Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
\sidecaptionvpos

figuret

Computational design of tunnel diodes with negative differential resistance and ultrahigh peak-to-valley current ratio based on two-dimensional cold metals: The case of NbSi2N4/HfSi2N4/NbSi2N4 lateral heterojunction diode

P. Bodewei1 paul.bodewei@student.uni-halle.de    E. Şaşıoğlu1    N. F. Hinsche1    I. Mertig1 1Institute of Physics, Martin Luther University Halle-Wittenberg, 06120 Halle (Saale), Germany
(June 18, 2024)
Abstract

Cold metals have recently gained attention as a promising platform for innovative devices, such as tunnel diodes with negative differential resistance (NDR) and field-effect transistors with subthreshold swings below the thermionic limit. Recently discovered two-dimensional (2D) MA2Z4 (M=Ti, Zr, Hf, Nb, Ta; A=Si, Ge; Z=N, P) compounds exhibit both cold metallic and semiconducting behavior. In this work, we present a computational study of lateral heterojunction tunnel diodes based on 2D NbSi2N4 and HfSi2N4 compounds. Employing density functional theory combined with a nonequilibrium Green function method, we investigate the current-voltage (I𝐼Iitalic_I-V𝑉Vitalic_V) characteristics of lateral tunnel diodes with varying barrier thicknesses in both zigzag and armchair orientations. We find that tunnel diodes in the zigzag orientation exhibit significantly higher peak current densities, while those in the armchair orientation display larger peak-to-valley current ratios (PVCRs) compared to the zigzag orientation. Our findings suggest that MA2Z4 materials are promising candidates for realizing NDR tunnel diodes with ultra-high PVCR values, which could have potential applications in memory, logic circuits, and other electronic devices.

I Introduction

Negative differential resistance (NDR) tunnel diodes offer unique functionalities and potential applications when integrated within conventional complementary metal-oxide-semiconductor (CMOS) transistors [1]. These include NDR-based multi-valued logic gates, static random-access memory (SRAM), magnetic random-access memory, and low-power oscillators [2, 3, 4]. Exploiting the NDR effect allows for logic and memory architectures with reduced device count, higher speed, and lower power consumption. For example, a tunnel SRAM requires only one transistor and two NDR tunnel diodes, instead of six transistors in a conventional SRAM. This results in a smaller footprint and lower power consumption [4, 5]. NDR diodes can be combined with CMOS transistors to implement various logic gates, while the extension to multi-valued logic increases information density and decreases system complexity [6, 7, 8].

The NDR effect has been demonstrated in a variety of devices and circuits, with a focus on two-terminal tunnel diodes like Esaki diodes and resonant tunneling diodes [9, 10, 11, 12]. Esaki diodes, which operate via quantum tunneling under forward bias, show NDR behavior. However, the application of these two-terminal NDR diodes faces challenges. They often have low peak-to-valley current ratios (PVCR), which are not suitable for memory applications like tunnel SRAMs [11, 13, 14, 15]. Additionally, III-V semiconductors, which offer high PVCR, are difficult to integrate within current CMOS technology [16, 17]. Research efforts have sought to enhance PVCR values over 100 through CMOS-compatible processes [4]. While some NDR circuits show extremely high PVCR values, their complex topology with multiple transistors makes them unsuitable for memory applications [18, 19, 20, 21, 22].

Refer to caption
Figure 1: (a) Schematic representation of the lateral and vertical negative differential resistance (NDR) tunnel diodes based on cold metals. CM and I stand for cold metal and insulator, respectively. (b) The I𝐼Iitalic_I-V𝑉Vitalic_V characteristics showing conventional N-type and (c) ΛΛ\Lambdaroman_Λ-type NDR.

Existing semiconductor-based NDR tunnel diodes suffer from low peak-to-valley current ratio (PVCR) values, which is attributed to the band tails tunneling that originates from the strong doping and dopant fluctuations [23, 24, 25]. To address this issue, we recently proposed a new semiconductor-free NDR tunnel diode concept with ultra-high PVCR [26]. Our proposed diode consists of two cold metal electrodes separated by a thin insulating tunnel barrier (see Fig. 1). The NDR effect arises from the unique electronic band structure of the cold-metal electrodes (see Fig. 2). A cold metal is obtained when the metallic, i.e. partially filled, bands are separated from all energetically higher and lower-lying bands by sufficiently large energy gaps. Specifically, the width of the isolated metallic bands around the Fermi level, as well as the energy gaps separating higher and lower-lying bands, determine the current-voltage (I𝐼Iitalic_I-V𝑉Vitalic_V) characteristics and PVCR of the tunnel diode. By choosing the appropriate cold metal electrode materials, either a conventional N𝑁Nitalic_N-type or ΛΛ\Lambdaroman_Λ-type NDR effect can be achieved as sketched in Fig. 1(b,c).

The intention of this paper is to computationally design lateral heterojunction NDR tunnel diodes based on the recently discovered family of 2D materials known as MA2Z4 (where M=Ti, Zr, Hf, Nb, Ta; A=Si, Ge; Z=N, P) [27]. These MA2Z4 compounds provide an exceptional platform for realizing NDR tunnel diodes due to their closely matched lattice parameters and composition, as well as their ability to exhibit both cold metallic and semiconducting properties within the same material class [28]. This characteristic enables the coherent growth of consecutive components within the device which naturally favors a lateral device geometry [29]. Furthermore, we prefer a lateral heterojunction design towards the vertical counterpart, as a higher current density can be expected, especially pronounced in van der Waals 2D materials [30]. An additional benefit of the MA2Z4 compounds is their superior strength and mechanical stability compared to most monolayer semiconductors like MoS2 [28]. Additionally, the electronic properties of MA2Z4 compounds are more resilient when it comes to stacking. For example, the cold metallic behavior observed in MX2 (where M=Nb, Ta; X=S, Se, Te) is typically limited to monolayers, disappearing after only a few layers [31]. In contrast, MA2Z4 compounds maintain this behavior even in their bulk phase.

Refer to caption
Figure 2: Schematic illustration of the density of states (DOS) depicting: (a) a normal metal, (b) a semiconductor, and (c) a cold metal. The symbols Egsubscript𝐸gE_{\mathrm{g}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT, EgIsubscriptsuperscript𝐸IgE^{\mathrm{I}}_{\mathrm{g}}italic_E start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT, and EgEsubscriptsuperscript𝐸EgE^{\mathrm{E}}_{\mathrm{g}}italic_E start_POSTSUPERSCRIPT roman_E end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT correspond to the band gap of the semiconductor, as well as the internal and external band gaps of the cold metal, respectively. The width of the metallic band for the cold metal is designated by Wmsubscript𝑊mW_{\mathrm{m}}italic_W start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT. The Fermi level is denoted by EFsubscript𝐸𝐹E_{F}italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT.

