Photon emitters are becoming increasingly important due to their ability to generate high brightness, low emittance, and spatiotemporally coherent electron bunches for multiple applications; however, these emitters rarely produce electrons solely due to photoemission. Often, photon emitters are prone to undesired thermionic emission; alternatively, some devices intentionally leverage field and thermionic emission to increase output current. Regardless, attempting to extract higher currents from these devices raises concerns about space-charge buildup. While theories have examined the transitions between many of these mechanisms, none have used a common framework to unify photo-, thermionic, field, and space-charge limited emission simultaneously, typically represented individually by the Fowler–Dubridge (FD), Richardson–Laue–Dushman (RLD), Fowler–Nordheim (FN), and Child–Langmuir (CL) equations, respectively. This paper derives an exact solution unifying these mechanisms and reports conditions where emission bypasses RLD to directly transition from FD to FN based on asymptotically matching the three models at a nexus point. Furthermore, we provide a step-by-step approach for developing nexus phase space plots exhibiting the operating conditions for transitions among FD, RLD, FN, CL, Mott–Gurney for space-charge limited current with collisions, and Ohm's law for an external resistor. We demonstrate the utility of nexus plots for assessing the applicability of the simple well-known theories based on a single mechanism or the necessity to use more complicated solutions combining multiple mechanisms. As such, nexus theory provides a simple framework for guiding theorists in model development, simulation experts in algorithm development and selection, and experimentalists in device design.

Electron emission is a critical phenomenon in multiple applications, including high power microwaves, electronic thrusters, nanovacuum transistors, and time-resolved electron microscopy, and across multiple phases, such as vacuum, gas, liquid, and solid.1–3 While many electron emission studies focus on vacuum,1–3 recent experiments and theories have examined the implications of field emission (FE) on gas breakdown mechanisms for microscale and smaller gaps.4–7 This has opened up a wider examination of the unification of electron emission mechanisms across various parameter spaces,4 including emitter temperature,8 charged particle mobility in the gap,9 external series resistance,10 and diode gap distance.11 

A seminal study unified field emission with space-charge limited current (SCLC) at vacuum by using the Fowler–Nordheim (FN) equation for field emission current density and solving the force equation to determine the transit time for a single electron to cross the diode.12 Performing a matched asymptotic analysis, this study recovered the Child–Langmuir law (CL), given by13,14

(1)

where ε0 is the permittivity of free space, e is the electron charge, m is the electron mass, D is the gap distance, and V is the bias voltage, for pure planar SCLC at high V or small D; at low V or high D, this study recovered the FN equation,12 given by15–17 

(2)

where

(3)
(4)

Φ is the electrode work function, Es is the electric field at the cathode, y=Φ14QeEs, where Q=e2/(16πϵ0), and AFN=e3/(16π2Φ) and BFN=42mΦ3/(3e) are FN constants, where is the reduced Planck's constant. Table I summarizes the values of key physical parameters used in these equations and throughout this paper.

TABLE I.

Physical constants used in calculations.

DescriptionSymbolValue
Speed of light c 2.997 92 × 108 m s−1 
Electron charge e 1.602 176 62 × 10−19 C 
Permittivity of free space ε0 8.854 187 812 8 × 10−12 F m−1 
Electron mass m 9.109 383 56 × 10−31 kg 
Reduced Planck's constant  1.054 571 8 × 10−34 J s 
Boltzmann's constant k 1.380 648 52 × 10−23 J K−1 
Work function constant Q=(16π)1ϵ01e2 1.602 176 62 × 10−19 J m 
Richardson constant ARLD=emkB22π22 1.2017 × 106 A K−2 m−2 
Fowler–Nordheim constant AFN 3.4254 × 10−7 A V−2 
Fowler–Nordheim constant BFN 6.5207 × 1010 V m−1 
DescriptionSymbolValue
Speed of light c 2.997 92 × 108 m s−1 
Electron charge e 1.602 176 62 × 10−19 C 
Permittivity of free space ε0 8.854 187 812 8 × 10−12 F m−1 
Electron mass m 9.109 383 56 × 10−31 kg 
Reduced Planck's constant  1.054 571 8 × 10−34 J s 
Boltzmann's constant k 1.380 648 52 × 10−23 J K−1 
Work function constant Q=(16π)1ϵ01e2 1.602 176 62 × 10−19 J m 
Richardson constant ARLD=emkB22π22 1.2017 × 106 A K−2 m−2 
Fowler–Nordheim constant AFN 3.4254 × 10−7 A V−2 
Fowler–Nordheim constant BFN 6.5207 × 1010 V m−1 

Motivated by a theoretical analysis of plasma sheath formation linking CL at vacuum with the Mott–Gurney law (MG)18 for SCLC when collisions with neutral particles (dissipation) are important, which is given by19 

(5)

where μ is electron mobility, which is generally inversely proportional to pressure, we modified the electron force law from the prior vacuum assessment12 to incorporate collisions through μ.9 Accounting for collisions causes electron emission to asymptotically approach MG for small D at low μ and moderately high V. The SCLC solution always transitions from MG to CL with any combination of increasing μ, increasing V, and decreasing D, all of which diminish collisional effects until the electron acts as if the gap were in vacuum.9 This approach yielded phase space plots based on equations (1), (2), and/or (5), representing the asymptotic solutions of the exact solution for electron transit time in the appropriate limits for FN, CL, and MG, respectively, to determine the transition point between dominant electron emission mechanisms.9 We subsequently defined matching such asymptotic solutions as “nexus theory.”4 We refer to matching two asymptotic solutions as a “second-order” nexus, matching three asymptotic solutions as a “third-order” nexus, and so forth. For the case of MG, CL, and FN, we obtained a third-order nexus where all three asymptotic solutions met and showed that defining any of V, μ, or D uniquely defines the other two parameters at the third-order nexus, where it is necessary to simultaneously account for all three mechanisms.9 

Practical devices are often more complicated than the simple planar diode described above. For instance, practical electron emitters use a series resistor to limit the current to the emitter to minimize damage.20,21 Placing a resistor in series with the diode divides the applied voltage between the two elements, which may be determined by applying Kirchhoff's law to a circuit composed of the resistor in series with the impedance of the diode, which may be modeled as a parallel plate capacitor in the circuit.10,21 This approach permitted the derivation of the transitions among FN, CL, and Ohm's law (OL);21 incorporating μ in the electron force law introduced transitions involving MG and a fourth-order nexus among all these mechanisms.10 We emphasize that while OL is not an electron emission mechanism, it represents a limit on emission current (just as SCLC does) and is important for characterizing the contribution of the series resistor to overall device operation.

In addition to circuit elements, not all devices use “cold cathodes” (zero initial electron velocity) for emission; many devices use higher temperature thermionic emission (TE) and there is a vast literature considering cold and thermionic cathodes driven by field and thermionic emission, respectively.22 Thermionic emission is typically defined by the Richardson–Laue–Dushman (RLD) law, given by23 

(6)

where T is the cathode temperature, kB is the Boltzmann constant, ARLD=qm/(2π23), and q is the particle charge. The transition between thermionic and field emission is fully accounted for by the general thermo-field current density equation JGTF.2,24–26 Additionally, many studies have examined Miram curves, which assess the transition from thermionic emission to SCLC,27,28 with recent theories and simulations extending this work to account for the influence of nonuniform emitters.29–32 We considered the simple case of emission in a planar diode and applied the JGTF equation to directly unify FN, RLD, and CL; we further used the asymptotic solutions for MG and OL to link all five mechanisms with nexus theory.8 Using an analogous approach, we directly unified quantum CL, nonquantum (standard) CL, MG, and FN and indirectly linked these mechanisms with field emission driven breakdown and Paschen's law.11 We are currently performing nanoscale (gaps between ∼20 and 800 nm) experiments at atmospheric pressure to assess the transitions between these mechanisms.4 

One important electron emission mechanism that we have not yet incorporated into nexus theory is photoemission (PE), where electron emission occurs due to the photoelectric effect, typically due to laser exposure.2,33 This leads to the Fowler–DuBridge (FD) equation,2,34 given by

(7)

where QE is the photoyield or quantum efficiency, ω is the photon energy (product of and photon frequency ω), JFD is the FD current density, and Iω is the laser intensity. Due to the ability to generate high brightness, low emittance, and spatiotemporally coherent electron bunches, photoemission is critical for multiple applications, including time-resolved electron microscopy,35,36 free-electron lasers,37 carrier-envelope-phase detection,38 and novel nanoelectronic devices.39–42 

We rewrite (7) as43–46 

(8)

where F is the cathode electric force (F=eE, where E is the electric field); ω=2πc/λ is the angular frequency; λ is the laser wavelength; c is the speed of light in vacuum;

(9)

is the emission probability, where ζ(2)π2/6 is the Riemann zeta function, φ=Φ4QF is the reduced work function, Q=e2/(16πϵ0); and

(10)

is the effective laser intensity, where Ii is the initial laser intensity in units of Wm2, Fλ quantifies the scattering, and R(θ) is the reflectivity of a laser with incident angle θ.

