Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Full loss compensation in hybrid plasmonic waveguides under electrical pumping

Open Access Open Access

Abstract

Surface plasmon polaritons (SPPs) give an opportunity to break the diffraction limit and design nanoscale optical components, however their practical implementation is hindered by high ohmic losses in a metal. Here, we propose a novel approach for efficient SPP amplification under electrical pumping in a deep-subwavelength metal-insulator-semiconductor waveguiding geometry and numerically demonstrate full compensation for the SPP propagation losses in the infrared at an exceptionally low pump current density of 0.8 kA/cm2. This value is an order of magnitude lower than in the previous studies owing to the thin insulator layer between a metal and a semiconductor, which allows injection of minority carriers and blocks majority carriers reducing the leakage current to nearly zero. The presented results provide insight into lossless SPP guiding and development of future high dense nanophotonic and optoelectronic circuits.

© 2015 Optical Society of America

1. Introduction

Surface plasmon polaritons (SPPs) being surface electromagnetic waves propagating along the interface between a metal and an insulator give a unique opportunity to overcome the conventional diffraction limit and thus can be used for light manipulation at the nanoscale [1]. This gives plasmonic components a significant advantage over their photonic counterparts in the integration density and strength of light-matter interaction. However, high mode confinement is paid off by a significant localization of the SPP electromagnetic field in the metal and the resulting propagation length in nanoscale plasmonic waveguides does not exceed a few tens of micrometers due to absorption in the metal [1–3]. Nevertheless, it is possible to overcome this limitation by compensating ohmic losses with optical gain in the adjacent dielectric. Full loss compensation was successfully demonstrated by optical pumping [4–6], which is easily implemented in a laboratory, but is impractical due to its very poor energy efficiency, stray illumination, and necessity of an external high power pump light source. In this regard, an efficient electrical pumping scheme is strongly needed for practical realization of SPP guides and circuits [7–9].

In this paper, we propose a novel SPP amplification scheme based on minority carrier injection in metal-insulator-semiconductor (MIS) structures and demonstrate numerically full loss compensation in an electrically pumped active hybrid plasmonic waveguide. In contrast to the techniques based on heterostructures and quantum wells [10], our approach gives a possibility to bring the active gain medium to the metal surface at a distance of a few nanometers. At the same time, a thin insulator layer can efficiently block the majority carrier current, which is quite high in metal-semiconductor Schottky-barrier-diode SPP amplifiers [11]. This guarantees high mode confinement to the active region and very low threshold currents.

2. Operating principle

Realization of an electrical pumping scheme for loss compensation in SPP waveguides and cavities is a significant challenge, since there are at least two limitations in the development of plasmonic components. Firstly, plasmonics gives a possibility to reduce the mode size well below the diffraction limit, but, in order to use this opportunity, the component size should be comparable with the mode size. Accordingly, one has to use the SPP supporting metal-dielectric or metal-semiconductor interface as an electrical contact. Secondly, a limited choice of low-loss materials with negative real part of permittivity poses a serious challenge for efficient carrier injection, because gold, silver and copper typically form Schottky contacts to direct bandgap semiconductors. As a result, either the modal gain is rather low for full loss compensation, or the threshold current density is quite high [12], which leads to significant power consumption and low energy efficiency of the amplification scheme.

Our technique based on an electrically pumped MIS structure (Fig. 1) is remarkably different from the previously proposed approaches [10, 11, 13–15]. First, SPP propagation losses are reduced by placing a very thin layer of a low refractive index insulator between the metal and semiconductor. In this case, the SPP electromagnetic field is pushed into the insulator reducing the portion of the SPP mode in the metal, which, however, does not impair the mode confinement [3]. Second, and more important, a thin insulator layer can play a dual role: it is semitransparent for electrons moving from the metal to semiconductor, which ensures favourable conditions for efficient minority carrier injection, and blocks the hole current. Such a unique feature gives a possibility to suppress high leakage currents [16], which are unavoidable in metal-semiconductor Schottky contacts [14], and not to form an ohmic contact to the semiconductor at the SPP supporting interface, as opposed to double-heterostructure and quantum-well SPP amplifiers [12]. At the same time, the proposed scheme provides high density of non-equilibrium carriers right near the metal surface, at the distance equal to the thickness of the insulator layer. At a high bias voltage, it becomes possible to create a sufficiently high electron concentration in the semiconductor and to satisfy the condition for population inversion [17], which provides gain for the plasmonic mode propagating in the MIS waveguide.

 figure: Fig. 1

Fig. 1 SPP amplification schemes based on a Schottky barrier diode [11] and tunnel MIS contact. Top panels: Energy band diagrams for the Schottky contact between gold and indium arsenide in equilibrium (a) and at high forward bias (b). The Schottky barrier height for electrons φBn at the Au/InAs contact is negative owing to the large density of surface states and, under high forward bias, electrons (minority carriers in p-type InAs) are freely injected in the bulk of the semiconductor. However, majority carriers (holes) also pass across the Au/InAs contact without any resistance, which results in a high leakage current. Bottom panels: Energy band diagrams for the tunnel Au/HfO2/InAs MIS structure in equilibrium (c) and at high forward bias (d). If the barrier height for holes χVI is substantially greater than that for electrons (φMI), the insulator layer can efficiently block majority carriers (holes), but be semi-transparent for minority carriers (electrons) escaping from the metal.

Download Full Size | PDF

Previously, an attempt to describe the carrier transport in an electrically pumped hybrid plasmonic waveguide was made in [18]. However, the authors assumed that both electrons and holes pass freely through the insulating layer and used the thermionic emission boundary conditions (literally the same as in the Schottky junction theory [9, 19]). This is erroneous in the presence of the insulating barrier and contradicts the well-established theory of current transfer through thin insulating films [20, 21]. In fact, the insulating barrier strongly suppresses the thermionic current, while the main current component is due to tunneling.

The electron and hole tunnel injection currents Jt,cb and Jt,vb across the MIS contact can be calculated by integrating the flux of carriers incident on the contact timed by the transparency of the insulating barrier (see Appendix A). According to the energy band diagram shown in Fig. 1 (c,d), the ratio of the electron tunneling probability Dcb to the hole tunneling probability Dvb can be easily estimated as

DcbDvd=exp[23/2mi1/2dih(χVI1/2φMI1/2)],
where χVI is the valence band offset between the insulator and semiconductor, and φMI is the barrier height for electrons at the metal-insulator interface. As evident from this simple expression, in order to block only one type of carriers, the effective barrier height for holes (χVI) must be substantially greater than that for electrons (φMI) or vice versa. Hafnium oxide appears to be a very promising material for electron injection into the p-type III–V materials thanks to the low tunneling mass mi = 0.1m0 [22] and large valence band and small conduction band offsets. Both theoretical [23] and experimental [24] studies report χCI = 2.4 eV and χVI = 3.2 eV for the HfO2/InAs interface. At the same time, φMI = 2.6 eV for the gold contact [23], and the ratio Dcb/Dvb equals five, which is high enough for practical applications.

