref. 1, Sec. II.15.
A description of the mechanism of phase focusing as a phenomenon of oscillating space-charge waves is only a mathematical description of a process, the essence of which is as follows: The initial velocity modulation gives rise to periodic concentration and dispersion of electron space charge. The amount of bunching and the associated alternating current increase through the bunching region, provided that there are no repulsive space-charge forces affecting this process. Space-charge forces oppose the initial velocity modulation and cause additional retardation and acceleration of the electrons. Thus, the law of conservation of energy is obeyed. On the other hand, the ballistic theory is fundamentally contradictory to this.
In fact, the ballistic theory of bunching assumes that the alternating velocity acquired by the electrons in the modulator remains constant along the whole path. However, the potential energy necessarily increases after electron bunching, and so the total energy of the electron beam constantly varies, and this conflicts with the law of conservation of energy. Despite this contradiction, the ballistic theory is a good enough approximation for many of the cases met with in practice. In this case, both ballistic and electron-wave theories lead to identical results.
In agreement with the above quotation, our field theory of the space-charge wave can be viewed as effective mathematical descriptions of the underlying physical complexity involving the electron velocity and the electron number densities.
As to the energy conservation, unlike the kinetic theory, our Lagrangian field theory surely provides for that. The field theory under some conditions agrees at least with some points of the kinetic/ballistic theory as we discuss below.
The hydrokinetic point of view on our simplifying assumption that the cavity width lg and the corresponding transit time τg are zeros, see Eq. (1.4) and assumption 1, is as follows (Ref. 1, Sec. II.5):
“Let us assume further that the transit time of electrons between grids 1 and 2 is infinitesimally small, which means a physically small transit time compared with the period of oscillation of the high-frequency field. If the transit time is negligible, electrons can be considered to move through a constant (momentarily) alternating field, i.e., virtually in a static field. The electrons acquire or lose an amount of energy equal to the product of the electron charge and the momentary value of the voltage. Therefore electrons entering the space between the grids at different moments in time, with equal velocities, pass out of this space at different velocities which are determined by the momentary value of the alternating voltage. The electron beam is thus velocity modulated and has a uniform density of space charge.”
Another link between the field and the kinetic theories comes from our analysis in Sec. X B. In view of Eqs. (10.3) and (10.4), jumps that are explicitly allowed by the field theory represent jumps and related to the kinetic properties of the electron flow; see Remark 1. Namely, jump manifests the electron bunching, jump manifests the ac electron velocity modulation, and Eq. (10.4) relates the two of them.
XI. LAGRANGIAN VARIATIONAL FRAMEWORK
We construct here the Lagrangian variational framework for our model of CCTWT. According to assumption 1, the model integrates into it quantities associated with continuum of real numbers, on one hand, and features associated with discrete points, on the other hand. The continuum features are represented by Lagrangian densities and in Eq. (3.10), whereas discrete features are represented by Lagrangian in Eq. (3.11) with energies concentrated in a set of discrete points . One possibility for constructing the desired Lagrangian variational framework is to apply the general approach developed in Ref. 41 when the “rigidity” condition holds. Another possibility is to directly construct the Lagrangian variational framework using some ideas from Ref. 41, and that is what we actually pursue here.
XII. ROOT DEGENERACY FOR A SPECIAL POLYNOMIAL OF THE FOURTH DEGREE
ACKNOWLEDGMENTS
This research was supported by AFOSR MURI under Grant No. FA9550-20-1-0409 administered through the University of New Mexico. The author is grateful to E. Schamiloglu for sharing his deep and vast knowledge of high power microwave devices and inspiring discussions.
AUTHOR DECLARATIONS
Conflict of Interest
The author has no conflicts to disclose.
Author Contributions
Alexander Figotin: Writing – original draft (lead).
DATA AVAILABILITY
The data that support the findings of this study are available within the article.
NOMENCLATURE
set of complex number
set of n dimensional column vectors with complex complex-valued entries
set of n × m matrices with complex-valued entries
CCTWT dispersion function
CCS dispersion function
MCK dispersion function
the determinant of matrix A
block diagonal matrix with indicated blocks
dimension of the vector space W
- EL
the Euler–Lagrange (equations)
identity matrix
kernel of matrix A, which is the vector space of vector x such that Ax = 0
- MT
matrix transposed to matrix M
- ODE
ordinary differential equation
complex-conjugate to complex number s
spectrum of matrix A
set of n × m matrices with real-valued entries
characteristic polynomial of a ν × ν matrix A
APPENDIX A: FOURIER TRANSFORM
APPENDIX B: JORDAN CANONICAL FORM
We provide here a very concise review of Jordan canonical forms following mostly Ref. 47, Sec. III.4 and Ref. 48, Secs. 3.1 and 3.2. As to a demonstration of how Jordan block arises in the case of a single nth order differential equation, we refer to Ref. 49, Sec. 25.4.
Let A be an n × n matrix and λ be its eigenvalue, and let be the least integer k such that , where is a null space of a matrix C. Then, we refer to as the generalized eigenspace of matrix A corresponding to eigenvalue λ. Then, the following statements hold (Ref. 47, Sec. III.4).
APPENDIX C: COMPANION MATRIX AND CYCLICITY CONDITION
An eigenvalue is called cyclic (nonderogatory) if its geometric multiplicity is 1. A square matrix is called cyclic (nonderogatory) if all its eigenvalues are cyclic (Ref. 51, Sec. 5.5). The following statement provides different equivalent descriptions of a cyclic matrix (Ref. 51, Sec. 5.5).
Furthermore, the following statements are equivalent:
.
A is cyclic.