Photoemission devices may undergo transitions between photoemission and other mechanisms over a range of physical parameters. In practical photocathodes, local electric fields may be higher than the applied electric field due to surface imperfections increasing field enhancement, which causes field emission through the reduced photoelectric barrier.47 This combined field and photoemission regime may also be applied practically by using laser intensities of ∼1 W cm−2 with a bias voltage <10 V to activate a semiconductor-free device whose small dimensions would permit integration with semiconductor devices.41 Additionally, one may leverage thermionic emission of photoexcited electrons from a semiconductor cathode at high temperature to operate in a combined thermionic and photoemission regime to exceed the theoretical limits of single-junction photovoltaic cells.48 

As for thermionic and field emitters, extracting the maximum current from a photoemitter, or a device that operates with a combination of thermionic, field, and/or photoemission, ultimately requires including space-charge. As discussed in detail in multiple reviews and references therein,1–4 fully accounting for the transitions between the multiple emission regimes becomes challenging, motivating the development of nexus theory. In the case of photoemission, while several theoretical studies have linked various combinations of thermionic, field, and photoemission,24,43–46,49–55 the connection between photoemission and space-charge limited emission (especially MG and CL simultaneously) has been less thoroughly studied.3 

Thus, this paper seeks to unify CL, MG, FD, RLD, and FN in a single theoretical framework, as shown schematically in Fig. 1. In the process, we will demonstrate a consistent, step-by-step process for unifying mechanisms to develop general nexus curves for given conditions. This will provide a procedure for constructing nexus curves based on the specific conditions under consideration to determine which mechanism(s) and theory(s) must be considered. Note that we use the general thermal-field-photoemission (GTFP) current density24,25 for our canonical current, adding photoemission to our prior nexus theory analysis,4,8 so we do not explicitly study time-dependent phenomena here.3 Upon understanding the relevant regime of operation, one may either use the full theory outlined in Sec. II or, more realistically, an appropriate simplified theory from the literature that considers one, two, or three mechanisms depending on the proximity of desired parameters relative to relevant nexus curves.

FIG. 1.

Circuit diagram demonstrating a gas-filled diode exposed to an applied DC voltage V with an external series resistor R undergoing PE, TE, and FE.

FIG. 1.

Circuit diagram demonstrating a gas-filled diode exposed to an applied DC voltage V with an external series resistor R undergoing PE, TE, and FE.

Close modal

Section II outlines the derivation of the exact theory unifying CL, MG, FD, RLD, and FE and the process for deriving nexus curves unifying two or more of these mechanisms. Section III details the process for constructing nexus curves as a function of different variables to demonstrate the conditions for various transitions. We provide concluding remarks in Sec. IV.

As in prior theoretical studies with planar diodes, we consider the motion of a single electron emitted from a grounded cathode with potential ϕ=0 at x=0 and an anode at x=D biased to ϕ=V. We can then write Poisson's equation, conservation of energy, and electron continuity as

(11)
(12)

and

(13)

respectively, where J is the current density, ρ is the charge density, v is electron velocity, and νi is the initial electron velocity.

Prior nexus theories have defined J as either JFN or JGTF when unifying space-charge limited emission with field emission9,10,12,21 or thermo-field emission,8 respectively. Since this study seeks to unify space-charge limited emission, field emission, thermionic emission, and photoemission, we define J as the GTFP current density,24,25 given by

(14)

where F is the cathode electric force (F=eE, where E is the electric field), JFD(F,T) is the Fowler–DuBridge equation for generalized photoemission given by (8)–(10), and

(15)

where ARLD=(emkb2)/(2π23) and

(16)

where nβT/βF, s=βF(Em)(ωφ), Q0.36eVnm,2,37 and Em=μF+φ, where μF is the Fermi energy; this model is valid provided that the emission remains below the barrier energy.26,52 Equation (16) accounts for the different regimes where field and thermionic emission occur.

The GTF equation becomes the GTFP equation when N(n,s) goes to N(n,s),2,24,52–55 yielding

(17)

where u=βF(Em)(μF+φω) and ns=βT(ωφ) for PE.26 Similar to R(θ) in (10), the scattering factor Fλ depends on various factors, including the incident angle; however, Fλ is calculated using a different analytical approach.36 Because of the complexity of R(θ) and Fλ due to their dependence on material-specific parameters, we will treat them as constants for this analysis. Performing a matched asymptotic analysis for nexus theory requires determining the appropriate limits for the transitions from photoemission. The emission mechanisms transition in JGTFP based on the values of βT, βF, n, and s.

We note that this analysis is a first step in unifying these multiple mechanisms and, as such, we do not attempt to use the most complete theory of photoemission. This is analogous to our recent work incorporating quantum space-charge phenomena where we used an early model of quantum SCLC to demonstrate nexus theory.11 The present study focuses on traditional photoemission due to multiphoton absorption at the large Keldysh parameter γ1, which corresponds to low laser power where the FD and GTFP theories are acceptable. Higher laser powers can cause electron tunneling through the time varying barrier and lead to optical tunneling. At its core level, this paper focuses on unifying the GTFP with SCLC by using a simpler version of photoemission that neglects optical tunneling. As such, it does not specifically address conditions at the frontier of photoemission research where 1γ10, where both multiphoton and optical field emission are important.56,57 Moreover, our approach does not directly account for laser-induced heating of the material, which is time-dependent and nonequilibrium during photoemission; this can also make the two-temperature assumption inaccurate.58 As we have demonstrated through prior studies and in subsequent sections in this paper, the concept of nexus theory may be modified and expanded to include additional mechanisms or more details of a given mechanism. Specifically, this paper provides a primer for developing nexus curves by either incorporating additional mechanisms or adding details within given mechanisms.

To eliminate physical parameters and more intuitively assess asymptotic behavior, we define the following nondimensionalized parameters:

(18)

with physical scaling parameters (with units in square brackets)

(19)

where S is emission area, R is resistance, and the dimensions are denoted in square brackets. Table II summarizes the values of the scaling parameters defined in (19).

TABLE II.

Calculated scaling parameters used in nondimensionalization.

Scaling constantValue
E0 6.5207 × 1010 V m−1 
J0 1.4565 × 1015 A m−2 
t0 3.9641 × 10−16 s 
T0 5.22 × 104 K 
F0 1.0447 × 10−8 eV m−1 
x0 1.8022 × 10−9 m 
ν0 4.5463 × 106 m s−1 
ϕ0 117.5152 V 
μ0 5.1346 × 1017 m2 V−1 s−1 
R0 806.8 Ω 
I0 2.2980 × 10−15 W m−2 
ω0 2.5227 × 1015 s−1 
Scaling constantValue
E0 6.5207 × 1010 V m−1 
J0 1.4565 × 1015 A m−2 
t0 3.9641 × 10−16 s 
T0 5.22 × 104 K 
F0 1.0447 × 10−8 eV m−1 
x0 1.8022 × 10−9 m 
ν0 4.5463 × 106 m s−1 
ϕ0 117.5152 V 
μ0 5.1346 × 1017 m2 V−1 s−1 
R0 806.8 Ω 
I0 2.2980 × 10−15 W m−2 
ω0 2.5227 × 1015 s−1 

Using these values yields the same asymptotic limits derived previously for FN, RLD, CL, generalized CL (GCL) (which models SCLC with nonzero injection velocity), MG, and OL as8–10 

(20)
(21)
(22)
(23)
(24)

and

(25)

respectively.