II Computational Method

The computational design of the presented lateral heterojunction tunnel diodes is based on density functional theory (DFT) within the QuantumATK T-2022.03 package [32]. We used a linear combination of atomic orbitals (LCAO) basis set (double zeta polarized) with norm-conserving FHI pseudopotentials [33]. The exchange-correlation functional was represented by a generalized gradient approximation (GGA) with Perdew-Burke-Ernzerhof (PBE) flavor[34]. The self-consistent DFT calculations are performed using a 24×24×12424124\times 24\times 124 × 24 × 1 k-point grid with a density mesh cut-off of 45Hartree45Hartree45\,\mathrm{Hartree}45 roman_Hartree and the total energy converged to at least 104Hartreesuperscript104Hartree10^{-4}\,\mathrm{Hartree}10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT roman_Hartree. To prevent interactions between the periodically repeated images, 20 Åtimes20angstrom20\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 20 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG of vacuum were added and Dirichlet and Neumann boundary conditions are employed. The transport calculations were performed using DFT combined with a nonequilibrium Green’s function method (NEGF). We use a 24×1×18424118424\times 1\times 18424 × 1 × 184 (14×1×31814131814\times 1\times 31814 × 1 × 318) 𝐤𝐤\mathbf{k}bold_k-point grid in self-consistent DFT-NEGF calculations of the lateral tunnel diodes along the armchair (zigzag) orientation. The I𝐼Iitalic_I-V𝑉Vitalic_V characteristics were calculated within a Landauer approach [35], where I(V)=2ehT(E,V)[fL(E,V)fR(E,V)]dE𝐼𝑉2𝑒𝑇𝐸𝑉delimited-[]subscript𝑓𝐿𝐸𝑉subscript𝑓𝑅𝐸𝑉differential-d𝐸I(V)=\frac{2e}{h}\int\,T(E,V)\left[f_{L}(E,V)-f_{R}(E,V)\right]\mathrm{d}Eitalic_I ( italic_V ) = divide start_ARG 2 italic_e end_ARG start_ARG italic_h end_ARG ∫ italic_T ( italic_E , italic_V ) [ italic_f start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_E , italic_V ) - italic_f start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_E , italic_V ) ] roman_d italic_E. Here V𝑉Vitalic_V denotes the bias voltage, T(E,V)𝑇𝐸𝑉T(E,V)italic_T ( italic_E , italic_V ) are the transmission coefficients and fL(E,V)subscript𝑓𝐿𝐸𝑉f_{L}(E,V)italic_f start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_E , italic_V ) and fR(E,V)subscript𝑓𝑅𝐸𝑉f_{R}(E,V)italic_f start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_E , italic_V ) are the Fermi-Dirac distributions of the left and right electrodes, respectively. The temperature was kept at T=𝑇absentT=italic_T = 300 Ktimes300K300\text{\,}\mathrm{K}start_ARG 300 end_ARG start_ARG times end_ARG start_ARG roman_K end_ARG throughout all calculations. The transmission coefficients T(E,V)𝑇𝐸𝑉T(E,V)italic_T ( italic_E , italic_V ) for the lateral tunnel diodes along the armchair (zigzag) orientation are calculated using a 255×12551255\times 1255 × 1 ( 147×11471147\times 1147 × 1) 𝐤𝐤\mathbf{k}bold_k-point grid and convergence tests can be found in the supplemental material [36].

III Results and Discussion

In Fig. 1 we present schematically the structure of the lateral and vertical NDR tunnel diode and the corresponding IV𝐼𝑉I-Vitalic_I - italic_V characteristics. The concept of the cold metal NDR tunnel diode was previously introduced by us in Ref. [26], and thus only a brief overview of the device will be given subsequently. Our NDR tunnel diode consists of two cold metal electrodes, which are separated by an insulating tunnel barrier. The schematic density of states (DOS) of a cold metal is presented in Fig. 2 and compared with a DOS of a conventional metal and a semiconductor. As seen in the schematic DOS cold metallic materials possess a unique band structure that has an isolated metallic band Wmsubscript𝑊mW_{\mathrm{m}}italic_W start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT around the Fermi level as well as the energy gaps separating higher- and lower-lying states. The latter is referred to as the internal gap EgIsubscriptsuperscript𝐸IgE^{\mathrm{I}}_{\mathrm{g}}italic_E start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT (below the Fermi level) and external EgEsubscriptsuperscript𝐸EgE^{\mathrm{E}}_{\mathrm{g}}italic_E start_POSTSUPERSCRIPT roman_E end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT (above the Fermi level). These three electronic structure parameters play a decisive role in determining the I𝐼Iitalic_I-V𝑉Vitalic_V characteristics of the tunnel diode and the type of the NDR effect (see Fig. 1 (b)). A conventional N𝑁Nitalic_N-type NDR effect is expected if these three parameters Wmsubscript𝑊mW_{\mathrm{m}}italic_W start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT, EgIsubscriptsuperscript𝐸IgE^{\mathrm{I}}_{\mathrm{g}}italic_E start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT, and EgEsubscriptsuperscript𝐸EgE^{\mathrm{E}}_{\mathrm{g}}italic_E start_POSTSUPERSCRIPT roman_E end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT are close to each other. On the other hand, the tunnel diode will show ΛΛ\Lambdaroman_Λ-type NDR effect if WmEgIEgEmuch-less-thansubscript𝑊msubscriptsuperscript𝐸Igsimilar-tosubscriptsuperscript𝐸EgW_{\mathrm{m}}\ll E^{\mathrm{I}}_{\mathrm{g}}\sim E^{\mathrm{E}}_{\mathrm{g}}italic_W start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT ≪ italic_E start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT ∼ italic_E start_POSTSUPERSCRIPT roman_E end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT.

Table 1: Lattice constants a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, work function ΦΦ\Phiroman_Φ, band gap Egsubscript𝐸gE_{\mathrm{g}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT, internal gap EgIsubscriptsuperscript𝐸IgE^{\mathrm{I}}_{\mathrm{g}}italic_E start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT, external gap EgEsubscriptsuperscript𝐸EgE^{\mathrm{E}}_{\mathrm{g}}italic_E start_POSTSUPERSCRIPT roman_E end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT and metallic bandwidth Wmsubscript𝑊mW_{\mathrm{m}}italic_W start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT of cold metallic and semiconducting MA2Z4 (M = Nb, Ta, Ti, Zr, Hf; A = Si, Ge; Z = N, P) monolayers. Lattice parameters are taken from Ref. [28]. The band gap of HfSi2N4 calculated within the GGA+U𝑈Uitalic_U method is given in parentheses.
Compound a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ΦΦ\Phiroman_Φ Egsubscript𝐸gE_{\mathrm{g}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT EgIsubscriptsuperscript𝐸IgE^{\mathrm{I}}_{\mathrm{g}}italic_E start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT EgEsubscriptsuperscript𝐸EgE^{\mathrm{E}}_{\mathrm{g}}italic_E start_POSTSUPERSCRIPT roman_E end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT Wmsubscript𝑊mW_{\mathrm{m}}italic_W start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT
(Å) (eV) (eV) (eV) (eV) (eV)
NbSi2N4 2.97 5.62 1.01 1.74 1.33
NbGe2N4 3.08 5.69 0.92 1.21 1.16
NbSi2P4 3.53 4.52 0.01 0.52 1.26
TaSi2N4 2.97 5.47 1.16 1.41 1.56
TaGe2N4 3.08 5.51 1.16 0.98 1.25
TaSi2P4 3.53 4.45 0.00 0.27 1.51
TiSi2N4 2.94 6.08 1.54
ZrSi2N4 3.04 5.94 1.46
HfSi2N4 3.03 5.92 1.61 (1.84)