We next explore the carrier transport in the semiconductor. It is important to note that the electric potential is abruptly screened near the interface between the insulator and heavily doped semiconductor, which directly follows from the solution of Poisson’s equation. At a doping density of NA = 1018 cm−3, the estimated screening length is equal to 2 nm (see Appendix B), while the electron mean free path limited by impurity scattering equals lfp = 10 nm. This implies that electrons move ballistically through the screening region. Out of the screening region, the driving electric field is weak and the electron transport is governed by diffusion. Accordingly, the density distribution of the injected electrons is given by

n(z)J0,neτRDnexp(zLD)+n0,
where J0,n is the electron current density at the insulator-semiconductor interface, LD=τRDn is the diffusion length, and n0 is the equilibrium electron density. LD is the most critical parameter determining the distribution of carrier density in InAs, and, consequently, the material gain profile. It depends on the electron mobility through the electron diffusion coefficient Dn and on the recombination rate through the electron lifetime τR. Equation (2) shows that, in order to compensate for the SPP propagation losses efficiently, LD should be comparable with the half of the SPP penetration depth in InAs (500 nm at λ = 3.22 µm) or greater than the latter. The electron mobility in p-type InAs is mainly limited by charged impurity scattering and is about µn = 2 × 103 cm2V−1s−1 at 77 K at a doping level of NA = 1018 cm−3 [25]. The main contributions to the carrier recombination in InAs at 77 K come from the nonradiative Auger process RA = (Cnn + Cpp)(npn0p0) and radiative recombination due to spontaneous emission Rsp = FpBsp(npn0p0), where n0 and p0 are the equilibrium electron and hole densities, Cn = Cp = 10−27 cm6/s are the Auger coefficients [26], Bsp is the bulk radiative recombination coefficient calculated to be 4 × 10−10 cm3s−1 and Fp is the Purcell factor, which is of the order of unity in plasmonic waveguides at mid-infrared wavelengths [11], as discussed in the next section. Since in a p-type semiconductor the density of electrons is much smaller than that of holes even at high injection levels, τR=[CpNA2+BspNA]1=700ps, which corresponds to a diffusion length of LD = 950 nm ensuring an excellent overlap between the distributions of the material gain (that is nearly proportional to the carrier concentration) and the SPP field.

The electron current J0,n at the semiconductor-insulator junction is mainly determined by quantum mechanical tunneling through the insulating layer. In the case of an ideal interface, J0,n is simply equal to the tunnel current Jt,cb. However, the presence of intrinsic and extrinsic surface states at the semiconductor-insulator interface has a significant impact on the characteristics of the tunnel MIS contact (see Appendix C). First, surface states act as charge storage centers and affect the voltage drop across the insulator layer and band bending in the semiconductor. Second, these states are recombination centers [26] and reduce the efficiency of electron injection. The electron current at the insulator-semiconductor interface J0,n is smaller than the tunnel injection current Jt,cb by the value of the surface recombination current Jsr,n. Third, direct tunneling from the metal to semiconductor through the surface states is also possible [27, 28]. This process contributes only to the majority carrier current and reduces the energy efficiency of the SPP amplification scheme. Fortunately, intrinsic surface states in InAs have a moderate density ρss ≃ 3 × 1012 cm−2eV−1 [29]. Deposition of metal [30] or insulator [31] layers produces additional defects and ρss can be increased up to 1014 cm−2eV−1 at HfO2/InAs interfaces [32]. Nevertheless, both intrinsic and extrinsic states can be treated by surface passivation before HfO2 deposition. The reported values of ρss at the trimethylaluminum-treated HfO2/InAs interface is about 1013 cm−2eV−1 [33], while advanced oxygen-termination and surface reconstruction techniques give a possibility to reduce ρss down to 2 × 1011 cm−2eV−1 [34], which eliminates the influence of surface states on carrier transport in high quality samples.

3. Results and discussion

3.1. Active hybrid plasmonic waveguide in the passive regime

In order to achieve strong SPP mode localization and efficiently inject both electrons and holes into the active semiconductor region, a modified T-shaped plasmonic waveguide approach [11] has been implemented [Fig. 2(a)]. InAs rib with an acceptor concentration of 1018 cm−3 on InAs substrate is covered by a thin HfO2 layer and surrounded by a low refractive index dielectric (SiO2) to confine optical modes in the lateral direction. Finally, a metal layer is placed on top of the semiconductor-insulator structure to form an SPP supporting interface and a tunnel MIS contact for electron injection, while the substrate plays a role of the ohmic contact for majority carrier (hole) injection.

 figure: Fig. 2

Fig. 2 Schematic of a T-shaped hybrid plasmonic waveguide based on the Au/HfO2/InAs MIS structure, w is the waveguide width, di is the thickness of the low refractive index insulator layer between the metal and semiconductor, and H is the waveguide height. (b) Distribution of the normalized energy density per unit length of the waveguide for the fundamental TM00 mode at λ = 3.22 µm, H = 2.5 µm, w = 400 nm and di = 3 nm. The dielectric functions of the materials are as follows: εSiO2=2.00 [35], εHfO2=3.84 [36], εInAs = 12.38 [37] and εAu = −545+38i [11].

Download Full Size | PDF

Two-dimensional eigenmode simulations using the finite element method (COMSOL Multiphysics software) reveal that the waveguide depicted in Fig. 2(a) supports a deep-subwavelength TM00 SPP mode [Fig. 2(b)]. For the chosen waveguide dimensions, at a free space wavelength of 3.22 µm, which appears to be optimal for SPP amplification (see the next section for details), the effective index of the TM00 mode is equal to 2.799 and the propagation length in the passive regime assuming the semiconductor to be lossless is about 66 µm corresponding to a modal loss of 2Imβpsv = 152 cm−1. The other modes supported by the T-shaped waveguide are very leaky photonic modes. In spite of poor field confinement to the lossy metal, propagation lengths of the photonic TM10, TE00, and TE10 modes are 5.8 µm, 12.8 µm and 2.7 µm, respectively, which are much shorter than that of the TM00 mode due to leakage into the high refractive index substrate. The rib height of H = 2.5 µm is close to the optimum, since, as H decreases, the radiation loss of the plasmonic mode becomes substantially greater than the absorption in the metal and the SPP propagation length decreases down to 39 µm at H = 2 µm. On the other hand, greater rib heights do not provide a significant loss reduction, while high-aspect-ratio structures are difficult in fabrication. For the selected optimal geometrical parameters, the SPP mode demonstrates an exceptionally high level of confinement [38] within the InAs region of more than 95% provided by the small mode width and high group index. At the same time, the spontaneous emission enhancement by the Purcell effect does not appreciably decrease the electron lifetime. Thanks to the moderate SPP confinement in the vertical direction, the Purcell factor Fp(di) in InAs near the metal surface does not exceed 1.28, which results in a maximum of 7% shortening of τR compared to the bulk case.

3.2. Full loss compensation in a deep-subwavelength waveguide

Applying a negative voltage to the top Au electrode of the active plasmonic waveguide shown in Fig. 2(a), one injects electrons into the p-type InAs region. This creates a high density of non-equilibrium electrons in the semiconductor rib, which reduces absorption in InAs [17]. As the bias voltage increases, the quasi-Fermi level for electrons shifts upward towards the conduction band. When the energy difference between quasi-Fermi levels for electrons and holes exceeds the energy ω of the SPP quantum, InAs starts to compensate for the SPP propagation losses. The modal gain G is given by the overlap integral of the material gain profile g(y, z) and the electric field distribution of the SPP mode [38]:

G=cε0nInAsw/2+w/2dydiHdzg(y,z)|E(y,z)|22+dy+dzPz(y,z)2Imβpsv.