Next, we define the relevant asymptotic solution for FD to develop nexus phase space plots assessing transitions between these emission mechanisms and demonstrate the transitions between the exact solution from (8) to (17) above to the asymptotic solutions under appropriate limits. We obtain the asymptotic solution for FD when F0 and T0. Nondimensionalizing (8) yields

(26)

with χ1=em/(AFNBFN2)=9.5×1031, χ2=AFNE0/(ϵ0Φ)=0.3690, χ3=e3BFN/(4πϵ0Φ2)=4.6368, and χ4=μFΦ1=1.5556 for the parameters in Table I. When T=0, (26) becomes

(27)

when F=0, (26) becomes

(28)

and when F=0 and T=0, (26) becomes

(29)

Note that (26)–(29) are not universal (i.e., independent of material) since χ2,χ3, and χ4 all depend on the work function Φ. Since (29) eliminates the temperature and field dependence and just retains the photon dependence, we use (29) as the asymptotic solution for the FD equation when developing nexus curves in subsequent sections. When assessing J¯ vs V¯ curves, we use the more complete version of FD from (26) to demonstrate the full, analytic transitions among FD, FN, and RLD since J¯GTFP from (14) includes the effects of E¯ and T¯ upon photoemission.

The process to obtain the exact solution is identical to that reported for our prior unification of thermionic, field, and space-charge limited emission8 except that we replace J¯GTF with J¯GTFP, which gives

(30)

where ξ=1+[1+(2J¯/E¯2)((2V¯+v¯i2)1/2v¯i)]1/2 and

(31)

where we use (26) for J¯FD(F¯,T¯) in the exact solutions (30) and (31).

As described briefly above and in detail elsewhere,4 setting two or more of the asymptotic solutions for current density equal yields nexuses representing the transition in dominance from one mechanism to another. A condition where two asymptotic solutions match is a second-order nexus; equating additional asymptotic solutions yields higher-order nexus points.

To begin, we write the conditions for the various second-order nexus curves by equating the various combinations of (20)–(25) and (29) as follows (the ones independent of FD have been derived previously):

(32)

where the subscript N denotes a nexus quantity, for MG to CL,

(33)

for MG to FN,

(34)

for MG to OL,

(35)

for MG to RLD,

(36)

for FN to CL,

(37)

for FN to RLD,

(38)

for FN to OL,

(39)

for CL to RLD,

(40)

for CL to OL,

(41)

for RLD to OL,

(42)

for FD to FN,

(43)

for FD to RLD,

(44)

for FD to CL,

(45)

for FD to MG, and

(46)

for FD to OL.

We may obtain higher-order nexuses by setting appropriate combinations of (32)–(46) equal. For instance, the third-order nexus between RLD, FN, and FD comes from equating the second-order nexuses between RLD and FN from (37) and FN and FD from (42) [or, alternatively, by matching their asymptotic solutions from (20), (21), and (29), or using the nexus between RLD and FD from (43) with (37) or (42)] to obtain

(47)

Note that (47) depends on D¯N/V¯N, not on V¯N and D¯N individually (i.e., V¯N and D¯N are linked and defining one defines the other), and is not universal (true for any material) since χ2 is material dependent. Due to the numerous combinations available for the number of mechanisms studied here, we do not perform these calculations for all potential third-order and higher nexuses; however, we have demonstrated the straightforward use of this process and discussed the many options available to arrive at additional third-order and higher nexuses.

This leads to a critical point concerning nexus theory. Matching the various asymptotic solutions invariably violates the assumptions for the individual asymptotic solutions. For instance, the asymptotic solution for FN assumes negligible space-charge, meaning that E=V/D; however, EV/D for the CL equation due to the presence of space-charge, so the CL and FN asymptotic relations must both be invalid at the nexus. Accurately determining the behavior near the nexus requires a theory combining both mechanisms.12 A device operating with parameters within one to two orders of magnitude of a nexus most likely requires using a theory combining multiple mechanisms, while parameters further from a nexus typically can be modeled asymptotically. In other words, nexuses can identify by proximity the mechanisms mostly likely to be relevant for a particular device. Adding phenomena makes deriving complete, exact solutions from first principles more tedious; however, the nexus process remains straightforward. For instance, adding MG, CL, and OL to (47) yields a sixth-order nexus between FD, RLD, FN, CL, MG, and OL, given by

(48)

The most critical parameters to determine when assessing the behavior for emission mechanism transitions across multiple regimes are the nexus points. Since we introduce photoemission to our prior nexus theory studies,4,8–10 we focus here on the third-order nexus between the emission source current densities from FD, RLD, and FN from (47). Figure 2 shows the resulting third-order nexus plot as a function of E¯=V¯/D¯,ω¯=λ¯1, and T¯ where defining one of these three parameters uniquely defines the other two. Note that because the asymptotic solutions used to obtain (47) assume no space-charge, the third-order nexus point depends on E¯ and not separately on V¯ and D¯, which means that defining either V¯ or D¯ uniquely determines the other parameter to achieve this order nexus point. Figure 2 provides important guidance for characterizing the mechanisms involved with changing parameters, much as the third-order nexus when coupling FN, MG, and CL determined when electron emission would either encounter or bypass the MG regime when transitioning from FN to CL with increasing V¯ or decreasing D¯.9 As we shall see, depending on the relationship with the third-order nexus, electron emission may transition from FD to RLD before FN or from FD to FN directly depending on the parameters relative to the third-order nexus conditions.

FIG. 2.

Third-order nexus between FD, FN, and RLD from (47) for I¯1=3.4813×1023(8×108Wm2),I¯2=3.4813×1025(8×1010Wm2),I¯3=3.4813×1027(8×1012Wm2) and work function Φ=4.5eV.

FIG. 2.

Third-order nexus between FD, FN, and RLD from (47) for I¯1=3.4813×1023(8×108Wm2),I¯2=3.4813×1025(8×1010Wm2),I¯3=3.4813×1027(8×1012Wm2) and work function Φ=4.5eV.

Close modal

As in our prior analyses,8–11 we next examine the full solution of J¯ as a function of V¯ under various conditions to demonstrate the transition between electron emission mechanisms with increasing V¯ by comparing to the relevant asymptotic solutions. Figure 3 shows the initial effect of changing T¯ relative to T¯N, the temperature representing the third-order nexus between FD, RLD, and FN [cf. (47) and Fig. 2], on J¯ as a function of V¯ with no series resistor with D¯=0.5549(1nm), ω¯=0.7018(1.7703×1015s1), λ¯=591(λ=1064×109m), and effective intensity I¯=3.4813×1025(8×1010Wm2) for T¯=0.03T¯N=0.0016(82K), 0.15T¯N=0.0392(2047K), T¯N=0.0523(2729K), and 1.5T¯N=0.0784(4094K). For all T¯, the exact solution follows FD at low V¯ since the input energy from the laser is sufficiently high for photoemission to dominate over either field or thermionic emission at these low electric fields. For T¯<T¯N, the exact solution next follows FN rather than RLD since field emission dominates over thermionic emission; ultimately, emission becomes space-charge limited and follows GCL at higher voltages. For T¯=T¯N, FD, RLD, and FN meet at the third-order nexus, and the exact solution undergoes a transition region after following FD and before following FN that follows neither asymptote. For T¯=1.5T¯N, the exact solution now transitions from FD to RLD to FN to GCL with increasing V¯. As observed previously when considering FN, RLD, and GCL,9 the full solution tracks with the highest J¯ from the emission laws until reaching GCL.