To design lateral NDR tunnel diodes, we first performed a material screening within the MoSi2N4 compound family, specifically focusing on the cold metal and semiconducting compounds, which are listed in Table 1. Our selection criteria for materials included cold metal electrode materials with large internal (EgIsuperscriptsubscript𝐸gIE_{\mathrm{g}}^{\mathrm{I}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT) and external (EgEsuperscriptsubscript𝐸gEE_{\mathrm{g}}^{\mathrm{E}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_E end_POSTSUPERSCRIPT) band gaps relative to their metallic band widths (Wmsubscript𝑊mW_{\mathrm{m}}italic_W start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT). For tunnel barriers, we chose materials with large band gaps (Egsubscript𝐸gE_{\mathrm{g}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT) and work functions that matched those of the cold metallic compounds. In Table 1, we provide the lattice parameters, work functions, band gaps (EgIsuperscriptsubscript𝐸gIE_{\mathrm{g}}^{\mathrm{I}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT, EgEsuperscriptsubscript𝐸gEE_{\mathrm{g}}^{\mathrm{E}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_E end_POSTSUPERSCRIPT, Egsubscript𝐸gE_{\mathrm{g}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT), and band widths for all the selected compounds. As presented, all the cold metal compounds under consideration, with few exceptions, have either internal or external band gaps smaller than their corresponding band widths, indicating that tunnel diodes based on these materials would exhibit the conventional N𝑁Nitalic_N-type NDR effect. Among the screened materials, we selected NbSi2N4 as the cold metal electrode and HfSi2N4 as the tunnel barrier for our lateral tunnel diode simulations. The calculated electronic band structures of these materials are presented in Fig. 3. As all members of the 2D MoSi2N4 compound family belong to the trigonal P6¯m2𝑃¯6𝑚2P\overline{6}m2italic_P over¯ start_ARG 6 end_ARG italic_m 2 (#187#187\#187# 187) spacegroup, the bands are shown along the high-symmetry points of the related hexagonal Brillouin zone. The monolayer NbSi2N4 possesses an internal and external band gap of 1.01 eVtimes1.01eV1.01\text{\,}\mathrm{e}\mathrm{V}start_ARG 1.01 end_ARG start_ARG times end_ARG start_ARG roman_eV end_ARG and 1.74 eVtimes1.74eV1.74\text{\,}\mathrm{e}\mathrm{V}start_ARG 1.74 end_ARG start_ARG times end_ARG start_ARG roman_eV end_ARG respectively, while its metallic band width is 1.33 eVtimes1.33eV1.33\text{\,}\mathrm{e}\mathrm{V}start_ARG 1.33 end_ARG start_ARG times end_ARG start_ARG roman_eV end_ARG within DFT-PBE. The tunnel barrier HfSi2N4 has a band gap of 1.61 eVtimes1.61eV1.61\text{\,}\mathrm{e}\mathrm{V}start_ARG 1.61 end_ARG start_ARG times end_ARG start_ARG roman_eV end_ARG within DFT-PBE. Note that DFT-PBE underestimates the band gap of semiconductors and insulators, as well as cold metals (including cold metals such as the MX2 (M=Nb, Ta; X=S, Se, Te) compounds [37, 38]). To improve the band gap of the tunnel barrier HfSi2N4, we employed a PBE+U𝑈Uitalic_U method with a U𝑈Uitalic_U value of 8 eVtimes8eV8\text{\,}\mathrm{e}\mathrm{V}start_ARG 8 end_ARG start_ARG times end_ARG start_ARG roman_eV end_ARG for the d𝑑ditalic_d orbitals of Hf in the transport calculations (see Table 1).

In Fig. 4, we show the atomic structure of the lateral tunnel diode, which is formed by attaching one monolayer of cold metal NbSi2N4 as the left and right electrodes and one monolayer of HfSi2N4 as the tunnel barrier. We consider both armchair and zigzag directions as the transport direction for the tunnel diodes, which we will refer to as armchair and zigzag tunnel diodes, respectively. The device is periodic in the x𝑥xitalic_x-direction, and we choose the z𝑧zitalic_z-direction as the transport direction. Since the lattice constants of the electrode and tunnel barrier materials are slightly different (less than 2%percent22\%2 % mismatch, see Table 1) for both directions, the tunnel barrier adopts the in-plane (x𝑥xitalic_x-direction) lattice constant of the electrode material, and its lattice constant along the transport direction (z𝑧zitalic_z-direction) is relaxed. The tunnel barrier thicknesses are chosen to be about 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG and 30 Åtimes30angstrom30\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG for both armchair and zigzag tunnel diodes. The total length of the scattering region ranges from 81 Åtimes81angstrom81\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 81 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG to 90 Åtimes90angstrom90\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 90 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG.

Refer to caption
Figure 3: Calculated PBE band structure of monolayer NbSi2N4 and PBE+U𝑈Uitalic_U band structure of HfSi2N4 along the high-symmetry lines in the 2D Brillouin zone. The dashed black lines denote the Fermi level, which is set to zero. The isolated metallic band of the NbSi2N4 is highlighted in blue.
Refer to caption
Figure 4: Schematic illustration of the atomic structure of the NbSi2N4/HfSi2N4/NbSi2N4 lateral heterojunction tunnel diode device in armchair orientation. The tunnel barrier width is 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG. Different atomic components are represented by distinct colors.
Refer to caption
Figure 5: (a) Calculated room temperature current-voltage characteristics for the NbSi2N4/HfSi2N4/NbSi2N4 lateral heterojunction tunnel diode with 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG tunnel barrier in both armchair and zigzag orientations. (b) The same as (a) for a tunnel barrier length of 30 Åtimes30angstrom30\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG.