Here, E(y, z) and Pz(y, z) are the complex amplitudes of the electric field and the local power density of the SPP mode, respectively, and 2Imβpsv is attributed to the SPP radiation and ohmic losses. In turn, the material gain coefficient g(y, z) is related to the carrier quasi-Fermi levels Fn(y, z) and Fp(y, z) by the integral over transitions between the energy levels in the conduction and valence bands separated by the energy ω (see Appendix D). In heavily doped semiconductors, the material gain strongly depends on the impurity concentration [39], so are the electrical properties (LD, µn, and τR). The optimal acceptor concentration is found to be 1018 cm−3. As the doping concentration increases, the Auger recombination rate and the impurity scattering rate rapidly increase greatly reducing the electron diffusion length. On the other hand, at lower doping levels, the number of free electron states in the valence band is insufficient to provide a pronounced gain upon electrical injection. At the given impurity concentration and a reasonably high density of injected electrons of 2×1016 cm−3 (which corresponds to FnEc = kBT), the material gain spectrum exhibits a maximum of 310 cm−1 at ω=0.385 eV (λ = 3.22 µm), sufficiently high for the net SPP amplification.

Figure 3 shows the simulated gain-current characteristic of the plasmonic TM00 mode in the active hybrid plasmonic waveguide cooled down to 77 K for the HfO2/InAs interfaces of different quality. All curves exhibit the same trend. At zero bias, the concentration of holes in the semiconductor is many orders of magnitude greater than that of electrons, InAs strongly absorbs the SPP propagating in the waveguide and the modal loss is two-and-a-half times greater than in the case of the lossless (e.g. wide bandgap) semiconductor. As the bias voltage increases, electrons are injected in the InAs rib [Fig. 4(a)], which reduces absorption in the semiconductor near the MIS contact, but still, InAs strongly absorbs at a distance greater than LD/2 [Fig. 4(b)]. Nevertheless, the SPP modal loss steadily decreases with the injection current (this region is shown in blue in Fig. 3). At some point [green curve in Fig. 4(b)], the overlap between the distributions of the material gain and the SPP field is zero and, at higher bias voltages, the gain in InAs produced by injected electrons partially compensates for the SPP propagation losses.

 figure: Fig. 3

Fig. 3 SPP modal gain versus pump current for different densities of surface states at the HfO2/InAs interface. Blue, yellow and red regions show three regimes of the active plasmonic waveguide: at low injection currents, the electron concentration in InAs is quite small for population inversion; in the yellow region, the SPP propagation losses are partially compensated by optical gain in InAs; and at high injection currents, the material gain is significantly large for net SPP gain. The black circles highlight the breakdown condition for the HfO2 insulating layer Vi/di = 1 V/nm [40].

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 (a) Contributions of different tunneling processes to the injection current. (b) Material gain profile across the InAs rib at different bias voltages and spatial profile of the SPP electric field averaged over the waveguide width |Eavg(z)|2=w/2w/2|Eavg(y,z)|2dy/w. For both panels, the density of surface states equals 1013 cm−2eV−1.

Download Full Size | PDF

Eventually, at an injection current of about 2.7 kA/cm2 (Fig. 3), ohmic losses and radiation losses of the plasmonic mode are fully compensated. For comparison, it was recently numerically shown [11] that in active plasmonic waveguides based on the Au/p-InAs Schottky contact under the similar conditions a current density of the order of 20 kA/cm2 is needed for full compensation of the SPP propagation losses. Here, we have achieved one order of magnitude reduction in the threshold current density using the tunneling Au/HfO2/InAs contact instead of the Schottky contact and blocking the majority carrier current with a thin insulator layer.

As evident from Fig. 3, the current density required for full loss compensation is robust against the quality of the HfO2/InAs interface. It is equal to 2.6 kA/cm2 in the case of the ideal interface and increases by only 12% as the density of surface states increases to 2 × 1013 cm−2eV−1. This small increase is completely attributed to the tunnel current from the metal to surface states [Fig. 4(a)], while the surface recombination current Jsr,n is more than six orders of magnitude lower than the total current. This is explained by the energy barrier for electrons in the semiconductor due to band bending at the HfO2/InAs interface under high forward bias [41].

Interesting is the possibility to control optical and electrical properties of the active T-shaped hybrid plasmonic waveguide by varying the thickness of the insulator layer. As di increases, the portion of the SPP field in the metal decreases, which results in a decrease of ohmic losses. At the same time, thicker insulating layers block the majority carrier current more efficiently as follows from Equation (1). However, it appears that the tunnel current is very sensitive to the thickness of the HfO2 layer and di = 3 nm is the optimal value. As di is decreased by 0.5 nm (one atomic layer), the majority carrier current drastically increases due to direct tunneling from the metal to surface states and, at a surface-state density of 1013 cm−2eV−1, the SPP propagation losses are fully compensated at a very high current density of 7.5 kA/cm2 (see Appendix E). On the other hand, the 3.5 nm thick HfO2 layer efficiently blocks the majority current, but the barrier width for minority carriers is also increased and higher bias voltages are needed for efficient electron injection. As a result, the electric field in HfO2 rapidly reaches the breakdown value of 1 V/nm [40] as the electron current increases. The minimum achievable SPP modal loss is 240 cm−1 that is much greater than 2Imβpsv = 155 cm−1 evaluated in the case of the lossless (e.g. wide bandgap) semiconductor. For the optimal HfO2 layer thickness of 3 nm, the breakdown condition Vi = 3 V corresponds to the net SPP modal gain of 40 cm−1, which can be increased up to 160 cm−1 in high-quality samples, where the breakdown electric field can be as high as 1.2 V/nm [40].

3.3. Low-current loss compensation in copper waveguide

The effective density of states in the valence band of InAs is 75 times greater than in the conduction band. Therefore, in spite of the relatively large ratio Dcb/Dvb=5 for the Au/HfO2/InAs contact, the electron tunnel current does not exceed that of holes [Fig. 4(a)]. In addition, under high forward bias, the effective barrier height for holes, which can be estimated as the energy difference between the hole quasi-Fermi level in InAs near the insulator-semiconductor contact and the valence band edge of HfO2 at z = di, is substantially reduced due to band bending and the ratio Jt,cb/Jt,vb rapidly decreases as the bias voltage increases [Fig. 4(a)]. To improve the efficiency of the amplification scheme, one should increases the difference between barrier heights χVI and φMI as it can be seen from equation (1). To satisfy this demand, one or more materials in the T-shaped active plasmonic waveguide [Fig. 2(a)] should be replaced.

It would be naive to expect better characteristics in plasmonic structures based on metals different from gold and silver, however, in electrically pumped active plasmonic waveguides, both electrical and optical processes are involved, and the optimal configuration is typically the result of the interplay between them. Copper is characterized by higher absorption at optical frequencies than gold [42, 43] and the modal loss of the plasmonic TM00 mode in the T-shaped waveguide increases from 152 cm−1 to 189 cm−1. On the other hand, the electron work function of copper is equal to 4.7 eV, which is 0.4 eV lower than that of gold. Such a small difference is sufficient to produce an appreciable effect on carrier tunneling through the insulating barrier and increase the efficiency of minority carrier injection in InAs. The electron-to-hole transparency ratio Dcb/Dvb of the Au/HfO2/InAs contact is as large as 20, and the electron tunnel current Jt,cb substantially exceeds the hole tunnel current Jt,vb (Fig. 5).

 figure: Fig. 5

Fig. 5 Gain-current characteristic for the SPP mode of the T-shaped active hybrid plasmonic waveguide based on the Cu/HfO2/InAs MIS structure and relative contributions of different tunnelling processes to the total current. The density of surface states is equal to 5×1012 cm−2eV−1 and the dimensions of the waveguide are the same as in Fig. 2(b).