FIG. 3.

Dimensionless current density J¯ as a function of dimensionless gap voltage V¯ at fixed dimensionless gap distance D¯=0.5549(1nm), ω¯=0.7018(1.7703×1015s1), λ¯=591(λ=1064nm), and effective intensity I¯=3.4813×1025(8×1010Wm2) for various nondimensional temperatures: (a) T¯=0.03T¯N=0.0016(82K), (b) 0.15T¯N=0.0392(2047K), (c) T¯N=0.0523(2729K), and (d) 1.5T¯N=0.0784(4094K), where T¯N corresponds to the third-order nexus between FN, RLD, and FD from (47). The asymptotic photoemission (FD), field emission (FN), thermionic emission (RLD), and space-charge limited emission (GCL) limits are calculated from (26), (20), (21), (22), and (23), respectively, and compared to the exact solution (30) and (31) (red boxes). The exact solution follows the highest J¯ from the relevant canonical current density (FD, FN, or RLD) until reaching the space-charge emission limit (GCL) at high V¯.

FIG. 3.

Dimensionless current density J¯ as a function of dimensionless gap voltage V¯ at fixed dimensionless gap distance D¯=0.5549(1nm), ω¯=0.7018(1.7703×1015s1), λ¯=591(λ=1064nm), and effective intensity I¯=3.4813×1025(8×1010Wm2) for various nondimensional temperatures: (a) T¯=0.03T¯N=0.0016(82K), (b) 0.15T¯N=0.0392(2047K), (c) T¯N=0.0523(2729K), and (d) 1.5T¯N=0.0784(4094K), where T¯N corresponds to the third-order nexus between FN, RLD, and FD from (47). The asymptotic photoemission (FD), field emission (FN), thermionic emission (RLD), and space-charge limited emission (GCL) limits are calculated from (26), (20), (21), (22), and (23), respectively, and compared to the exact solution (30) and (31) (red boxes). The exact solution follows the highest J¯ from the relevant canonical current density (FD, FN, or RLD) until reaching the space-charge emission limit (GCL) at high V¯.

Close modal

Figure 4 shows the initial effect of changing λ¯ relative to λ¯N, representing the wavelength corresponding to the third-order nexus between FN, RLD, and FD from (47), on J¯ as a function of V¯ with no series resistor for D¯=0.5549(1nm), T¯=0.0523, and effective intensity I¯=3.4813×1025(8×1010Wm2) for λ¯=0.05λ¯N, λ¯N, 50λ¯N, and 500λ¯N, where λ¯N=416 (750 nm). As for Fig. 3 with FN, the exact solution initially follows FD at low V¯ since J¯FD is higher than both J¯RLD and J¯FN. With increasing V¯, electron emission transitions to FN and GCL in all cases independent of λ¯. For λ¯>0.05λ¯N in Fig. 4(a), electron emission never follows RLD. Prior to reaching FN, the exact solution appears to undergo a transition from FD to RLD; however, the exact solution begins to approach the FN asymptote before reaching RLD, so the exact solution never follows RLD. Changing device operating conditions, most notably T¯, would most likely make the transition to RLD more distinct. Figure 4(a) highlights the importance of characterizing the contribution of multiple emission mechanisms. Although emission never follows RLD in Fig. 4(a), thermionic emission does contribute to the emission between the FD and FN regimes, which would necessitate the use of the exact solution. For λ¯λ¯N in Figs. 4(b)4(d), thermionic emission is completely bypassed and emission transitions from FD directly to FN with transition occurring more rapidly (i.e., a smaller transition region in D¯ requiring the exact solution) with increasing λ¯ beyond λ¯N. Again, the exact solution follows the emission law (FD, RLD, or FN) giving the highest J¯ until emission is space-charge limited (GCL).

FIG. 4.

Dimensionless current density J¯ as a function of dimensionless gap voltage V¯ at fixed dimensionless gap distance D¯=0.5549(1nm), T¯=0.0523, and effective intensity I¯=3.4813×1025(8×1010Wm2) for various nondimensional wavelengths: (a) λ¯=0.05λ¯N, (b) λ¯N, (c) 50λ¯N, and (d) 500λ¯N, where λ¯N=416 (750 nm) corresponds to the third-order nexus between FN, RLD, and FD from (47). The asymptotic limits for photoemission (FD), field emission (FN), thermionic emission (RLD), and space-charge limited current (GCL) are calculated from (26), (20), (21), (22), and (23), respectively, and compared to the exact solution (30) and 31) (boxes). The exact solution follows the highest J¯ from the relevant canonical current density (FD, FN, or RLD) until reaching the space-charge emission limit (GCL) at high V¯.

FIG. 4.

Dimensionless current density J¯ as a function of dimensionless gap voltage V¯ at fixed dimensionless gap distance D¯=0.5549(1nm), T¯=0.0523, and effective intensity I¯=3.4813×1025(8×1010Wm2) for various nondimensional wavelengths: (a) λ¯=0.05λ¯N, (b) λ¯N, (c) 50λ¯N, and (d) 500λ¯N, where λ¯N=416 (750 nm) corresponds to the third-order nexus between FN, RLD, and FD from (47). The asymptotic limits for photoemission (FD), field emission (FN), thermionic emission (RLD), and space-charge limited current (GCL) are calculated from (26), (20), (21), (22), and (23), respectively, and compared to the exact solution (30) and 31) (boxes). The exact solution follows the highest J¯ from the relevant canonical current density (FD, FN, or RLD) until reaching the space-charge emission limit (GCL) at high V¯.

Close modal

We next construct nexus curves highlighting the transitions between the different emission mechanisms by matching the solutions for the different canonical currents, SCLC, and OL, which arise under appropriate asymptotic conditions4 as we have outlined previously for the following conditions: (a) FN, MG, and CL;9 (b) FN, MG, CL, and OL;10 (c) FN, MG, CL, RLD, and OL.8 While we have previously developed these nexus curves, we have not specifically reported the step-by-step process used to generate them. This is straightforward when only considering a second-order nexus (e.g., FN to CL),12 which requires only matching two asymptotic solutions; however, this becomes increasingly challenging with additional mechanisms. Furthermore, as Refs. 8–10 and 12 demonstrate, adding mechanisms also requires considering more parameters for nexus development. For instance, FN to CL12 requires only considering D¯ and V¯. Incorporating MG (Ref. 9) requires also considering μ¯, while including OL (Ref. 10) and RLD (Ref. 8) requires adding R¯ and T¯, respectively. Incorporating photoemission requires including λ¯, which we mathematically incorporate by studying ω¯ in dimensionless units (while we could more precisely consider λ¯ since most photoemission literature considers wavelength rather than angular frequency, the mathematical analysis is more convenient with ω¯, and since ω¯=λ¯1, the transformation between the two is straightforward). Thus, Figs. 5–8 consider V¯ as a function of D¯,μ¯,T¯,ω¯ at low I¯ and ω¯ at high I¯. For each figure, we provide the step-by-step procedure for developing the nexus curves by starting from a single second-order nexus and adding other second-order nexuses to form the full plot, which “slices” the phase space into regimes of primary influence for each mechanism. Each panel introduces a new nexus curve as a dashed curve, which is replaced by a solid curve in subsequent panels. In particular, we emphasize the slicing of phase space into the various emission regimes as we incorporate additional nexuses since adding a new mechanism will often introduce nexuses with other existing mechanisms and not just one. This slicing of multiple regimes is what introduces additional complexity in adding more mechanisms to the nexus plots and motivates us to detail the philosophy and procedure of plot formation in the following figures.

FIG. 5.