In Fig. 5, we present the calculated I𝐼Iitalic_I-V𝑉Vitalic_V curves for tunnel diodes with tunnel barrier widths of 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG and 30 Åtimes30angstrom30\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG, for both armchair (black lines) and zigzag (red lines) directions. A maximum bias voltage of 1.5 Vtimes1.5V1.5\text{\,}\mathrm{V}start_ARG 1.5 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG was chosen. We report a conventional N𝑁Nitalic_N-type NDR effect in both transport directions with a valley voltage VVsubscript𝑉𝑉V_{V}italic_V start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT of approximately 1.25 Vtimes1.25V1.25\text{\,}\mathrm{V}start_ARG 1.25 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG, which is primarily set by the metallic band width of the cold metal electrodes (see Table 1). Besides this similarity, the I𝐼Iitalic_I-V𝑉Vitalic_V curves of the armchair and zigzag directions exhibit distinct behavior. In the armchair case, the peak current IPsubscript𝐼𝑃I_{P}italic_I start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT (cf. Fig. 2 )is achieved at very low bias voltages of about 0.2 Vtimes0.2V0.2\text{\,}\mathrm{V}start_ARG 0.2 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG, while in the zigzag case it appears at higher bias voltages of V>0.7 V𝑉times0.7VV>$0.7\text{\,}\mathrm{V}$italic_V > start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG. These details can be directly attributed to the overlap of contributing electronic states of the left and right cold metal electrodes in different areas of the Brillouin zone. The mentioned overlap differs for armchair and zigzag-oriented tunnel diodes resulting in the presented I𝐼Iitalic_I-V𝑉Vitalic_V curves in Fig. 5. In a first approximation, the spin-independent transmission through an insulating barrier of width d𝑑ditalic_d at zero bias is T(kz,E)=tL(kz,E)exp[2κ(kz,E)d]tR(kz,E)𝑇subscript𝑘𝑧𝐸subscript𝑡𝐿subscript𝑘𝑧𝐸2𝜅subscript𝑘𝑧𝐸𝑑subscript𝑡𝑅subscript𝑘𝑧𝐸T(k_{z},E)=t_{L}(k_{z},E)\exp\bigl{[}-2\kappa(k_{z},E)d\bigr{]}t_{R}(k_{z},E)italic_T ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) = italic_t start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) roman_exp [ - 2 italic_κ ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) italic_d ] italic_t start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) [39]. Here tL(kz,E)subscript𝑡𝐿subscript𝑘𝑧𝐸t_{L}(k_{z},E)italic_t start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) and tR(kz,E)subscript𝑡𝑅subscript𝑘𝑧𝐸t_{R}(k_{z},E)italic_t start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) are the interface transmission functions of the left and right metallic leads, respectively. Assuming perfect symmetry matching of the wave functions at the interfaces, the product tL(kz,E)tR(kz,E)subscript𝑡𝐿subscript𝑘𝑧𝐸subscript𝑡𝑅subscript𝑘𝑧𝐸t_{L}(k_{z},E)\cdot t_{R}(k_{z},E)italic_t start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) ⋅ italic_t start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) coincides with the joint density of states of left and right electrodes. The remaining factor involves the exponential decay length 1/κ(kz,E)1𝜅subscript𝑘𝑧𝐸1/\kappa(k_{z},E)1 / italic_κ ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) of the wave function within the insulating barrier. The latter is directly connected to the complex band structure E(kz+ıκ)𝐸subscript𝑘𝑧italic-ı𝜅E(k_{z}+\imath\kappa)italic_E ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + italic_ı italic_κ ) of the barrier, which will be discussed later in the letter. We introduce further insights on the specific structure of the I𝐼Iitalic_I-V𝑉Vitalic_V curves maxima via the energy- and state-resolved transmission, as well as projected interface band structures in the supplementary material [36].

From Fig. 5 we furthermore identify vanishing valley current in a rather large bias voltage interval of 1.0 Vtimes1.0V1.0\text{\,}\mathrm{V}start_ARG 1.0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG to 1.3 Vtimes1.3V1.3\text{\,}\mathrm{V}start_ARG 1.3 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG for armchair tunnel diodes at both barrier thicknesses. We note, that this is in contrast to conventional Esaki diodes where a vanishing valley current is barely obtainable. For the zigzag tunnel diodes, the transition from the NDR to second positive differential resistance (PDR) takes place in a very narrow voltage window. Note that for the first PDR and NDR regions the current is formed by intra-band tunneling, i.e. from the single Nb d-band in the NbSi2N4 electrodes, while in the second PDR region, the current is due to inter-band tunneling.

The most striking difference is the absolute value of the peak current density in both device geometries. The maximum current density of the armchair tunnel diode located close to 0.25 Vtimes0.25V0.25\text{\,}\mathrm{V}start_ARG 0.25 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG is noticeably smaller than the peak current density of the zigzag tunnel diode located at V𝑉Vitalic_V>0.7 Vtimes0.7V0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG. In particular we find a current density ratio of Izz/Iarm4subscript𝐼𝑧𝑧subscript𝐼𝑎𝑟𝑚4\nicefrac{{I_{zz}}}{{I_{arm}}}\approx 4/ start_ARG italic_I start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_a italic_r italic_m end_POSTSUBSCRIPT end_ARG ≈ 4 for the shorter 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG device and Izz/Iarm40subscript𝐼𝑧𝑧subscript𝐼𝑎𝑟𝑚40\nicefrac{{I_{zz}}}{{I_{arm}}}\approx 40/ start_ARG italic_I start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_a italic_r italic_m end_POSTSUBSCRIPT end_ARG ≈ 40 for the longer 30 Åtimes30angstrom30\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG device. This difference can be readily explained with the help of the complex band structure E(kz+ıκ)𝐸subscript𝑘𝑧italic-ı𝜅E(k_{z}+\imath\kappa)italic_E ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + italic_ı italic_κ ), whereas κ=mkz𝜅𝑚subscript𝑘𝑧\kappa=\Im mk_{z}italic_κ = roman_ℑ italic_m italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. The latter is displayed for the tunnel barrier material HfSi2N4 for both devices orientations in Fig. 6. We recall that the transmission T(kz,E)=tL(kz,E)exp[2κ(kz,E)d]tR(kz,E)𝑇subscript𝑘𝑧𝐸subscript𝑡𝐿subscript𝑘𝑧𝐸2𝜅subscript𝑘𝑧𝐸𝑑subscript𝑡𝑅subscript𝑘𝑧𝐸T(k_{z},E)=t_{L}(k_{z},E)\exp\bigl{[}-2\kappa(k_{z},E)d\bigr{]}t_{R}(k_{z},E)italic_T ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) = italic_t start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) roman_exp [ - 2 italic_κ ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) italic_d ] italic_t start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) through an insulating barrier is to a good approximation only dependent on the available state overlap of the left and right electrodes, their interface symmetry matching and the exponential decay length 1/κ(kz,E)1𝜅subscript𝑘𝑧𝐸1/\kappa(k_{z},E)1 / italic_κ ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_E ) within the insulating barrier. It is worth mentioning, that in the case of the orthorhombic transport geometry of the explicit diode devices, the available states of the electrodes as well as barrier material within the range of the applied bias voltages consist only of Asuperscript𝐴A^{\prime}italic_A start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, A′′superscript𝐴′′A^{\prime\prime}italic_A start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT symmetry characters of the C1hsubscript𝐶1C_{1h}italic_C start_POSTSUBSCRIPT 1 italic_h end_POSTSUBSCRIPT group, i.e. states symmetric or anti-symmetric with respect to reflection through the Hf/Nb mirror plane [40, 41]. As a consequence, there is almost no state-filtering at the interface and the electronic transport can be well understood by an analysis of the complex band structure. Obviously states with a small imaginary part κ𝜅\kappaitalic_κ will have the weakest decay within the HfSi2N4 barrier and will dominate the contribution to the transmission, i.e. the current density.

Refer to caption
Figure 6: Complex band structure of the barrier material HfSi2N4 in (a) armchair and (b) zigzag direction. The corresponding right panels show the imaginary parts κ=mkz𝜅𝑚subscript𝑘𝑧\kappa=\Im mk_{z}italic_κ = roman_ℑ italic_m italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT of the in general complex wavevector kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. The colorbar reflects the value of the real part of kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, i.e. purple bands indicate imaginary bands of the first kind with ekz=0𝑒subscript𝑘𝑧0\Re ek_{z}=0roman_ℜ italic_e italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0; yellow bands correspond to imaginary bands of the second kind with ekz=max𝑒subscript𝑘𝑧𝑚𝑎𝑥\Re ek_{z}=maxroman_ℜ italic_e italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_m italic_a italic_x and all other colors refer to complex bands with ekz0𝑒subscript𝑘𝑧0\Re ek_{z}\neq 0roman_ℜ italic_e italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ≠ 0 and mkz0𝑚subscript𝑘𝑧0\Im mk_{z}\neq 0roman_ℑ italic_m italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ≠ 0.