Download Full Size | PDF

At a surface-state density of 5 × 1012 cm−2eV−1, the propagation losses of the plasmonic mode are fully compensated at a current density of 0.95 kA/cm2, which is significantly smaller than in the case of the gold active plasmonic waveguide despite higher absorption losses of the SPP mode. Passivation of the InAs surface reduces the density of surface states by orders of magnitude [34], and the regime of full loss compensation is reached at a very low current J = 0.82 kA/cm2, which is comparable with the threshold currents in double-heterostructure InAs lasers [44]. In addition, the copper waveguide operates at smaller bias voltages than the gold one. First, this results in lower energy consumption per unit waveguide length, which decreases from 30 mW/mm to 8.2 mW/mm in the regime of full loss compensation. Second, and more importantly, smaller bias voltages create lower electric fields in the HfO2 layer. The breakdown condition (Vi/di = 1 V/nm) corresponds to the net SPP modal gain of 220 cm−1 that is much higher than in the similar gold waveguide. This allows one to use the proposed amplification scheme in nanolasers [45], where not only ohmic, but also high radiation losses have to be compensated.

4. Conclusion

In summary, we have proposed a novel approach for SPP amplification in deep-subwavelength structures under electrical pumping and demonstrated full loss compensation in an active hybrid plasmonic waveguide. The amplification scheme is based on efficient minority carrier injection in metal-insulator-semiconductor contacts at large forward bias, which creates a population inversion in the semiconductor and provides optical gain for the plasmonic mode propagating along the metal interface. At the same time, a thin insulator layer blocks majority carriers giving a possibility to reduce the leakage current to that in double-heterostructure lasers. Comprehensive electronic/photonic simulations show that the SPP propagation losses can be fully compensated in the electrically driven active hybrid plasmonic waveguide based on the Au/HfO2/InAs structure at a current density of only 2.6 kA/cm2. Moreover, by replacing gold with copper, one significantly improves the efficiency of minority carrier injection, and, in spite of higher ohmic losses, the current density in the regime of lossless SPP propagation is reduced down to 0.8 kA/cm2. Such an exceptionally low value demonstrates the potential of electrically pumped active plasmonic waveguides and plasmonic nanolasers for future high-density photonic integrated circuits.

Appendix A. Calculation of the tunnel current

The tunnel injection current across the metal-insulator-semiconductor (MIS) contact from the metal to the k-th band of the semiconductor [k = cb stands for the conduction and k = vb stands for the valence band] can be expressed as an integral over the transverse momenta p and energies E of the incident electrons [20]:

Jt,k=4πe(2π)3Ec+dE[fm(E)fk(E)]0p,maxpdpDk(E,p).

Here, Dk (E, p) is the probability of carrier tunneling to the k-th band of the semiconductor, p⊥,max is the maximum transverse momentum of the carrier in the semiconductor at a given energy E and fm(E), fcb(E), and fvb(E) are the electron distribution functions in the metal and respective bands of the semiconductor. For electron tunneling to the conduction band, p,max=2me(EEc), where me = 0.026m0 is the effective electron mass [51] and Ec is the conduction band edge in InAs. In the case of tunneling to the heavy (light) hole band, p,max=2mhh(lh)(Ev|z=diE), where mlh = 0.024m0 and mhh = 0.36m0 are the light and heavy hole effective masses, respectively, and Ev|z=di is the valence band edge in InAs at the interface. The contributions from light and heavy holes are summed over when calculating the net hole current.

The tunneling probability can be directly obtained in the WKB-approximation

Dk(E,p)=exp[22mihz1z2Ep22miUk(z)dz].

Here, z1,2 are the classical turning points, mi is the effective mass in the insulator [22], and Ucb(z) and Uvb(z) are the conduction and valence band edges in the insulator. To obtain closed-form expressions for the tunneling probability, we introduce the following notations [see Fig. 6(a)]: φMI is the barrier height at the metal-insulator interface; Ec and Ec z=d are the conduction band edges in the bulk and at the interface, respectively; Fi = Vi/di is the the electric field in the insulator, and Fs is the electric field in the semiconductor at the semiconductor-insulator interface. The height and width of the barrier and, consequently, its transparency strongly depend on the electron energy E which we reckon from the Fermi level in the metal.

  1. Electrons with energies Ep22mi>φMIeVi pass through the triangular barrier in the insulator [red arrow in Fig. 6(a)]. The corresponding transparency, Dcb,I, is given by
    lnDcb,I(E,p)=42mi3eFi[φMI+p22miE]3/2.
  2. Electrons with energies Ec|z=di<Ep2/2mi<φMIeVi tunnel through the trapezoid barrier in the insulator [green arrow in Fig. 6(a)]. The transparency Dcb,II of the trapezoid barrier is
    lnDcb,II(E,p)=42mi3eFi{[φMI+p22miE]3/2[φMIeVi+p22miE]3/2}.
  3. Finally, electrons with energies Ec < Ep2 ⊥/2me < Ec z=d tunnel through the trapezoid barrier in the insulator followed by a short section of the bandgap in the semiconductor [blue arrow in Fig. 6(a)]. It should be noted here that the electric field in heavily doped semiconductors is abruptly screened near the surface. At a doping density of NA = 1018 cm−3, the screening length in InAs is about ls=2 nm [see Fig. 6(a)]. Hence, the contribution of these electrons to the tunnel current is non-negligible. The transparency of such a compound barrier is given by
    lnDcb,III(E,p)=lnDII(E,p)432meeFs[Ec|z=diE+p22me]3/2.

 figure: Fig. 6

Fig. 6 (a) Illustration of three different components of the electron tunnel current: tunneling through the triangular barrier (I), tunneling through the trapezoid barrier (II), and tunneling through the trapezoid barrier followed by a short (~ 2 nm) section of the bandgap in the semiconductor (III). (b) Schematic illustration of carrier transport in the MIS structure under forward bias. Black arrows indicate the tunnel currents between the metal and the conduction band (Jt,cb) and between the metal and the valence band (Jt,vb). Color arrows show the processes related to the population and depopulation of surface states: capture and thermal emission of electrons (blue arrows), capture and thermal emission of holes (red arrows), direct tunneling from the metal to surface states (green arrows).

Download Full Size | PDF

The main contribution to the tunnel current emerges from the low transverse momenta p. Thus, one can expand the right-hand sides of Eqs. (68) over p and integrate over the transverse momenta in Eq. (4). The tunnel current density is then written as

Jt,cb=4πemi(2π)3Ec+[fm(E)fcb(E)]S(E)dE,
where the functional form of S(E) depends on the energy of the incident electron. For E > φMIeVi (triangular barrier),
S(E)=Et,IDcb,I(E,0)[1exp(memiEEcEt,I)],
where the characteristic tunneling energy Et,I is
Et,I=eFi2[2mi(φMIE)]1/2.

For Ec|z=di<E<φMIeVi (trapezoid barrier),

S(E)=Et,IIDcb,II(E,0)[1exp(memiEEcEt,II)],
where
Et,II=eFi22mi([φMIE]1/2[φMIEeVi]1/2).

Finally, for low-energy electrons (Ec<E<Ec|z=di),

S(E)=Et,IIIDcb,III(E,0)[1exp(memiEEcEt,III)],
Et,III={22mi(|φMIE|1/2|φMIEeVi|1/2)eFi+mime2[2me(EEc)]1/2eFs}1.

The probability Dvb(E, p) of electron tunneling from the metal to the valence band of the semiconductor or, equivalently, the hole tunneling probability is obtained in a similar way. Namely, holes near the valence band edge (Ev|z=diEp2/2mi<χVIeVi) tunnel through the trapezoid barrier, whose transparency is given by

lnDvb,I(E,p)=42mi3eFi{[χVI+p22miEv|z=di+E]3/2[χVIeVi+p22miEv|z=di+E]}3/2,
while holes with sufficiently high energies (Ev|z=diEp2/2mi>χVIeVi) tunnel through the triangular barrier with transparency
lnDvb,II(E,p)=42mi3eFi[χVI+p22miEv|z=di+E]3/2.