Development of the nexus plot of V¯ as a function of D¯ between FD, FN, RLD, MG, CL, and OL, where T¯=7.5×102(3915K), I¯=3.4813×1025(8×1010Wm2), ω¯=0.7018(1.7703×1015s1),F¯=1(9.5718×107eV/m),R¯=0.01(8.07Ω), and μ¯=0.05(9.7378×1020m2V1s1) by introducing second-order nexuses in each panel as dashed lines that may divide previously defined mechanisms into smaller regions. (a) Start with the second-order nexus between CL and MG from (32) to define the CL and MG regimes. (b) Create the FN regime by including the nexuses of FN with CL (36) and MG (33). (c) Incorporate the RLD regime by adding the nexuses of RLD with MG (35) and FN (37). Since JFD>JRLD for these parameters, we omit the RLD regime in subsequent panels. (d) Incorporate the FD region by adding the nexuses of FD with FN (42) and MG (45). (e) Incorporate the OL regime by including the nexuses between OL with CL (40) and FN (38). Note that we omit the nexus between OL and MG here since this device conditions here yield a fourth-order nexus between OL, CL, MG, and FN.

FIG. 5.

Development of the nexus plot of V¯ as a function of D¯ between FD, FN, RLD, MG, CL, and OL, where T¯=7.5×102(3915K), I¯=3.4813×1025(8×1010Wm2), ω¯=0.7018(1.7703×1015s1),F¯=1(9.5718×107eV/m),R¯=0.01(8.07Ω), and μ¯=0.05(9.7378×1020m2V1s1) by introducing second-order nexuses in each panel as dashed lines that may divide previously defined mechanisms into smaller regions. (a) Start with the second-order nexus between CL and MG from (32) to define the CL and MG regimes. (b) Create the FN regime by including the nexuses of FN with CL (36) and MG (33). (c) Incorporate the RLD regime by adding the nexuses of RLD with MG (35) and FN (37). Since JFD>JRLD for these parameters, we omit the RLD regime in subsequent panels. (d) Incorporate the FD region by adding the nexuses of FD with FN (42) and MG (45). (e) Incorporate the OL regime by including the nexuses between OL with CL (40) and FN (38). Note that we omit the nexus between OL and MG here since this device conditions here yield a fourth-order nexus between OL, CL, MG, and FN.

Close modal
FIG. 6.

Development of the nexus plot of V¯ as a function of μ¯ between FD, FN, MG, CL, and OL, where I¯=1×105(2.30×1020Wm2), D¯=1×107(1.80×102m), ω¯=0.7018(1.7703×1015s1),F¯=1(9.5718×107eV/m), and R¯=1×109(8.07×1011Ω) by introducing second-order nexuses in each panel as dashed lines that may divide previously defined mechanisms into smaller regions. (a) Start with the second-order nexus between FN and FD from (42). (b) Add the second-order nexus between FN and MG from (33). (c) Incorporate the CL regime by adding the second-order nexuses of CL with MG (32) and FN (36). (d) Incorporate the OL regime by adding second-order nexuses of OL with CL (40) and MG (34). As in Fig. 5, JFD>JRLD for these parameters, so we omit the RLD regime in our analysis.

FIG. 6.

Development of the nexus plot of V¯ as a function of μ¯ between FD, FN, MG, CL, and OL, where I¯=1×105(2.30×1020Wm2), D¯=1×107(1.80×102m), ω¯=0.7018(1.7703×1015s1),F¯=1(9.5718×107eV/m), and R¯=1×109(8.07×1011Ω) by introducing second-order nexuses in each panel as dashed lines that may divide previously defined mechanisms into smaller regions. (a) Start with the second-order nexus between FN and FD from (42). (b) Add the second-order nexus between FN and MG from (33). (c) Incorporate the CL regime by adding the second-order nexuses of CL with MG (32) and FN (36). (d) Incorporate the OL regime by adding second-order nexuses of OL with CL (40) and MG (34). As in Fig. 5, JFD>JRLD for these parameters, so we omit the RLD regime in our analysis.

Close modal
FIG. 7.

Development of the nexus plot of V¯ as a function of T¯ between FD, FN, RLD, MG, CL, and OL, where ω¯=3.8769×109(9.78×106s1),I¯=1×1014(4.35×1029Wm2),D¯=1×108(1.8×101m),F¯=6×108(2.3929×109eV/m),R¯=1×1011(8.07×1013Ω), and μ¯=1000(1.95×1015m2V1s1) by introducing second-order nexuses in each panel as dashed lines that may divide previously defined mechanisms into smaller regions. (a) Start with the second-order nexus between FN and MG (33). (b) Add the second-order nexus between CL and MG (32). (c) Add the second-order nexus between CL and OL (40). (d) Incorporate the RLD regime by adding second-order nexuses of RLD with OL (41), CL (39), MG (35), and FN (37). (e) Incorporate the FD regime by adding second-order nexuses of FD with FN (42) and RLD (43). The second-order nexus between FD and RLD is vertical since it is independent of V¯.

FIG. 7.

Development of the nexus plot of V¯ as a function of T¯ between FD, FN, RLD, MG, CL, and OL, where ω¯=3.8769×109(9.78×106s1),I¯=1×1014(4.35×1029Wm2),D¯=1×108(1.8×101m),F¯=6×108(2.3929×109eV/m),R¯=1×1011(8.07×1013Ω), and μ¯=1000(1.95×1015m2V1s1) by introducing second-order nexuses in each panel as dashed lines that may divide previously defined mechanisms into smaller regions. (a) Start with the second-order nexus between FN and MG (33). (b) Add the second-order nexus between CL and MG (32). (c) Add the second-order nexus between CL and OL (40). (d) Incorporate the RLD regime by adding second-order nexuses of RLD with OL (41), CL (39), MG (35), and FN (37). (e) Incorporate the FD regime by adding second-order nexuses of FD with FN (42) and RLD (43). The second-order nexus between FD and RLD is vertical since it is independent of V¯.

Close modal
FIG. 8.

Development of the nexus plot of V¯ as a function of ω¯ between FD, FN, RLD, MG, and CL, where T¯=1.5×104(7.83K),I¯=3.4813×1025(8×1010Wm2),D¯=0.9(1.62×109m),F¯=25(2.3929×109eV/m), and μ¯=0.084(1.636×1019m2V1s1) by introducing second-order nexuses in each panel as dashed lines that may divide previously defined mechanisms into smaller regions. (a) Start with the second-order nexus between FN and MG (33). (b) Add the second-order nexus between CL and MG (32). (c) Add the second-order nexus between RLD and FN (37). (d) Add the FD regime by adding second-order nexuses of FD with FN (42), MG (45), and CL (44)—note that RLD is subsumed at this higher intensity and no longer has any defined regime in this phase plot.

FIG. 8.

Development of the nexus plot of V¯ as a function of ω¯ between FD, FN, RLD, MG, and CL, where T¯=1.5×104(7.83K),I¯=3.4813×1025(8×1010Wm2),D¯=0.9(1.62×109m),F¯=25(2.3929×109eV/m), and μ¯=0.084(1.636×1019m2V1s1) by introducing second-order nexuses in each panel as dashed lines that may divide previously defined mechanisms into smaller regions. (a) Start with the second-order nexus between FN and MG (33). (b) Add the second-order nexus between CL and MG (32). (c) Add the second-order nexus between RLD and FN (37). (d) Add the FD regime by adding second-order nexuses of FD with FN (42), MG (45), and CL (44)—note that RLD is subsumed at this higher intensity and no longer has any defined regime in this phase plot.

Close modal

First, Fig. 5 describes the step-by-step process for constructing a nexus plot with V¯ as a function of D¯ for T¯=7.5×102 (3915 K), I¯=3.4813×1025(8×1010Wm2), ω¯=0.7018(1.7703×1015s1),F¯=1(9.5718×107eV/m),R¯=1×105(8.07×107Ω), and μ¯=0.05(9.7378×1020m2V1s1). While |I| in dimensional units for some later figures may be insignificant compared to the value used here, we note the conversion from I¯ to I is material dependent, and the choice of I¯ for each plot is motivated to maximize the number of emission mechanism transitions to demonstrate the nexus phase space plot generation process, irrespective of the particular physical I.