From Fig. 6 is seen that the fundamental band gap for real-valued kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT is bridged by loops with imaginary-valued κ=mkz𝜅𝑚subscript𝑘𝑧\kappa=\Im mk_{z}italic_κ = roman_ℑ italic_m italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. Comparing the complex band structures of the zigzag and armchair-oriented barrier material we immediately conclude that a single loop will dictate the transport in the zigzag geometry, while a combination of at least two crossing loops will describe the tunneling in the case of the armchair geometry. In all cases, imaginary loops have to connect real-valued kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT of the same symmetry character [42]. The fact that an imaginary loop in the zigzag direction is considerably smaller than in the armchair direction directly reflects the reduced size of the current densities shown in Fig. 5. For simplicity let E=EF𝐸subscript𝐸𝐹E=E_{F}italic_E = italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT and we find minimal values of κzz0.22 1/Åsubscript𝜅𝑧𝑧times0.221Å\kappa_{zz}\approx$0.22\text{\,}\mathrm{1}\mathrm{/}\mathrm{% \SIUnitSymbolAngstrom}$italic_κ start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT ≈ start_ARG 0.22 end_ARG start_ARG times end_ARG start_ARG 1 / roman_Å end_ARG and κarm0.32 1/Åsubscript𝜅𝑎𝑟𝑚times0.321Å\kappa_{arm}\approx$0.32\text{\,}\mathrm{1}\mathrm{/}\mathrm{% \SIUnitSymbolAngstrom}$italic_κ start_POSTSUBSCRIPT italic_a italic_r italic_m end_POSTSUBSCRIPT ≈ start_ARG 0.32 end_ARG start_ARG times end_ARG start_ARG 1 / roman_Å end_ARG for the mid-gap states. Approximating the current density by Iexp[κ(kz=0,EF)d]𝐼𝜅subscript𝑘𝑧0subscript𝐸𝐹𝑑I\approx\exp\bigl{[}-\kappa(k_{z}=0,E_{F})d\bigr{]}italic_I ≈ roman_exp [ - italic_κ ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 , italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) italic_d ] we will obtain ratios of Izz/Iarm8subscript𝐼𝑧𝑧subscript𝐼𝑎𝑟𝑚8\nicefrac{{I_{zz}}}{{I_{arm}}}\approx 8/ start_ARG italic_I start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_a italic_r italic_m end_POSTSUBSCRIPT end_ARG ≈ 8 for the shorter 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG device and Izz/Iarm20subscript𝐼𝑧𝑧subscript𝐼𝑎𝑟𝑚20\nicefrac{{I_{zz}}}{{I_{arm}}}\approx 20/ start_ARG italic_I start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_a italic_r italic_m end_POSTSUBSCRIPT end_ARG ≈ 20 for the longer 30 Åtimes30angstrom30\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG device which is in excellent qualitative agreement with the results of our full NEGF calculations presented in Fig. 5. Due to the evanescent nature of the states the ratio Izz/Iarmsubscript𝐼𝑧𝑧subscript𝐼𝑎𝑟𝑚\nicefrac{{I_{zz}}}{{I_{arm}}}/ start_ARG italic_I start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_a italic_r italic_m end_POSTSUBSCRIPT end_ARG will naturally increase for larger barrier widths.

While the peak current density in zigzag tunnel diodes is notably higher, it is essential to note that these exhibit smaller PVCR values. The PVCR values obtained for both tunnel barrier thicknesses are approximately 500. In contrast, for armchair tunnel diodes, the PVCR values are significantly higher, reaching 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT and 105superscript10510^{5}10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT for tunnel barriers of 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG and 30 Åtimes30angstrom30\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG, respectively. These findings highlight the trade-off between peak current density and PVCR, and the unique characteristics of armchair and zigzag tunnel diodes for diverse applications. We note that these PVCR values are order of magnitudes higher than found in conventional Ge-based NDR diodes (PVCR 210absent210\approx 2\dots 10≈ 2 … 10), CaF2-based NDR diodes (PVCR 100absent100\approx 100≈ 100) [1] or even in MoS2 homojunctions with a PVCR 200400absent200400\approx 200\dots 400≈ 200 … 400 [30].

We want to emphasize that the choice of pseudopotentials and basis sets can be crucial in the evaluation of the complex band structure and thus of the related I𝐼Iitalic_I-V𝑉Vitalic_V characteristics within the LCAO-NEGF formalism. Pay attention to the vertical complex bands appearing with an almost constant decay length 1/κz1subscript𝜅𝑧1/\kappa_{z}1 / italic_κ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT over a wide energy range - e.g. the imaginary band of second kind at κ=mkz0.35 1/Å𝜅𝑚subscript𝑘𝑧times0.351Å\kappa=\Im mk_{z}\approx$0.35\text{\,}\mathrm{1}\mathrm{/}\mathrm{% \SIUnitSymbolAngstrom}$italic_κ = roman_ℑ italic_m italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ≈ start_ARG 0.35 end_ARG start_ARG times end_ARG start_ARG 1 / roman_Å end_ARG in the case of zigzag orientation shown in Fig. 6(b). These spurious states referred to as ghost or phantom modes, are associated with virtual molecular orbitals of the barrier area and might lead to an inaccurate description of the tunneling [43]. Interestingly, the ghost modes are an apparently unavoidable consequence of large numerical basis sets, so an allegedly increase in computational accuracy might increase the number of ghost states [43]. A separation of ghost modes and true CBS modes is challenging. In our case, the ghost states do have large κ𝜅\kappaitalic_κ, are heavily damped, and thus only contribute negligibly to the transport. Generally, they may exhibit very slow decay and can dominate the transport leading to fundamental qualitative and quantitative errors. We suggest using the complex band structure and its validity concerning the ghost states as a first test before the computationally demanding Landauer approach is applied.

The I𝐼Iitalic_I-V𝑉Vitalic_V characteristics and high PVCR values reported for the tunnel diodes are based on the assumption of coherent tunneling transport. This means that mechanisms like electron-electron scattering or electron-phonon scattering are not included in the calculations. Electron-phonon scattering is a well-known dissipation mechanism, particularly efficient in 2D semiconductors compared to 3D[44]. It can influence current flow in devices based on 2D materials [45]. While software packages like QuantumATK can include electron-phonon scattering, the computational cost is significant and often limited to very simple models[46]. Therefore, for our presented tunnel diodes, such detailed simulations are currently not practical. Consequently, the peak current in Fig. 5 and the PVCR obtained here represent upper limits within the coherent tunneling framework. Introducing inelastic scattering mechanisms, like electron-phonon interactions, is expected to modify the I𝐼Iitalic_I-V𝑉Vitalic_V characteristics. This would likely decrease the peak current and increase the valley current, resulting in a lower PVCR value. It is important to note that other factors, such as defects, interface atomic mixing, and interactions with surrounding materials, can also influence the I𝐼Iitalic_I-V𝑉Vitalic_V characteristics. However, the overall trends observed in the I𝐼Iitalic_I-V𝑉Vitalic_V curves are expected to remain largely unchanged.