Appendix B. Distribution of electric potential

As in most semiconductor tunnel structures [20], the injection current density in the MIS structure [Eq. (9)] depends exponentially on the voltage drop Vi=Eidi across the thin insulator layer [see Fig. 6(a)]. Hence, the accurate determination of Vi becomes critical. It can be found from the boundary condition for the electric displacement at the semiconductor-insulator interface:

εiVi/diεsEs=Qss/ε0.

Here, εi = 25, [46] and εs = 15, [47] are the static dielectric constants of the insulator and semiconductor, respectively, and Es is the electric field in the semiconductor at the insulator-semiconductor interface. This field is found from Poisson’s equation for the electric potential φ in the semiconductor

d2φdx2=e(NAp)εsε0,
where p is the density of holes. The contribution of electrons to the charge density can be neglected, since the population inversion in p-type InAs is achieved at an electron density of the order of 1016 cm−3, which is far below the hole density and the the acceptor concentration at the chosen doping level.

By multiplying Eq. (19) by /dx and integrating over dx, one easily obtains the expression for the electric field in the semiconductor at the semiconductor-insulator interface

Es={2εsε0ζpζp+eVs[(εF)NA]dεF}1/2,
where ζp = EvFp is the hole Fermi energy in the bulk of the semiconductor and Vs = VVi + (χCIEGζp)/e is the voltage drop across the semiconductor [see Fig. 6(b)]. Here, we have assumed that the distribution of holes is equilibrium, which is justified by a large barrier height for holes at the semiconductor-insulator interface.

Appendix C. Surface states at the semiconductor-insulator interface: charge density, surface recombination and tunnel current through the surface states

The presence of the surface states at the semiconductor-insulator interface has a significant impact on the characteristics of the tunnel MIS contact [21]. First, surface states accumulate electric charge and affect the band bending and voltage drop Vi across the insulator layer. Second, the direct tunnel current from the metal to the surface states contributes to the net current across the junction [27]. Finally, surface states act as recombination centers [26] and reduce the electron injection current:

J0,n=Jt,cbJsr,n,
where J0,n is the electron current density at the insulator-semiconductor interface, Jt,cb is the electron tunnel current density across the MIS contact, and Jsr,n is the surface recombination current density.

Following [48, 49], we assume that surface states ’communicate’ with the electron states in metal and semiconductor bands, but not with each other. Thus, the occupancy of each state fss(E) with energy E is independent from the other states. fss(E) is found from the detailed balance principle between three population/depopulation processes schematically shown in Fig. 6(b):

  1. electron capture from the conduction band of the semiconductor to the surface states and their thermal emission to the band;
  2. hole capture and thermal emission;
  3. direct tunnelling from the metal to surface states.

In a steady state, fss is found to be given by

fss(E)=vn+vpe(FpE)/kBT+vt(E)fm(E)vn[e(EFn)/kBT+1]+vp[e(FpE)/kBT+1]+vt(E),
where we have introduced the rates of electron capture vn=n|z=diσnvn/4, hole capture vp=n|z=diσpvp/4, and electron tunnel escape from the surface state to the metal νt (E). In the above equations, σn and σp are the electron and hole capture cross-sections, vn and vp are the carrier thermal velocities.

The capture-cross section depends on whether the surface trap is donor type or acceptor type. Experimental studies have revealed the donor-type behavior of surface states in InAs [31]. The hole capture cross-sections by a neutral trap is approximately equal to 10−15 cm−2 [26], while the electron capture cross-section σn by a positively charged center is calculated using the theory of carrier capture mediated by cascade phonon emission [50]:

σn=4π3rT3lε1.

Here, rT = e2/4πε0εskBT is the Bohr radius for an electron with an energy equal to the thermal energy, and lε is the energy relaxation length. The latter is smaller than the momentum relaxation length lp by a factor of kT/2mes2 due to the quasi-elastic character of electron-phonon relaxation (here s is the sound velocity in the semiconductor). The momentum relaxation length lp = vnτp is estimated using the electron mobility in intrinsic InAs µn,i = p/me. Substituting the values appropriate to InAs (effective mass me = 0.026m0, average sound velocity s = 3 × 105 m/s, mobility µn,i, = 105 cm2 V−1 s−1 [51]), we obtain lp = 500 nm, kT/2mes2 ≈ 2350, rT = 14 nm, and σn = 10−16 cm−2. The above derivation shows that such a small capture cross section is due to the large energy relaxation length in InAs.

Following [28], we can write the tunnel escape rate as

vt(E)=τ01D(E,0),
where D(E,0) is the tunneling probability evaluated at zero transverse momentum and τ0 is the characteristic escape time:
τ0=mi2di2m0κ2+kF2κ2kF[1κdikF(dikF2κ)]1.

In the above expression, κ=2miχCI/ is the electron wave function decay constant in the insulator, kF is the Fermi wave vector in the metal, m0 is the electron effective mass in metal assumed to be equal to the free electron mass, and I(x) stands for the Dawson integral:

(x)=ex20xet2dt.

In fact, the characteristic escape time weakly depends on the barrier width di and the Fermi energy of the metal. For the 3-nm-thick HfO2 layer and Au contact (Fermi energy 5.5 eV), τ0 = 0.25 fs, while for Cu contact (Fermi energy 7.0 eV) τ0 = 0.29 fs.

Assuming the density of surface states ρss to be uniform over the band gap, we readily express the surface charge density

Qss=eρssEvEc[1fss(E)]dE,
the surface recombination current
Jsr,n=14evnn|z=diσnρssEvEc{[1fss(E)]fss(E)e(EFn)/kBT}dE,
and the direct tunnel current through these states
Jt,ss=eρssEvEcvt(E)[fm(E)fss(E)]dE.

We finally note that the electron density at the semiconductor-insulator interface n|z=di entering the expressions for fss (E) and Jsr,n is considerably smaller than the electron density nbulk in the bulk of the semiconductor (i.e. at z > ls + di, where ls is the screening length) at large forward bias. The reason for this is that the energy barrier prevents the electrons from reaching the surface [41] and n|z=dinbulkeeVs/*kBT, where Vs is the voltage drop across the semiconductor.

Appendix D. Calculation of the material gain

Optical gain in the semiconductor is obtained by integrating the transition probabilities between electron states in the conduction and valence bands over the states’ energies:

g(Fn,Fp)=πe2nInAsm02ε0|Mb|26ω+ρcb(E)ρvb(Eω)|Menv(E,Eω)|2[fcb(E,En)fvb(Eω,Fp)]dE.

Here, ρcb(E) and ρvb(E) are the densities of states in the conduction and valence bands of the semiconductor, respectively, fcb(E, Fn) and fvb(E, Fp) are the Fermi-Dirac distribution functions for the conduction and valence bands, Mb is the average matrix element connecting Bloch states near the band edge [52], and Menv is the envelope matrix element calculated using Stern’s model [39]. In heavily doped semiconductors, ρcb, ρvb, Menv and consequently the material gain, strongly depend on the impurity concentration, while at low doping levels Menv is strongly peaked at the energy corresponding to the exact momentum conservation for the electron-photon system [17, 39], and equation (30) is reduced to the that in the band-to-band transition model with the k-selection rule.

The calculated gain spectra at an acceptor density of 1018 cm−3 are shown in Fig. 7. The maximum of the gain spectrum shifts slightly towards shorter wavelengths as the electron density increases. At an electron concentration of 2 × 1016 cm−3 near the MIS contact, which corresponds to the full loss compensation regime, the maximum is achieved at λ = 3.22 µm.

 figure: Fig. 7

Fig. 7 Gain spectra of InAs calculated using Stern’s model [see Eq. (30)] at different electron densities, the acceptor density is equal to 1018 cm−3 for all curves.