Figure 5(a) starts with the second-order nexus between CL and MG from (32); CL dominates at high V¯ since J¯CLV¯3/2 in (22) scales slower than J¯MGV¯2 in (24), and the space-charge limited condition with the lower J¯ determines this limiting behavior. As discussed earlier when developing Figs. 3 and 4, the dominant nonlimiting emission condition is the one with highest J¯. Thus, the general principle is that the emission mechanism with the highest J¯ dominates until the J¯ for each emission mechanism under consideration exceeds J¯ for the lowest space-charge limited condition. We then add FN to this plot, analogous to our earlier study;9 however, because FN intersects with both CL and MG in Fig. 5(b), we must include both the second-order nexuses of FN with CL and MG from (36) and (33), respectively. In the process of uniting these second-order nexuses, we also create a third-order nexus between CL, MG, and FN.9 Since neither FD nor RLD depend on V¯ or D¯, only one of them may appear in this nexus plot; for the conditions chosen, J¯FD>J¯RLD, so we will ultimately incorporate FD rather than RLD. However, we include the second-order nexuses of RLD with FN and MG from (37) and (35), respectively, in Fig. 5(c) to show how the RLD regime would appear if J¯FD were lower than J¯RLD for reference. This emphasizes the six-dimensional nature of this iteration of nexus theory even if RLD does not appear on the V¯D¯ phase plot; if the operating temperature is close to a nexus (which will be elucidated by a V¯T¯ plot shortly), RLD must be considered when predicting behavior. For the present case, we add the second-order nexus between FD and FN from (42), which eliminates the RLD regime introduced in Fig. 5(c), to obtain Fig. 5(d). Finally, in Fig. 5(e), we incorporate OL by starting from the CL to OL nexus from (40) since we have previously shown the transition from CL to OL (Ref. 10) with increasing V¯ or decreasing D¯. Because of how OL intersects with CL, we also know that reducing D¯ will require including the second-order nexus of OL with FN from (38) to complete the nexus plot in Fig. 5(e). Note that because OL, CL, FN, and MG intersect in a fourth-order nexus, we do not include the nexus between OL and MG; this second-order nexus between OL and MG from (34) must be considered under general conditions.

One common issue with nexus plots is that not every transition for a given plot is necessarily physically relevant for a given set of physical conditions;9 however, it provides a framework to assess the most relevant mechanisms across a broad range of potential conditions to guide the selection of the appropriate theory or combination of theories for assessing a given condition. As such, Fig. 5 provides a comprehensive guide to developing such a nexus plot.

Figure 6 demonstrates the construction of a nexus plot for V¯ as a function of μ¯ for I¯=1×105(2.30×1020Wm2), D¯=1×107(1.80×102m), ω¯=0.7018(1.7703×1015s1),F¯=1(9.5718×107eV/m), and R¯=1×109(8.07×1011Ω) by incorporating second-order nexuses in a step-by-step process. We also note that not every mechanism under consideration here depends on μ¯; in fact, only MG does. Thus, when constructing second-order (or higher) nexuses for mechanisms without μ¯ dependence, the resulting nexus curve will be a horizontal line with constant V¯. This will also hold for certain electron emissions as functions of ω¯ and T¯ in Figs. 7 and 8, respectively. For instance, the second-order nexus between FN and FD, given by (42), does not depend on μ¯, resulting in a constant V¯ in Fig. 6(a). We next introduce the MG to FN nexus from (33) in Fig. 6(b) since that is the next emission transition with increasing V¯, noting that this nexus varies with μ¯ due to the μ¯ dependence of MG. As in Fig. 5, we neglect any nexuses with RLD here since it is subsumed by the FD regime because JFD>JRLD for these conditions. Further increasing V¯ leads to the MG to CL nexus from (32) in Fig. 6(c). Introducing the CL regime also requires including the CL to FN nexus from (36), represented by the second dashed horizontal line in Fig. 6(c) since neither CL nor FN depend on μ¯. Finally, further increasing V¯ leads to the transition from CL to OL, given by (40), which we introduce in Fig. 6(d). To finalize the OL regime, we must also introduce the second-order nexus between OL and MG from (34).

We construct a nexus plot for V¯ as a function of T¯ for ω¯=3.8769×109(9.78×106s1),I¯=1×1014(4.35×1029Wm2),D¯=1×108(1.8×101m),F¯=6×108(2.3929×109eV/m),R¯=1×1011(8.07×1013Ω), and μ¯=1000(1.95×1015m2V1s1) by incorporating second-order nexuses in a step-by-step process in Fig. 7. In this case, only nexuses with RLD, which depends on T¯, exhibit nonconstant V¯. Figure 7(a) shows the nexus between FN and MG from (33). As before, increasing V¯ causes the transition from MG to CL, given by (32) and included as a nexus curve in Fig. 7(b). Figure 7(c) incorporates the nexus between CL and OL from (40), which occurs with further increasing V¯. Electron emission transitions to RLD as T¯ increases, so Fig. 7(d) shows the second-order nexuses of RLD with FN, MG, CL, and OL, given by (37), (35), (39), and (41), respectively, at high T¯. Note that for a given T¯, a sufficiently high V¯ will ultimately cause a transition to MG, CL, or OL. Finally, we incorporate FD in Fig. 7(e) by adding second-order nexuses of FD with FN and RLD, given by (42) and (43), respectively. The nexus between FD and RLD is vertical since it is independent of V¯ but not T¯, while the nexus between FD and FN is horizontal since it is independent of T¯ but not V¯.

Figure 8 demonstrates the construction of a nexus plot for V¯ as a function of ω¯ for T¯=1.5×104(7.83K),I¯=3.4813×1025(8×1010Wm2),D¯=0.9(1.62×109m),F¯=25(2.3929×109eV/m), and μ¯=0.084(1.636×1019m2V1s1) by incorporating second-order nexuses in a step-by-step process. In this case, only nexuses with FD, which depends on ω¯, exhibit nonconstant V¯. Figure 8(a) starts with the FN/MG second-order nexus from (33). Further increasing V¯ leads to the transition from MG to CL given by (32) and leading to the additional nexus curve in Fig. 8(b). Figure 8(c) shows the second-order nexus between FN and RLD from (37). Introducing the second-order nexuses of FD with OL, CL, MG, and FN, given by (46), (44), (45), and (42), respectively, leads to the nonconstant V¯ nexus curves in Fig. 8(d). Note that the RLD regime vanishes since JFD>JRLD for these conditions. We omit the OL to CL nexus from (40) because it will simply be another horizontal line since it does not depend on ω¯ [although it would also introduce an OL to FD nexus from (46)] and will add clutter to the nexus phase plot while providing limited additional information. Depending on the operational conditions for a given device, it is straightforward to incorporate these additional nexus curves onto a phase space plot.

We have derived an exact theory that includes photoemission (FD), thermionic emission (RLD), field demission (FN), and vacuum space-charge limited emission (CL) and used nexus theory to incorporate space-charge limited emission with collisions (MG) and an external resistor (OL). Nexus theory can be adapted to include a subset or all of the mechanisms under consideration depending on device conditions. Moreover, nexus theory provides a standard process for determining the applicability of the standard emission or circuit laws under consideration (e.g., OL, CL, MG, FN, RLD, or FD) or more complicated theories unifying multiple mechanisms.

As a tool for assessing the appropriate solutions to use, we have described a step-by-step process for constructing nexus phase space plots demonstrating the transitions between these various mechanisms by matching individual asymptotic equations, or second-order nexuses, in a step-by-step process. While simple conceptually, linking so many equations may initially appear overwhelming without having a simple process based on the variation of the dominant mechanisms with changing parameters, such as D, μ, ω, and T. To maximize mathematical simplicity and model generality, we used nondimensional units in our derivations that may easily be converted to physical parameters by using the definitions in (18) and (19) for the conditions under consideration.