For every ζj, the Jordan form of A contains exactly one block associated with ζj.
A is similar to the companion matrix .
The following statement summarizes important information on the Jordan form of the companion matrix and the generalized Vandermonde matrix (Ref. 51, Sec. 5.16, Ref. 52, Sec. 2.11, and Ref. 50, Sec. 7.9).
APPENDIX D: MATRIX POLYNOMIALS
Polynomial vector identity (D13) readily follows from Eqs. (D11) and (D12). Identities (D14) for the determinants follow straightforwardly from Eqs. (D12), (D15), and (D9). If , then the degree of the polynomial has to be mν since and are mν × mν matrices.
Suppose that Eq. (D15) holds. Then, combining them with proven identity (D13), we get , proving that expressions (D15) define an eigenvalue s and an eigenvector .
(characteristic polynomial degree). Note that according to Theorem 15, the characteristic polynomial for the m × m matrix polynomial has the degree mν, whereas in the linear case for the m × m identity matrix and m × m matrix A0, the characteristic polynomial is of the degree m. This can be explained by observing that in the non-linear case of m × m matrix polynomial , we are dealing effectively with many more m × m matrices A than just a single matrix A0.
Another problem of our particular interest related to the theory of matrix polynomials is the degeneracy of eigenvalues and eigenvectors and, consequently, the existence of non-trivial Jordan blocks, that is, Jordan blocks of dimensions higher or equal to 2. The general theory addresses this problem by introducing the so-called “Jordan chains,” which are intimately related to the theory of system of differential equations expressed as and their solutions of the form , where is a vector polynomial; see Appendix E and Ref. 53, Chap. I, II and Ref. 54, Chap. 9. Avoiding the details of Jordan chain developments, we simply note that an important point of Theorem 15 is that there is one-to-one correspondence between solutions of equations and , and it has the following immediate implication.
The next statement shows that if the geometric multiplicity of an eigenvalue is strictly less than its algebraic one, then there exist non-trivial Jordan blocks, that is, Jordan blocks of dimensions higher or equal to 2.
The Proof of Theorem 18 follows straightforwardly from the definition of the Jordan canonical form and its basic properties. Note that if Eq. (D23) holds, it implies that eigenvalue 0 is cyclic (nonderogatory) for matrix and eigenvalue s0 is cyclic (nonderogatory) for matrix , provided that exists; see Appendix C.
APPENDIX E: VECTOR DIFFERENTIAL EQUATIONS AND THE JORDAN CANONICAL FORM
In this section, we relate the vector ordinary differential equations to the matrix polynomials reviewed in Appendix D following Ref. 55, Secs. 5.1 and 5.7, Ref. 53, Sec. II.8.3, Ref. 47, Sec. III.4, and Ref. 50, Sec. 7.9.
1. Constant coefficients case.
APPENDIX F: FLOQUET THEORY
The monodromy matrix is integrated into the formulation of the main statement of the Floquet theory describing the structure of solutions to Eq. (F11) for ς-periodic matrix function .
The eigenvalues of the monodromy matrix are called Floquet (characteristic) multipliers, and their logarithms (not uniquely defined) are called characteristic exponents.
(Floquet eigenmodes). If is the Floquet eigenmode defined by (F18) and or, equivalently, , then grows exponentially as z → +∞, and we refer to such as the exponentially growing Floquet eigenmode. In the case when or equivalently , function is bounded and we refer to such as an oscillatory Floquet eigenmode.
APPENDIX G: HAMILTONIAN SYSTEM OF LINEAR DIFFERENTIAL EQUATIONS
1. Symmetry of the spectra.
G-unitary, G-skew-Hermitian, and G-Hermitian matrices have special properties described in Table VII. These properties can viewed as symmetries, and not surprisingly, they imply consequent symmetries of the spectra of the matrices. Let denote the spectrum of matrix A. It is a straightforward exercise to verify based on matrix properties described in Table VII that the following statements hold.
G-unitary, G-skew-Hermitian, and G-Hermitian matrices.
G-unitary . | G-skew-Hermitian . | G-Hermitian . |
---|---|---|
, | A† = G−1A*G = −A, | A† = G−1A*G = A, |
A* = GA−1G−1 | GA + A*G = 0 | GA − A*G = 0 |
A*GA = G | A = iG−1H, H = H* | A = G−1H, H = H* |
G-unitary . | G-skew-Hermitian . | G-Hermitian . |
---|---|---|
, | A† = G−1A*G = −A, | A† = G−1A*G = A, |
A* = GA−1G−1 | GA + A*G = 0 | GA − A*G = 0 |
A*GA = G | A = iG−1H, H = H* | A = G−1H, H = H* |
(spectral symmetries). Suppose that matrix A is either G-unitary or G-skew-Hermitian or G-Hermitian. Then, the following statements hold:
- If A is G-unitary, then is symmetric with respect to the unit circle, that is,(G10)
- If A is G-skew-Hermitian, then is symmetric with respect the imaginary axis, that is,(G11)
- If A is G-Hermitian, then is symmetric with respect to real axis, that is,(G12)
The following statement describes the G-orthogonality of invariant subspaces of G-unitary, G-skew-Hermitian, and G-Hermitian matrices (Ref. 56, Sec. 1.8).
The statement below describes a special property of eigenvectors of a G-unitary matrix.
2. Special Hamiltonian systems.
APPENDIX H: FOLDED WAVEGUIDE TWT DISPERSION RELATIONS
APPENDIX I: CAPACITANCE
According to Ref. 59, Sec. I.1 and Ref. 60, Sec. 3.5.2, the following formulas hold for the capacitance of capacitors of different geometries in Gaussian system of units.