Eventually, we discuss the anticipated I𝐼Iitalic_I-V𝑉Vitalic_V characteristics of tunnel diodes based on other cold metals listed in Table 1. As mentioned, the N𝑁Nitalic_N-type NDR effect is expected to be common in tunnel diodes using these materials. For example, TaSi2N4, which is isoelectronic to NbSi2N4, is expected to exhibit qualitatively similar I𝐼Iitalic_I-V𝑉Vitalic_V curves. However, quantitative differences, such as peak current density, valley voltage, and valley current, are likely due to TaSi2N4’s larger metallic band width (Wmsubscript𝑊mW_{\mathrm{m}}italic_W start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT) and slightly smaller internal and external band gaps. Similarly, Ge-based compounds are expected to exhibit analogous I𝐼Iitalic_I-V𝑉Vitalic_V characteristics. On the other hand, tunnel diodes based on P-containing compounds like NbSi2P4 may exhibit significantly smaller PVCR values, primarily due to the vanishing internal band gap in these compounds. This insight into the expected behavior of tunnel diodes based on various cold metals offers a glimpse into the diverse possibilities for tailoring their performance in electronic applications and underscores the role of material properties in shaping device behavior.

IV Conclusions

In conclusion, our computational investigation of prototype lateral heterojunction tunnel diodes unveils the remarkable potential of cold metallic MA2Z4 (M=Nb, Ta; A=Si, Ge; Z=N, P) compounds for harnessing the NDR phenomenon with very high PVCR values orders of magnitude higher than in conventional NDR tunnel diodes. By considering the exemplary cold metal NbSi2N4 as the prototype electrode material and HfSi2N4 as the tunnel barrier, our calculations consistently demonstrate the achievement of N𝑁Nitalic_N-type NDR behavior in both armchair and zigzag-oriented tunnel diodes. Moreover, our findings reveal intriguing differences between these orientations. Zigzag-oriented tunnel diodes exhibit significantly higher peak current densities, indicating their potential for high-speed electronic applications, while armchair-oriented diodes achieve very high PVCR values. This duality in performance highlights the versatility of MA2Z4 materials and their promising role in enabling future electronic devices with enhanced functionality and efficiency. Our findings not only broaden our understanding of the NDR effect in cold metal based heterojunction tunnel diodes, but also pave the way for exciting new directions in materials engineering for next-generation electronics.

Acknowledgements.
This work was supported by SFB CRC/TRR 227 of Deutsche Forschungsgemeinschaft (DFG) and by the European Union (EFRE) via Grant No: ZS/2016/06/79307.