Download Full Size | PDF

Appendix E. Gain characteristics at different thicknesses of the insulator layer

We have varied the thickness of the insulator layer in order to achieve the regime of full loss compensation at the lowest possible injection current. The simulated gain-current and current-voltage characteristics are presented in Fig. 8. One readily notes that the insulator thickness of di = 3 nm is optimal. At higher thicknesses, full loss compensation cannot be achieved before the breakdown of the HfO2 layer at Vi/di ≈ 1 V [40]. At lower oxide thicknesses, the current density in the regime of full loss compensation is greater than 5 kA/cm2 due to the large tunnel current Jt,ss through the surface states [Fig. 8(c)].

 figure: Fig. 8

Fig. 8 (a) SPP modal gain versus injection current density for three different thicknesses of the HfO2 layer of the Au/HfO2/InAs active hybrid plasmonic waveguide. Dashed curves correspond to the voltage range above the breakdown voltage of HfO2. (b) Current-voltage characteristics (color coding is the same as in panel (a)). (c, d) Contribution of different tunneling processes to the injection current for the 2.5-nm-thick (c) and 3.5-nm-thick (d) HfO2 layers. For all panels, the density of surface states is equal to 1013 cm−2eV−1.

Download Full Size | PDF

Acknowledgments

The work was supported by the Russian Science Foundation, grant no. 14-19-01788.

References and links

1. A. Kriesch, S. P. Burgos, D. Ploss, H. Pfeifer, H. A. Atwater, and U. Peschel, “Functional plasmonic nanocircuits with low insertion and propagation losses,” Nano Lett. 13, 4539–4545 (2013). [CrossRef]   [PubMed]  

2. J. A. Conway, S. Sahni, and T. Szkopek, “Plasmonic interconnects versus conventional interconnects: a comparison of latency, crosstalk and energy costs,” Opt. Express 15, 4474–4484 (2007). [CrossRef]   [PubMed]  

3. V. J. Sorger, Z. Ye, R. F. Oulton, Y. Wang, G. Bartal, X. Yin, and X. Zhang, “Experimental demonstration of low-loss optical waveguiding at deep sub-wavelength scales,” Nat. Commun. 2, 331 (2011). [CrossRef]  

4. S. Kena-Cohen, P. N. Stavrinou, D. D. Bradley, and S. A. Maier, “Confined surface plasmon-polariton amplifiers,” Nano Lett. 13, 1323–1329 (2013). [CrossRef]   [PubMed]  

5. R. F. Oulton, V. J. Sorger, T. Zentgraf, R.-M. Ma, C. Gladden, L. Dai, G. Bartal, and X. Zhang, “Plasmon lasers at deep subwavelength scale,” Nature 461, 629–632 (2009). [CrossRef]   [PubMed]  

6. R. Flynn, C. Kim, I. Vurgaftman, M. Kim, J. Meyer, A. Mäkinen, K. Bussmann, L. Cheng, F.-S. Choa, and J. Long, “A room-temperature semiconductor spaser operating near 1.5 µ m,” Opt. Express 19, 8954–8961 (2011). [CrossRef]   [PubMed]  

7. A. W. Fang, R. Jones, H. Park, O. Cohen, O. Raday, M. J. Paniccia, and J. E. Bowers, “Integrated AlGaInAs-silicon evanescent race track laser and photodetector,” Opt. Express 15, 2315–2322 (2007). [CrossRef]   [PubMed]  

8. D. Y. Fedyanin and A. V. Arsenin, “Semiconductor surface plasmon amplifier based on a Schottky barrier diode,” in AIP Conference Proceedings1291 (2010), pp. 112–114.

9. D. Y. Fedyanin and A. V. Arsenin, “Surface plasmon polariton amplification in metal-semiconductor structures,” Opt. Express 19, 12524–12531 (2011). [CrossRef]   [PubMed]  

10. D. Costantini, A. Bousseksou, M. Fevrier, B. Dagens, and R. Colombelli, “Loss and gain measurements of tensile strained quantum well diode lasers for plasmonic devices at telecom wavelengths,” IEEE J. Quantum Electron. 48, 73–78 (2012). [CrossRef]  

11. D. Y. Fedyanin, A. V. Krasavin, A. V. Arsenin, and A. V. Zayats, “Surface plasmon polariton amplification upon electrical injection in highly integrated plasmonic circuits,” Nano Lett. 12, 2459–2463 (2012). [CrossRef]   [PubMed]  

12. D. Costantini, “Compact generation and amplification of surface plasmon polaritons at telecom wavelengths,” Ph.D. thesis, University Paris Sud, ParisXI (2013).

13. C. Wang, H. J. Qu, W. X. Chen, G. Z. Ran, H. Y. Yu, B. Niu, J. Q. Pan, and W. Wang, “Polarization of the edge emission from Ag/InGaAsP Schottky plasmonic diode,” Appl. Phys. Lett. 102, 061112 (2013). [CrossRef]  

14. D. Y. Fedyanin, “Toward an electrically pumped spaser,” Opt. Lett. 37, 404–406 (2012). [CrossRef]   [PubMed]  

15. M. T. Hill, M. Marell, E. S. Leong, B. Smalbrugge, Y. Zhu, M. Sun, P. J. van Veldhoven, E. J. Geluk, F. Karouta, Y.-S. Oei, et al., “Lasing in metal-insulator-metal sub-wavelength plasmonic waveguides,” Opt. Express 17, 11107–11112 (2009). [CrossRef]   [PubMed]  

16. G. Wade, C. Wheeler, and R. Hunsperger, “Inherent properties of a tunnel-injection laser,” Proc. IEEE 53, 98–99 (1965). [CrossRef]  

17. H. Panish Jr and M. Casey, Heterostructure Lasers: Part A: Fundamental Principles (Academic Press, 1978).

18. T. Wijesinghe, M. Premaratne, and G. P. Agrawal, “Electrically pumped hybrid plasmonic waveguide,” Opt. Express 22, 2681–2694 (2014). [CrossRef]   [PubMed]  The results obtained in this reference assume free passage of carriers through the insulator layer with the injection velocity equal to the thermal velocity. This contradicts to the established theories of carrier transport in MIS structures [20,21]. Due to the high potential barriers at the metal-insulator and semiconductor-insulator interfaces, the thermionic injection cannot be actually efficient, while the dominating transport mechanism is due to tunneling.

19. M. Green and J. Shewchun, “Minority carrier effects upon the small signal and steady-state properties of Schottky diodes,” Solid-State Electron. 16, 1141–1150 (1973). [CrossRef]  

20. R. Stratton, “Volt-current characteristics for tunneling through insulating films,” J. Phys. Chem. Solids 23, 1177–1190 (1962). [CrossRef]  

21. H. Card and E. Rhoderick, “Studies of tunnel MOS diodes I. Interface effects in silicon Schottky diodes,” J. Phys. D: Appl. Phys. 4, 1589 (1971). [CrossRef]  

22. Y. C. Chang, M. L. Huang, K. Y. Lee, Y. J. Lee, T. D. Lin, M. Hong, J. Kwo, T. S. Lay, C. C. Liao, and K. Y. Cheng, “Atomic-layer-deposited HfO2 on In0.53Ga0.47As: Passivation and energy-band parameters,” Appl. Phys. Lett. 92, 072901 (2008). [CrossRef]  

23. J. Robertson and B. Falabretti, “Band offsets of high k gate oxides on III-V semiconductors,” J. Appl. Phys. 100, 014111 (2006). [CrossRef]  

24. D. Wheeler, “High-k InAs metal-oxide-semiconductor capacitors formed by atomic layer deposition,” Ph.D. thesis, University of Notre Dame, Notre Dame, Indiana (2009).