As noted previously,8 a given nexus plot may not be physically relevant for all possible transitions; the goal of nexus theory is to provide a simple, consistent framework to assess the physically relevant parameters for a given condition. For instance, a subset of emission mechanism transitions from one of the nexus plots developed in this study may be relevant for one set of parameters while another subset of transitions may be relevant under different conditions. The nexus plot simply provides a standard method to make this assessment and determine whether one of the canonical equations may be used to represent emission behavior if the operating conditions are sufficiently far from any of the nexus curves or if a combined theory unifying multiple mechanisms is necessary due to proximity to a nexus curve.

As a simple example, consider Fig. 7. If we select V¯=108 and T¯=1 as a relevant operating condition, we would only need to consider RLD since we are far from a nexus; however, reducing T¯ to 0.1 moves the condition closer to the MG/RLD nexus, which would require incorporating collisions into the emission theory derived to link thermionic emission and space-charge limited emission.8 Since the phase space for nexus theory in this paper is six-dimensional (voltage, gap distance, temperature, mobility, resistance, and photon frequency), it is important to use all relevant two-dimensional phase space nexus plots to examine the physical parameters relative to all nexuses of interest. While complicated, these nexus plots provide physical intuition and are especially useful when one or two parameters are intended to vary for a given application. Even when multiple parameters may be varied, the nexus conditions introduced here and extensions to higher-order nexuses may provide guidance to either reduce the dimensionality of the design space or shift focus to a specific electron emission mechanism or mechanisms.

For photoemission, which we have added to our prior first-principles analytic and nexus theories,4,8–10 we have demonstrated conditions where FD transitions to either RLD or FN directly before reaching the space-charge limited current. The nexus plots also show the conditions for the transition from FD to other mechanisms depending on system parameters. Of specific note, Fig. 8 shows that for certain ω¯=λ¯1, increasing the voltage can cause FD to directly transition to MG. Although not shown, FD may directly transition to CL to sufficiently low ω¯ and high V¯, which may or may not be physically relevant depending on device operating conditions. Moreover, based on prior nexus theory studies,9 one may conjecture that selecting an appropriately high μ¯ (or low pressure) will cause FD to directly transition to CL for sufficiently high V¯. The continuing decrease in photoemission device sizes coupled with applications requiring larger currents necessitates increased awareness of the possibility of achieving a space-charge limited condition due to photoemission. As discussed above, this becomes even more challenging when photoemission may be produced either intentionally or unintentionally (due to the low work function of photocathodes) combined with thermionic or field emission. The nexus approach presented here provides a flexible, rapid tool for combining these multiple theories to assess the relevant conditions to facilitate theoretical development and system design for a given condition mathematically using a single equation or visually using a plot.

Thus, the theory and procedure provided here is remarkably general for most emission mechanisms considered in the literature. We note that it is not complete since other mechanisms may be linked (if not directly unified), such as quantum SCLC or gas breakdown,11 as appropriate or necessary. Furthermore, the present theory only considers parallel plate geometries, meaning that it does not account for surface roughness,59 electrode curvature,60 or magnetic fields.61,62 Surface roughness and electrode curvature may be taken into account by applying variational calculus63 or conformal mapping,64 which have recently been applied to derive SCLC for nonplanar diodes. For a magnetic field orthogonal to the electric field, we may follow the same process as above by incorporating the appropriate electron emission source equations into prior derivations of SCLC for such a crossed-field diode.61,62 For a crossed-field diode, it is important to separately address the relevant physics where the emitted electrons cross the gap and reach the anode61 or are magnetically insulated and return to the cathode62 depending on the intensity of the applied magnetic field.

In summary, even for the relatively simple case considered here, nexus theory provides important details on the relevant transitions between different mechanisms depending on device parameters. Moreover, it may be further extended to address more physically relevant diode geometries or include additional emission or electromagnetic physics. In particular, future work could focus on incorporating optical tunneling of an electron exposed to a laser field into the GTFP and nexus theories to more accurately account for behavior for values of the Keldysh parameter where both multiphoton and optical field emission are important.56,57 A future study could also account for time-dependent and nonequilibrium temperature changes during laser exposure (particularly for ultrafast lasers) that would cause deviation from the standard equation for thermionic emission.58 These physical conditions become particularly important as space-charge contributions increase, which could be added to future nexus theory studies with photoemission. Nexus theory provides a straightforward tool for theorists for determine the appropriate theory/theories to consider and incorporate more or fewer physical phenomena depending on the experimental conditions of interest, computational physicists for determining appropriate algorithms, and experimentalists for selecting device parameters to achieve desired emission behavior.

This material is based on work supported by the Air Force Office of Scientific Research under Award No. FA9550-18-1-0218.

The data that support the findings of this study are available within the article.