References

  • Berger and Ramesh [2011] P. R. Berger and A. Ramesh, Negative differential resistance devices and circuits, in Comprehensive semiconductor science and technology (Elsevier Inc., 2011) pp. 176–241.
  • Jo et al. [2021] S. B. Jo, J. Kang, and J. H. Cho, Recent advances on multivalued logic gates: a materials perspective, Advanced Science 8, 2004216 (2021).
  • Wang et al. [2017] S. Wang, A. Pan, C. Grezes, P. K. Amiri, K. L. Wang, C. O. Chui, and P. Gupta, Leveraging nmos negative differential resistance for low power, high reliability magnetic memory, IEEE Transactions on Electron Devices 64, 4084 (2017).
  • van der Wagt [1999] J. P. A. van der Wagt, Tunneling-based sram, Proceedings of the IEEE 87, 571 (1999).
  • Karda et al. [2009] K. Karda, J. Brockman, S. Sutar, A. Seabaugh, and J. Nahas, One-transistor bistable-body tunnel sram, in 2009 IEEE International Conference on IC Design and Technology (IEEE, 2009) pp. 233–236.
  • Micheel and Paulus [1990] L. J. Micheel and M. J. Paulus, Differential multiple-valued logic using resonant tunneling diodes, in Proceedings of the Twentieth International Symposium on Multiple-Valued Logic (IEEE Computer Society, 1990) pp. 189–190.
  • Lin [1994] H. Lin, Resonant tunneling diodes for multi-valued digital applications, in Proceedings of 24th International Symposium on Multiple-Valued Logic (ISMVL’94) (IEEE, 1994) pp. 188–195.
  • Jin et al. [2004] N. Jin, S.-Y. Chung, R. M. Heyns, P. R. Berger, R. Yu, P. E. Thompson, and S. L. Rommel, Tri-state logic using vertically integrated si-sige resonant interband tunneling diodes with double ndr, IEEE electron device letters 25, 646 (2004).
  • Esaki [1958] L. Esaki, New phenomenon in narrow germanium p- n junctions, Physical review 109, 603 (1958).
  • Esaki and Tsu [1970] L. Esaki and R. Tsu, Superlattice and negative differential conductivity in semiconductors, IBM Journal of Research and Development 14, 61 (1970).
  • Ramesh et al. [2012] A. Ramesh, P. R. Berger, and R. Loo, High 5.2 peak-to-valley current ratio in si/sige resonant interband tunnel diodes grown by chemical vapor deposition, Applied Physics Letters 100 (2012).
  • Bruce et al. [2020a] A. V. Bruce, S. Liu, J. N. Fry, and H.-P. Cheng, Insights into negative differential resistance in mos 2 esaki diodes: A first-principles perspective, Physical Review B 102, 115415 (2020a).
  • Fung et al. [2011] W. Y. Fung, L. Chen, and W. Lu, Esaki tunnel diodes based on vertical si-ge nanowire heterojunctions, Applied Physics Letters 99 (2011).
  • Duschl et al. [1999] R. Duschl, O. Schmidt, G. Reitemann, E. Kasper, and K. Eberl, High room temperature peak-to-valley current ratio in si based esaki diodes, Electronics Letters 35, 1111 (1999).
  • Schmid et al. [2012] H. Schmid, C. Bessire, M. T. Björk, A. Schenk, and H. Riel, Silicon nanowire esaki diodes, Nano letters 12, 699 (2012).
  • Chow et al. [1992] D. Chow, J. Schulman, E. Özbay, and D. Bloom, Investigation of in0. 53ga0. 47as/alas resonant tunneling diodes for high speed switching, Applied physics letters 61, 1685 (1992).
  • Tsai et al. [1994] H. Tsai, Y. Su, H. Lin, R. Wang, and T. Lee, Pn double quantum well resonant interband tunneling diode with peak-to-valley current ratio of 144 at room temperature, IEEE electron device letters 15, 357 (1994).
  • Chung et al. [2004] S.-Y. Chung, N. Jin, P. R. Berger, R. Yu, P. E. Thompson, R. Lake, S. L. Rommel, and S. K. Kurinec, Three-terminal si-based negative differential resistance circuit element with adjustable peak-to-valley current ratios using a monolithic vertical integration, Applied physics letters 84, 2688 (2004).
  • Chen et al. [2009] S.-L. Chen, P. B. Griffin, and J. D. Plummer, Negative differential resistance circuit design and memory applications, IEEE transactions on electron devices 56, 634 (2009).
  • Gan et al. [2007] K.-J. Gan, C.-S. Tsai, and W.-L. Sun, Fabrication and application of mos-hbt-ndr circuit using standard sige process, Electronics Letters 43, 1 (2007).
  • Duane et al. [2003] R. Duane, A. Mathewson, and A. Concannon, Bistable gated bipolar device, IEEE Electron Device Letters 24, 661 (2003).
  • Fang et al. [2022] L. Fang, C. Qiu, H. Zhang, Y. Hu, and L.-M. Peng, Giant negative differential resistance effect caused by cutting off acceptable quantum states in carbon nanotube tunneling devices, Advanced Electronic Materials 8, 2101314 (2022).
  • Sant and Schenk [2017] S. Sant and A. Schenk, The effect of density-of-state tails on band-to-band tunneling: Theory and application to tunnel field effect transistors, J. Appl. Phys. 122, 135702 (2017).
  • Bizindavyi et al. [2018] J. Bizindavyi, A. S. Verhulst, Q. Smets, D. Verreck, B. Sorée, and G. Groeseneken, Band-tails tunneling resolving the theory-experiment discrepancy in esaki diodes, IEEE J. Electron Devices Soc. 6, 633 (2018).
  • Schenk and Sant [2020] A. Schenk and S. Sant, Tunneling between density-of-state tails: Theory and effect on esaki diodes, J. Appl. Phys. 128, 014502 (2020).
  • Şaşıoğlu and Mertig [2023] E. Şaşıoğlu and I. Mertig, Theoretical prediction of semiconductor-free negative differential resistance tunnel diodes with high peak-to-valley current ratios based on two-dimensional cold metals, ACS Applied Nano Materials 6, 3758 (2023).
  • Hong et al. [2020] Y.-L. Hong, Z. Liu, L. Wang, T. Zhou, W. Ma, C. Xu, S. Feng, L. Chen, M.-L. Chen, D.-M. Sun, et al., Chemical vapor deposition of layered two-dimensional mosi2n4 materials, Science 369, 670 (2020).
  • Yin et al. [2023] Y. Yin, Q. Gong, M. Yi, and W. Guo, Emerging versatile two-dimensional mosi2n4 family, Advanced Functional Materials 33, 2214050 (2023).
  • Ichinose et al. [2022] N. Ichinose, M. Maruyama, Z. L. Takato Hotta, R. Canton-Vitoriaa, S. Okada, F. Zeng, T. T. Feng Zhang, K. Watanabe, and R. Kitaura, Two-dimensional atomic-scale ultrathin lateral heterostructures, arXiv preprint arXiv:2208.12696v2  (2022).
  • Bruce et al. [2020b] A. V. Bruce, S. Liu, J. N. Fry, and H.-P. Cheng, Insights into negative differential resistance in mos2 esaki diodes: A first-principles perspective, Phys. Rev. B 102, 115415 (2020b).
  • Hinsche and Thygesen [2017] N. F. Hinsche and K. S. Thygesen, Electron–phonon interaction and transport properties of metallic bulk and monolayer transition metal dichalcogenide tas2, 2D Materials 5, 015009 (2017).
  • Smidstrup et al. [2020] S. Smidstrup, T. Markussen, P. Vancraeyveld, J. Wellendorff, J. Schneider, T. Gunst, B. Verstichel, D. Stradi, P. A. Khomyakov, U. G. Vej-Hansen, et al., Quantumatk: An integrated platform of electronic and atomic-scale modelling tools, J. Phys: Condens. Matter 32, 015901 (2020).
  • Troullier and Martins [1991] N. Troullier and J. L. Martins, Efficient pseudopotentials for plane-wave calculations, Physical review B 43, 1993 (1991).
  • Perdew et al. [1996] J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized gradient approximation made simple, Physical review letters 77, 3865 (1996).
  • Büttiker et al. [1985] M. Büttiker, Y. Imry, R. Landauer, and S. Pinhas, Generalized many-channel conductance formula with application to small rings, Phys. Rev. B 31, 6207 (1985).
  • [36] See Supplemental Material at [URL-will-be-inserted-by-publisher] for further insights into the characteristics of the devices.
  • Heil et al. [2018] C. Heil, M. Schlipf, and F. Giustino, Quasiparticle gw band structures and fermi surfaces of bulk and monolayer nbs2, Phy. Rev. B 98, 075120 (2018).
  • Kim and Son [2017] S. Kim and Y.-W. Son, Quasiparticle energy bands and fermi surfaces of monolayer nbse2, Phy. Rev. B 96, 155439 (2017).
  • Belashchenko et al. [2004] K. D. Belashchenko, E. Y. Tsymbal, M. van Schilfgaarde, D. A. Stewart, I. I. Oleynik, and S. S. Jaswal, Effect of interface bonding on spin-dependent tunneling from the oxidized co surface, Phys. Rev. B 69, 174408 (2004).
  • Terzibaschian and Enderlein [1986] T. Terzibaschian and R. Enderlein, The irreducible representations of the two-dimensional space groups of crystal surfaces. theory and applications, Physica Status Solidi (b) 133, 443 (1986).
  • Dresselhaus and Dresselhaus [2008] M. S. Dresselhaus and A. J. Dresselhaus, G., Group theory: Application to the Physics of Condensed Matter (Springer-Verlag,, Berlin-Heidelberg, 2008).
  • Hinsche et al. [2010] N. F. Hinsche, M. Fechner, P. Bose, S. Ostanin, J. Henk, I. Mertig, and P. Zahn, Strong influence of complex band structure on tunneling electroresistance: A combined model and ab initio study, Phys. Rev. B 82, 214110 (2010).
  • Herrmann et al. [2010] C. Herrmann, G. C. Solomon, J. E. Subotnik, V. Mujica, and M. A. Ratner, Ghost transmission: How large basis sets can make electron transport calculations worse, The Journal of Chemical Physics 132, 024103 (2010).
  • Cheng et al. [2020] L. Cheng, C. Zhang, and Y. Liu, Why two-dimensional semiconductors generally have low electron mobility, Phys. Rev. Lett. 125, 177701 (2020).
  • Afzalian [2021] A. Afzalian, Ab initio perspective of ultra-scaled cmos from 2d-material fundamentals to dynamically doped transistors, NPJ 2D Mater. Appl. 5, 1 (2021).
  • Smidstrup et al. [2019] S. Smidstrup, T. Markussen, P. Vancraeyveld, J. Wellendorff, J. Schneider, T. Gunst, B. Verstichel, D. Stradi, P. A. Khomyakov, U. G. Vej-Hansen, M.-E. Lee, S. T. Chill, F. Rasmussen, G. Penazzi, F. Corsetti, A. Ojanperä, K. Jensen, M. L. N. Palsgaard, U. Martinez, A. Blom, M. Brandbyge, and K. Stokbro, QuantumATK: an integrated platform of electronic and atomic-scale modelling tools, J. Phys. Condens. Matter 32, 015901 (2019).