25. S. M. Sze and K. K. Ng, Physics of Semiconductor Devices (John Wiley & Sons, 2006).

26. V. Abakumov, V. I. Perel, and I. Yassievich, Nonradiative Recombination in Semiconductors (Elsevier, 1991).

27. H. Card, “On the direct currents through interface states in metal-semiconductor contacts,” Solid-State Electron. 18, 881–883 (1975). [CrossRef]  

28. I. Lundström and C. Svensson, “Tunneling to traps in insulators,” J. Appl. Phys. 43, 5045–5047 (1972). [CrossRef]  

29. M. Noguchi, K. Hirakawa, and T. Ikoma, “Intrinsic electron accumulation layers on reconstructed clean InAs (100) surfaces,” Phys. Rev. Lett. 66, 2243 (1991). [CrossRef]   [PubMed]  

30. M. G. Betti, G. Bertoni, V. Corradini, V. De Renzi, and C. Mariani, “Metal-induced gap states at InAs (110) surface,” Surf. Sci. 454, 539–542 (2000). [CrossRef]  

31. W. Mönch, Semiconductor Surfaces and Interfaces (Springer Science & Business Media, 2001). [CrossRef]  

32. D. Wheeler, L.-E. Wernersson, L. Fröberg, C. Thelander, A. Mikkelsen, K.-J. Weststrate, A. Sonnet, E. Vogel, and A. Seabaugh, “Deposition of HfO2 on InAs by atomic-layer deposition,” Microelectron. Eng. 86, 1561–1563 (2009). [CrossRef]  

33. C. Wang, G. Doornbos, G. Astromskas, G. Vellianitis, R. Oxland, M. Holland, M. Huang, C. Lin, C. Hsieh, Y. Chang, et al., “High-k dielectrics on (100) and (110) n-InAs: physical and electrical characterizations,” AIP Adv. 4, 047108 (2014). [CrossRef]  

34. C. Wang, S. Wang, G. Doornbos, G. Astromskas, K. Bhuwalka, R. Contreras-Guerrero, M. Edirisooriya, J. Rojas-Ramirez, G. Vellianitis, R. Oxland, M.C. Holland, C. H. Hsieh, P. Ramvall, E. Lind, W.C. Hsu, L.-E. Wernersson, R. Droopad, M. Passlack, and C.H. Diaz, “InAs hole inversion and bandgap interface state density of 2 × 1011 cm2eV−1 at HfO2/InAs interfaces,” Appl. Phys. Lett. 103, 143510 (2013). [CrossRef]  

35. I. Malitson, “Interspecimen comparison of the refractive index of fused silica,” J. Opt. Soc. Am. 55, 1205–1208 (1965). [CrossRef]  

36. T. Bright, J. Watjen, Z. Zhang, C. Muratore, and A. Voevodin, “Optical properties of HfO2 thin films deposited by magnetron sputtering: From the visible to the far-infrared,” Thin Solid Films 520, 6793–6802 (2012). [CrossRef]  

37. Y. Tsou, A. Ichii, and E. M. Garmire, “Improving InAs double heterostructure lasers with better confinement,” IEEE J. Quantum Electron. 28, 1261–1268 (1992). [CrossRef]  

38. J. T. Robinson, K. Preston, O. Painter, and M. Lipson, “First principle derivation of gain in high-index-contrast waveguides,” Opt. Express 16, 16659–16669 (2008). [CrossRef]   [PubMed]  

39. H. Casey Jr and F. Stern, “Concentration-dependent absorption and spontaneous emission of heavily doped GaAs,” J. Appl. Phys. 47, 631–643 (1976). [CrossRef]  

40. C.-H. Chang and J.-G. Hwu, “Characteristics and reliability of hafnium oxide dielectric stacks with room temperature grown interfacial anodic oxide,” IEEE Trans. Device Mater. Rel. 9, 215–221 (2009). [CrossRef]  

41. J. Rimmer, J. Langer, M. Missous, J. Evans, I. Poole, A. Peaker, and K. Singer, “Minority-carrier confinement by doping barriers,” Mater. Sci. Eng. B 9, 375–378 (1991). [CrossRef]  

42. M. Ordal, L. Long, R. Bell, S. Bell, R. Bell, R. Alexander, and C. Ward, “Optical properties of the metals Al, Co, Cu, Au, Fe, Pb, Ni, Pd, Pt, Ag, Ti, and W in the infrared and far infrared,” Appl. Opt. 22, 1099–1119 (1983). [CrossRef]   [PubMed]  

43. A. Lenham and D. Treherne, “Applicability of the anomalous skin-effect theory to the optical constants of Cu, Ag, and Au in the infrared,” J. Opt. Soc. Am. 56, 683–685 (1966). [CrossRef]  

44. M. Aydaraliev, N. Zotova, S. Karandashov, B. Matveev, G. Talalakin, et al., “Low-threshold long-wave lasers (λ = 3.0–3.6µ m) based on III–V alloys,” Semicond. Sci. Technol. 8, 1575 (1993). [CrossRef]  

45. M. I. Stockman, “Nanoplasmonics: past, present, and glimpse into future,” Opt. Express 19, 22029–22106 (2011). [CrossRef]   [PubMed]  

46. J. Robertson, “High k dielectrics for future CMOS devices,” ECS Trans. 19, 579–591 (2009). [CrossRef]  

47. O. Madelung, Semiconductors: Group IV elements and III–V compounds (Springer Verlag, 1991), Vol. 1. [CrossRef]  

48. P. Landsberg and C. Klimpke, “Surface recombination effects in an improved theory of a p-type MIS solar cell,” Solid-State Electron. 23, 1139–1145 (1980). [CrossRef]  

49. K. Ng and H. Card, “A comparison of majority- and minority-carrier silicon MIS solar cells,” IEEE Trans. Electron. Dev. 27, 716–724 (1980). [CrossRef]  

50. V. Abakumov and I. Iassievich, “Cross section for the recombination of an electron on a positively charged center in a semiconductor,” Sov. Phys. JETP 71, 657–664 (1976).

51. S. Adachi, Properties of Semiconductor Alloys: Group-IV, III–V and II–VI Semiconductors (John Wiley & Sons, 2009), vol. 28. [CrossRef]  

52. L. A. Coldren, S.W. Corzine, and M. L. Mashanovitch, Diode Lasers and Photonic Integrated Circuits (John Wiley & Sons, 2012). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (8)