1.
P.
Zhang
,
A.
Valfells
,
L. K.
Ang
,
J. W.
Luginsland
, and
Y. Y.
Lau
,
Appl. Phys. Rev.
4
,
011304
(
2017
).
2.
K. L.
Jensen
,
IEEE Trans. Plasma Sci.
46
,
1881
(
2018
).
3.
P.
Zhang
,
Y. S.
Ang
,
A. L.
Garner
,
Á
Valfells
,
J. W.
Luginsland
, and
L. K.
Ang
,
J. Appl. Phys.
129
,
100902
(
2021
).
4.
A. L.
Garner
,
G.
Meng
,
Y.
Fu
,
A. M.
Loveless
,
R. S.
Brayfield
 II
, and
A. M.
Darr
,
J. Appl. Phys.
128
,
210903
(
2020
).
5.
A. L.
Garner
,
A. M.
Loveless
,
J. N.
Dahal
, and
A.
Venkattraman
,
IEEE Trans. Plasma Sci.
48
,
808
(
2020
).
6.
D. B.
Go
and
A.
Venkattraman
,
J. Phys. D: Appl. Phys.
47
,
503001
(
2014
).
7.
Y.
Fu
,
P.
Zhang
,
J. P.
Verboncoeur
, and
X.
Wang
,
Plasma Res. Express
2
,
013001
(
2020
).
8.
A. M.
Darr
,
C. R.
Darr
, and
A. L.
Garner
,
Phys. Rev. Res.
2
,
033137
(
2020
).
9.
A. M.
Darr
,
A. M.
Loveless
, and
A. L.
Garner
,
Appl. Phys. Lett.
114
,
014103
(
2019
).
10.
S. D.
Dynako
,
A. M.
Darr
, and
A. L.
Garner
,
IEEE J. Electron Devices Soc.
7
,
650
(
2019
).
11.
A. M.
Loveless
,
A. M.
Darr
, and
A. L.
Garner
,
Phys. Plasmas
28
,
042110
(
2021
).
12.
Y. Y.
Lau
,
Y.
Liu
, and
R. K.
Parker
,
Phys. Plasmas
1
,
2082
(
1994
).
13.
C. D.
Child
,
Phys. Rev. Ser. I
32
,
492
(
1911
).
14.
15.
R. H.
Fowler
and
L.
Nordheim
,
Proc. R. Soc. London Ser. A
119
,
173
(
1928
).
16.
E. L.
Murphy
and
R. H.
Good
, Jr.
,
Phys. Rev.
102
,
1464
(
1956
).
17.
R. G.
Forbes
,
Proc. R. Soc. London Ser. A
469
,
20130271
(
2013
).
18.
M. S.
Benilov
,
Plasma Sources Sci. Technol.
18
,
014005
(
2009
).
19.
N. F.
Mott
and
R. W.
Gurney
,
Electronic Processes in Ionic Crystals
(
Clarendon
,
Oxford
,
1940
).
20.
J. D.
Levine
,
R.
Meyer
,
R.
Baptist
,
T. E.
Felter
, and
A. A.
Talin
,
J. Vac. Sci. Technol. B
13
,
474
(
1995
).
21.
J. W.
Luginsland
,
A.
Valfells
, and
Y. Y.
Lau
,
Appl. Phys. Lett.
69
,
2770
(
1996
).
22.
A. K.
Singh
,
S. K.
Shukla
,
M.
Ravi
, and
R. K.
Barik
,
IEEE Trans. Plasma Sci.
48
,
3446
(
2020
).
23.
O. W.
Richardson
and
A. F. A.
Young
,
Proc. R. Soc. London Ser. A
107
,
377
(
1925
).
24.
K. L.
Jensen
,
M.
McDonald
,
O.
Chubenko
,
J. R.
Harris
,
D. A.
Shiffler
,
N. A.
Moody
,
J. J.
Petillo
, and
A. J.
Jensen
,
J. Appl. Phys.
125
,
234303
(
2019
).
25.
K. L.
Jensen
,
P. G.
O’Shea
, and
D. W.
Feldman
,
Appl. Phys. Lett.
81
,
3867
(
2002
).
26.
K. L.
Jensen
,
J. Appl. Phys.
102
,
024911
(
2007
).
27.
Y.
Wang
,
J.
Wang
,
W.
Liu
,
L.
Li
,
Y.
Wang
, and
X.
Zhang
,
IEEE Trans. Electron Devices
56
,
776
(
2009
).
28.
W.
Liu
,
Y.
Wang
,
J.
Wang
,
Y.
Wang
, and
B.
Vancil
,
IEEE Trans. Electron Devices
58
,
1241
(
2011
).
29.
A.
Sitek
,
K.
Torfason
,
A.
Manolescu
, and
Á.
Valfells
,
Phys. Rev. Appl.
15
,
014040
(
2021
).
30.
D.
Chernin
,
Y. Y.
Lau
,
J. J.
Petillo
,
S.
Ovtchinnikov
,
D.
Chen
,
A.
Jassem
,
R.
Jacobs
,
D.
Morgan
, and
J. H.
Booske
,
IEEE Trans. Plasma Sci.
48
,
146
(
2020
).
31.
A.
Jassem
,
D.
Chernin
,
J. J.
Petillo
,
Y. Y.
Lau
,
A.
Jensen
, and
S.
Ovtchinnikov
,
IEEE Trans. Plasma Sci.
49
,
749
(
2021
).
32.
D.
Chen
,
R.
Jacobs
,
D.
Morgan
, and
J.
Booske
,
IEEE Trans. Electron Devices
68
,
3576
(
2021
).
33.
H. P.
Bonzel
and
Ch.
Kleint
,
Prog. Surf. Sci.
49
,
107
(
1995
).
34.
L. A.
DuBridge
,
Phys. Rev.
43
,
727
(
1933
).
35.
A. H.
Zewail
and
J. M.
Thomas
,
4D Electron Microscopy Imaging in Space and Time
, 1st ed. (
Imperial College
,
London
,
2009
).
36.
S.
Sun
,
X.
Sun
,
D.
Bartles
,
E.
Wozniak
,
J.
Williams
,
P.
Zhang
, and
C.-Y.
Ruan
,
Struct. Dyn.
7
,
064301
(
2020
).
37.
W. A.
Barletta
 et al.,
Nucl. Instrum. Methods Phys. Res. Sect. A
618
,
69
(
2010
).
38.
B.
Piglosiewicz
,
S.
Schmidt
,
D. J.
Park
,
J.
Vogelsang
,
P.
Groß
,
C.
Manzoni
,
P.
Farinello
,
G.
Cerullo
, and
C.
Lienau
,
Nat. Photonics
8
,
37
(
2014
).
39.
F.
Rezaeifar
,
R.
Ahsan
,
Q.
Lin
,
H. U.
Chae
, and
R.
Kapadia
,
Nat. Photonics
13
,
843
(
2019
).
40.
H. U.
Chae
,
R.
Ahsan
,
Q.
Lin
,
D.
Sarkar
,
F.
Rezaeifar
,
S. B.
Cronin
, and
R.
Kapadia
,
Nano Lett.
19
,
6227
(
2019
).
41.
E.
Forati
,
T. J.
Dill
,
A. R.
Tao
, and
D.
Sievenpiper
,
Nat. Commun.
7
,
13399
(
2016
).
42.
S.
Piltan
and
D.
Sievenpiper
,
J. Opt. Soc. Am. B
35
,
208
(
2018
).
43.
K. L.
Jensen
,
D. W.
Feldman
,
M.
Virgo
, and
P. G.
O’Shea
,
Phys. Rev. Spec. Top. Accel. Beams
6
,
083501
(
2003
).
44.
K. L.
Jensen
,
D. W.
Feldman
, and
P. G.
O’Shea
,
J. Vac. Sci. Technol. B
23
,
621
(
2005
).
45.
K. L.
Jensen
,
D. W.
Feldman
,
N. A.
Moody
, and
P. G.
O’Shea
,
J. Appl. Phys.
99
,
124905
(
2006
).
46.
K. L.
Jensen
, in
Modern Developments in Vacuum Electron Sources
, Topics in Applied Physics, edited by
G.
Gaertner
,
W.
Knapp
, and
R. G.
Forbes
(
Springer
,
Cham
,
2020
), pp.
345
385
.
47.
W.
Liu
,
M.
Poelker
,
J.
Smedley
, and
R.
Ganter
, in
Modern Developments in Vacuum Electron Sources
, Topics in Applied Physics, edited by
G.
Gaertner
,
W.
Knapp
, and
R. G.
Forbes
(
Springer
,
Cham
,
2020
), pp.
293
344
.
48.
J. W.
Schwede
 et al.,
Nat. Mater.
9
,
762
(
2010
).
49.
P.
Zhang
and
Y. Y.
Lau
,
Sci. Rep.
6
,
19894
(
2016
).
50.
E.
Forati
and
D.
Sievenpiper
,
J. Appl. Phys.
124
,
083101
(
2018
).
51.
K. L.
Jensen
,
N. A.
Moody
,
D. W.
Feldman
,
E. J.
Montgomery
, and
P. G.
O’Shea
,
J. Appl. Phys.
102
,
074902
(
2007
).
52.
K. L.
Jensen
,
D. A.
Shiffler
,
J. J.
Petillo
,
Z.
Pan
, and
J. W.
Luginsland
,
Phys. Rev. Spec. Top. Accel. Beams
17
,
043402
(
2014
).
53.
K. L.
Jensen
 et al.,
J. Vac. Sci. Technol. B.
26
,
831
(
2008
).
54.
K. L.
Jensen
,
J. Appl. Phys.
111
,
054916
(
2012
).
55.
K. L.
Jensen
,
J. Appl. Phys.
126
,
065302
(
2019
).
56.
L. K.
Ang
and
M.
Pant
,
Phys. Plasmas
20
,
056705
(
2013
).
57.
M.
Pant
and
L. K.
Ang
,
Phys. Rev. B
86
,
045423
(
2012
).
58.
L.
Wu
and
L. K.
Ang
,
Phys. Rev. B
78
,
224112
(
2008
).
59.
R. S.
Brayfield
 II
,
A. J.
Fairbanks
,
A. M.
Loveless
,
S.
Gao
,
A.
Dhanabal
,
W.
Li
,
C.
Darr
,
W.
Wu
, and
A. L.
Garner
,
J. Appl. Phys.
125
,
203302
(
2019
).
60.
Y. B.
Zhu
and
L. K.
Ang
,
Phys. Plasmas
22
,
052106
(
2015
).
61.
Y. Y.
Lau
,
P. J.
Christenson
, and
D.
Chernin
,
Phys. Fluids B
5
,
4486
(
1993
).
62.
P. J.
Christenson
and
Y. Y.
Lau
,
Phys. Plasmas
1
,
3725
(
1994
).
63.
A. M.
Darr
and
A. L.
Garner
,
Appl. Phys. Lett.
115
,
054101
(
2019
).
64.
N. R.
Sree Harsha
and
A. L.
Garner
,
IEEE Trans. Electron Devices
68
,
264
(
2021
).