Supplemental Material: Computational design of tunnel diodes with negative differential resistance and ultrahigh peak-to-valley current ratio based on two-dimensional cold metals: The case of NbSi2N4/HfSi2N4/NbSi2N4 lateral heterojunction diode

P. Bodewei1, E. Şaşıoğlu1, N. F. Hinsche1, I. Mertig1 1Institute of Physics, Martin Luther University Halle-Wittenberg, 06120 Halle (Saale), Germany

Refer to caption
Refer to caption
Figure 7: Schematic illustration of the atomic structure of the NbSi2N4/HfSi2N4/NbSi2N4 lateral heterojunction tunnel diode device in armchair orientation. The tunnel barrier width is 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG. The side view of the device is shown on the right side. Different atomic components are represented by distinct colors.
Refer to caption
Refer to caption
Figure 8: Schematic illustration of the atomic structure of the NbSi2N4/HfSi2N4/NbSi2N4 lateral heterojunction tunnel diode device in zigzag orientation. The tunnel barrier width is 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG. The side view of the device is shown on the right side. Different atomic components are represented by distinct colors.
Refer to caption
Figure 9: Device density of states (DDOS) of the armchair device with the 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG barrier for a bias voltage of 0 Vtimes0V0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG.
Refer to caption
Figure 10: Convergence test for the k𝑘kitalic_k-point sampling for the calculation of the I𝐼Iitalic_I-V𝑉Vitalic_V-curve for the armchair device with 30 Åtimes30angstrom30\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG barrier thickness for negative bias applied with constant cut-offs, pseudo- and exchange-correlation-potentials
Refer to caption
Figure 11: *

(a) 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG barrier

Refer to caption
Figure 12: *

(b) 30 Åtimes30angstrom30\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG barrier

Figure 13: I𝐼Iitalic_I-V𝑉Vitalic_V-characteristics for positive and negative bias for the devices with 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG and 30 Åtimes30angstrom30\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG tunnel barrier

In Fig. 13 the I𝐼Iitalic_I-V𝑉Vitalic_V-curves for positive and negative bias are shown. While the I𝐼Iitalic_I-V𝑉Vitalic_V-curve for the zigzag device with 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG and 30 Åtimes30angstrom30\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG barrier thickness are perfectly antisymmetric, we see that the I𝐼Iitalic_I-V𝑉Vitalic_V-curves for devices in armchair direction show a slight asymmetry. The reason for that is the interface potential, building up between NbSi2N4 and HfSi2N4. Geometrically, we can see that the termination of the interface between the right electrode and left electrode with the barrier is different in armchair devices. Thus, a different interface potential at both interfaces builds up, leading to different current densities for positive and negative bias voltages, respectively.

Refer to caption
Figure 14: *

(a) armchair device with 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG tunnel barrier

Refer to caption
Figure 15: *

(b) zigzag device with 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG tunnel barrier

Figure 16: Transmission of the armchair and zigzag device for different energies and k-points in the direction perpendicular to the (a) armchair and (b) zigzag transport direction

In Fig. 16 we see the transmission at different k𝑘kitalic_k-points for different energies. Firstly, we see that there is no certain k𝑘kitalic_k-point dominating the transmission, i.e. the transmission is distributed throughout the whole range of k𝑘kitalic_k-points for both the armchair and the zigzag device. Further, we see that for higher energies of about 1 eVtimes1eV1\text{\,}\mathrm{e}\mathrm{V}start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_eV end_ARG we see a strong contribution of the transmission at the ΓΓ\Gammaroman_Γ-point.

Refer to caption
Figure 17: *

0.0 Vtimes0.0V0.0\text{\,}\mathrm{V}start_ARG 0.0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG Bias

Refer to caption
Figure 18: *

0.2 Vtimes0.2V0.2\text{\,}\mathrm{V}start_ARG 0.2 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG Bias

Refer to caption
Figure 19: *

0.75 Vtimes0.75V0.75\text{\,}\mathrm{V}start_ARG 0.75 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG Bias

Figure 20: k𝑘kitalic_k-point- and energy-resolved transmission spectrum of the armchair device with 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG barrier for different bias voltages. The vertical red lines show the bias voltage window.
Refer to caption
Figure 21: *

0.0 Vtimes0.0V0.0\text{\,}\mathrm{V}start_ARG 0.0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG Bias

Refer to caption
Figure 22: *

0.2 Vtimes0.2V0.2\text{\,}\mathrm{V}start_ARG 0.2 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG Bias

Refer to caption
Figure 23: *

0.75 Vtimes0.75V0.75\text{\,}\mathrm{V}start_ARG 0.75 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG Bias

Figure 24: k𝑘kitalic_k-point- and energy-resolved transmission spectrum of the zigzag device with 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG barrier for different bias voltages. The vertical red lines show the bias voltage window.

In Fig. 20 and Fig. 24 the k𝑘kitalic_k- and energy-resolved transmission spectrum is shown for three different bias voltages. Each black curve represents the transmission for a single k𝑘kitalic_k-point, and the red lines show the bias window ([Vbias2,Vbias2]subscript𝑉𝑏𝑖𝑎𝑠2subscript𝑉𝑏𝑖𝑎𝑠2\left[-\frac{V_{bias}}{2},\frac{V_{bias}}{2}\right][ - divide start_ARG italic_V start_POSTSUBSCRIPT italic_b italic_i italic_a italic_s end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG , divide start_ARG italic_V start_POSTSUBSCRIPT italic_b italic_i italic_a italic_s end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ]). In the following, we will call T(E)𝑇𝐸T(E)italic_T ( italic_E ) the transmission and T(E,k)𝑇𝐸𝑘T(E,k)italic_T ( italic_E , italic_k ) the k𝑘kitalic_k-resolved transmission. Since we are calculating the current within the Landauer-Büttiker approach, the main quantity is the transmission, which can be calculated by integrating T(E,k)𝑇𝐸𝑘T(E,k)italic_T ( italic_E , italic_k ) over all k𝑘kitalic_k-points. This is shown by the blue curve. Finally, one obtains the I𝐼Iitalic_I-V𝑉Vitalic_V-characteristics by integrating T(E)𝑇𝐸T(E)italic_T ( italic_E ) over a broad energy range. Here, we show k𝑘kitalic_k- and E𝐸Eitalic_E-resolved transmissions for the 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG-barrier armchair (Fig. 20) and the 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG-barrier zigzag device (Fig. 24). The bias voltages of 0.0 Vtimes0.0V0.0\text{\,}\mathrm{V}start_ARG 0.0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG, 0.2 Vtimes0.2V0.2\text{\,}\mathrm{V}start_ARG 0.2 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG, and 0.75 Vtimes0.75V0.75\text{\,}\mathrm{V}start_ARG 0.75 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG were chosen, as the maximum of the current in the armchair device occurs at 0.2 Vtimes0.2V0.2\text{\,}\mathrm{V}start_ARG 0.2 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG and the maximum of the zigzag around 0.7 Vtimes0.7V0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG. For the armchair device, we deduce that the area under the transmission curve is significantly higher at 0.2 Vtimes0.2V0.2\text{\,}\mathrm{V}start_ARG 0.2 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG than for 0.75 Vtimes0.75V0.75\text{\,}\mathrm{V}start_ARG 0.75 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG, underpinning the maxima in the I𝐼Iitalic_I-V𝑉Vitalic_V-curves (fig. 13). For the zigzag device, we spot an increase in the area under the transmission curve for the higher bias voltages. Furthermore, one notices that the areas under the transmission curves for the armchair device are much smaller than for the zigzag device, supporting the difference in the current densities absolute values for both devices.

Refer to caption
Figure 25: Temperature dependence of the armchair-device with the 21 Åtimes21angstrom21\text{\,}\mathrm{\SIUnitSymbolAngstrom}start_ARG 21 end_ARG start_ARG times end_ARG start_ARG roman_Å end_ARG barrier. An increase in temperature results in a slight decrease in the peak current as well as a smaller current valley region.