Fig. 1
Fig. 1 SPP amplification schemes based on a Schottky barrier diode [11] and tunnel MIS contact. Top panels: Energy band diagrams for the Schottky contact between gold and indium arsenide in equilibrium (a) and at high forward bias (b). The Schottky barrier height for electrons φBn at the Au/InAs contact is negative owing to the large density of surface states and, under high forward bias, electrons (minority carriers in p-type InAs) are freely injected in the bulk of the semiconductor. However, majority carriers (holes) also pass across the Au/InAs contact without any resistance, which results in a high leakage current. Bottom panels: Energy band diagrams for the tunnel Au/HfO2/InAs MIS structure in equilibrium (c) and at high forward bias (d). If the barrier height for holes χVI is substantially greater than that for electrons (φMI), the insulator layer can efficiently block majority carriers (holes), but be semi-transparent for minority carriers (electrons) escaping from the metal.
Fig. 2
Fig. 2 Schematic of a T-shaped hybrid plasmonic waveguide based on the Au/HfO2/InAs MIS structure, w is the waveguide width, di is the thickness of the low refractive index insulator layer between the metal and semiconductor, and H is the waveguide height. (b) Distribution of the normalized energy density per unit length of the waveguide for the fundamental TM00 mode at λ = 3.22 µm, H = 2.5 µm, w = 400 nm and di = 3 nm. The dielectric functions of the materials are as follows: ε Si O 2 = 2.00 [35], ε Hf O 2 = 3 . 84 [36], εInAs = 12.38 [37] and εAu = −545+38i [11].
Fig. 3
Fig. 3 SPP modal gain versus pump current for different densities of surface states at the HfO2/InAs interface. Blue, yellow and red regions show three regimes of the active plasmonic waveguide: at low injection currents, the electron concentration in InAs is quite small for population inversion; in the yellow region, the SPP propagation losses are partially compensated by optical gain in InAs; and at high injection currents, the material gain is significantly large for net SPP gain. The black circles highlight the breakdown condition for the HfO2 insulating layer Vi/di = 1 V/nm [40].
Fig. 4
Fig. 4 (a) Contributions of different tunneling processes to the injection current. (b) Material gain profile across the InAs rib at different bias voltages and spatial profile of the SPP electric field averaged over the waveguide width | E avg ( z ) | 2 = w / 2 w / 2 | E avg ( y , z ) | 2 d y / w. For both panels, the density of surface states equals 1013 cm−2eV−1.
Fig. 5
Fig. 5 Gain-current characteristic for the SPP mode of the T-shaped active hybrid plasmonic waveguide based on the Cu/HfO2/InAs MIS structure and relative contributions of different tunnelling processes to the total current. The density of surface states is equal to 5×1012 cm−2eV−1 and the dimensions of the waveguide are the same as in Fig. 2(b).
Fig. 6
Fig. 6 (a) Illustration of three different components of the electron tunnel current: tunneling through the triangular barrier (I), tunneling through the trapezoid barrier (II), and tunneling through the trapezoid barrier followed by a short (~ 2 nm) section of the bandgap in the semiconductor (III). (b) Schematic illustration of carrier transport in the MIS structure under forward bias. Black arrows indicate the tunnel currents between the metal and the conduction band (Jt,cb) and between the metal and the valence band (Jt,vb). Color arrows show the processes related to the population and depopulation of surface states: capture and thermal emission of electrons (blue arrows), capture and thermal emission of holes (red arrows), direct tunneling from the metal to surface states (green arrows).
Fig. 7
Fig. 7 Gain spectra of InAs calculated using Stern’s model [see Eq. (30)] at different electron densities, the acceptor density is equal to 1018 cm−3 for all curves.
Fig. 8
Fig. 8 (a) SPP modal gain versus injection current density for three different thicknesses of the HfO2 layer of the Au/HfO2/InAs active hybrid plasmonic waveguide. Dashed curves correspond to the voltage range above the breakdown voltage of HfO2. (b) Current-voltage characteristics (color coding is the same as in panel (a)). (c, d) Contribution of different tunneling processes to the injection current for the 2.5-nm-thick (c) and 3.5-nm-thick (d) HfO2 layers. For all panels, the density of surface states is equal to 1013 cm−2eV−1.

Equations (30)

Equations on this page are rendered with MathJax. Learn more.

D cb D vd = exp [ 2 3 / 2 m i 1 / 2 d i h ( χ VI 1 / 2 φ MI 1 / 2 ) ] ,
n ( z ) J 0 , n e τ R D n exp ( z L D ) + n 0 ,
G = c ε 0 n In As w / 2 + w / 2 d y d i H d z g ( y , z ) | E ( y , z ) | 2 2 + d y + d z P z ( y , z ) 2 Im β psv .
J t , k = 4 π e ( 2 π ) 3 E c + d E [ f m ( E ) f k ( E ) ] 0 p , max p d p D k ( E , p ) .
D k ( E , p ) = exp [ 2 2 m i h z 1 z 2 E p 2 2 m i U k ( z ) d z ] .
ln D cb , I ( E , p ) = 4 2 m i 3 e F i [ φ MI + p 2 2 m i E ] 3 / 2 .
ln D cb , II ( E , p ) = 4 2 m i 3 e F i { [ φ MI + p 2 2 m i E ] 3 / 2 [ φ MI e V i + p 2 2 m i E ] 3 / 2 } .
ln D cb , III ( E , p ) = ln D II ( E , p ) 4 3 2 m e e F s [ E c | z = d i E + p 2 2 m e ] 3 / 2 .
J t , cb = 4 π e m i ( 2 π ) 3 E c + [ f m ( E ) f cb ( E ) ] S ( E ) d E ,
S ( E ) = E t , I D cb , I ( E , 0 ) [ 1 exp ( m e m i E E c E t , I ) ] ,
E t , I = e F i 2 [ 2 m i ( φ MI E ) ] 1 / 2 .
S ( E ) = E t , II D cb , II ( E , 0 ) [ 1 exp ( m e m i E E c E t , II ) ] ,
E t , II = e F i 2 2 m i ( [ φ MI E ] 1 / 2 [ φ MI E e V i ] 1 / 2 ) .
S ( E ) = E t , III D cb , III ( E , 0 ) [ 1 exp ( m e m i E E c E t , III ) ] ,
E t , III = { 2 2 m i ( | φ MI E | 1 / 2 | φ MI E e V i | 1 / 2 ) e F i + m i m e 2 [ 2 m e ( E E c ) ] 1 / 2 e F s } 1 .
ln D vb , I ( E , p ) = 4 2 m i 3 e F i { [ χ VI + p 2 2 m i E v | z = d i + E ] 3 / 2 [ χ VI e V i + p 2 2 m i E v | z = d i + E ] } 3 / 2 ,
ln D vb , II ( E , p ) = 4 2 m i 3 e F i [ χ VI + p 2 2 m i E v | z = d i + E ] 3 / 2 .
ε i V i / d i ε s E s = Q ss / ε 0 .
d 2 φ d x 2 = e ( N A p ) ε s ε 0 ,
E s = { 2 ε s ε 0 ζ p ζ p + e V s [ ( ε F ) N A ] d ε F } 1 / 2 ,
J 0 , n = J t , cb J sr , n ,
f ss ( E ) = v n + v p e ( F p E ) / k B T + v t ( E ) f m ( E ) v n [ e ( E F n ) / k B T + 1 ] + v p [ e ( F p E ) / k B T + 1 ] + v t ( E ) ,
σ n = 4 π 3 r T 3 l ε 1 .
v t ( E ) = τ 0 1 D ( E , 0 ) ,
τ 0 = m i 2 d i 2 m 0 κ 2 + k F 2 κ 2 k F [ 1 κ d i k F ( d i k F 2 κ ) ] 1 .
( x ) = e x 2 0 x e t 2 d t .
Q ss = e ρ ss E v E c [ 1 f ss ( E ) ] d E ,
J sr , n = 1 4 e v n n | z = d i σ n ρ ss E v E c { [ 1 f ss ( E ) ] f ss ( E ) e ( E F n ) / k B T } d E ,
J t , ss = e ρ ss E v E c v t ( E ) [ f m ( E ) f ss ( E ) ] d E .
g ( F n , F p ) = π e 2 n In As m 0 2 ε 0 | M b | 2 6 ω + ρ cb ( E ) ρ vb ( E ω ) | M env ( E , E ω ) | 2 [ f cb ( E , E n ) f vb ( E ω , F p ) ] d E .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.