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

Design of an ultra-compact electro-absorption modulator comprised of a deposited TiN/HfO2/ITO/Cu stack for CMOS backend integration

Open Access Open Access

Abstract

An ultra-compact electro-absorption (EA) modulator operating around 1.55-μm telecom wavelengths is proposed and theoretically investigated. The modulator is comprised of a stack of TiN/HfO2/ITO/Cu conformally deposited on a single-mode stripe waveguide to form a hybrid plasmonic waveguide (HPW). Since the thin ITO layer can behave as a semiconductor, the stack itself forms a MOS capacitor. A voltage is applied between the Cu and TiN layers to change the electron concentration of ITO (NITO), which in turn changes its permittivity as well as the propagation loss of HPW. For a HPW comprised of a Cu/3-nm-ITO/5-nm-HfO2/5-nm-TiN stack on a 400-nm × 340-nm-Si stripe waveguide, the propagation loss for the 1.55-μm TE (TM) mode increases from 1.6 (1.4) to 23.2 (23.9) dB/μm when the average NITO in the 3-nm ITO layer increases from 2 × 1020 to 7 × 1020 cm−3, which is achieved by varying the voltage from −2 to 4 V if the initial NITO is 3.5 × 1020 cm−3. As a result, a 1-μm-long EA modulator inserted in the 400-nm × 340-nm-Si stripe waveguide exhibits insertion loss of 2.9 (3.2) dB and modulation depth of 19.9 (15.2) dB for the TE (TM) mode. The modulation speed is ~11 GHz, limited by the RC delay, and the energy consumption is ~0.4 pJ/bit. The stack can also be deposited on a low-index-contrast waveguide such as Si3N4. For example, a 4-μm-long EA modulator inserted in an 800-nm × 600-nm-Si3N4 stripe waveguide exhibits insertion loss of 6.3 (3.5) dB and modulation depth of 16.5 (15.8) dB for the TE (TM) mode. The influences of the ITO, TiN, HfO2 layers and the beneath dielectric core, as well as the processing tolerance, on the performance of the proposed EA modulator are systematically investigated.

© 2014 Optical Society of America

1. Introduction

Complementary metal-oxide-semiconductor (CMOS)-compatible silicon electro-optical (EO) modulators are basic building blocks for optoelectronic systems [1]. Recently developed Si modulators typically rely on electrically altering the Si refractive index through mechanisms such as carrier injection, accumulation, or depletion in either a PN diode or a metal-oxide-semiconductor (MOS) capacitor, and converting the phase variation into the optical intensity variation through a Mach-Zehnder interferometer (MZI) or a ring resonator [2, 3]. Due to the weak plasma dispersion effect of Si and the diffraction limit of the Si waveguides, the Si MZI modulators suffer from large footprints of ~103–104 μm2. The ring modulators have compact footprints of ~102–103 μm2, but at a price of higher temperature sensitivity and lower optical bandwidth. Ultra-compact modulators with μm2 footprint and even higher performance (i.e., higher speed and lower energy consumption) are highly desired for further high-density Si electronic photonic integrated circuits (Si-EPICs). On the other hand, optical circuits based on deposited materials such as amorphous-Si and Si3N4 are emerging recently for three-dimensional (3D) integration of multiple photonic layers on Si electronic circuits [4, 5] or for flexible photonics [6]. Modulators which can be integrated with these deposited waveguides are needed.

Plasmonis provides an approach to miniaturize optical devices beyond the diffraction limit [7]. Several ultra-compact plasmonic modulators have been reported, but they usually rely on non-CMOS-compatible materials or processes, making it difficult to integrate them into the existing Si-EPICs [8, 9]. A fully CMOS-compatible plasmonic modulator utilizing Si as the active material is demonstrated recently [10, 11]. However, it requires a very large modification of the Si electron concentration to reach a realistic modulation depth. This is because Si is not a good active material for plasmonic modulator, as will be explained in this paper. Moreover, it limits the modulator on the single-crystal Si layer.

Recently, transparent conductor oxides (TCOs) such as indium tin oxide (ITO) are emerging as an attractive active material for new concept optical modulators because of its unique property such as epsilon-near-zero (ENZ) in the near infrared regime and electrically-tunable permittivity [1220]. Based on the Drude model, the real part of the ITO’s permittivity can be tuned between positive and negative by varying its electron concentration (NITO). As a result, its optical property changes dramatically between dielectric-like and metal-like [12, 13]. Since ITO can behave as a highly-doped n-type semiconductor, its NITO can be electrically changed through mechanisms such as accumulation or depletion in a MOS capacitor. In order to form the MOS capacitor and also to enhance the overlap between the optical mode and the thin active ITO layer, the ITO layer together with a gate oxide layer is usually sandwiched between two Si layers to form a slot waveguide [14], or a Si and a metal layer to form a metal-dielectric-Si hybrid plasmonic waveguide (HPW) [15], or two metal layers to form a metal-dielectric-metal (MDM) plasmonic waveguide [1618]. The first two configurations can be easily integrated with conventional Si waveguides with high coupling efficiency because Si is utilized as both the waveguide material and the electrode. However, the Si layer needs to be highly doped and a rib waveguide structure is required for serving as the electrode. Moreover, it limits the modulator on the single-crystal Si layer. The third configuration can be integrated with any kind of waveguide in principle. However, it usually has a large insertion loss because the MDM waveguide has large propagation loss and also large coupling loss to the dielectric waveguide when the sandwiched dielectric layer is very thin. Moreover, these three configurations work only for one polarization mode whose dominate electric field is perpendicular to the ITO/Si or ITO/metal interfaces.

In this paper, a new design of ITO-based plasmonic modulator is proposed, which is comprised of a Cu/ITO/HfO2/TiN stack deposited on a stripe waveguide. The stack itself forms a MOS capacitor where ITO behaves as a semiconductor, HfO2 serves as a high-κ gate dielectric, and Cu and TiN serve as electrodes. The top Cu layer is much thicker and the bottom TiN layer is much thinner than the light penetration depth in a metal (~26 nm), so that the stack together with the beneath stripe waveguide form a hybrid plasmonic waveguide. The stack can be deposited on either high-index-contrast waveguide such as Si or low-index-contrast waveguide such as Si3N4 to form an EA modulator for both transverse electric (TE) and transverse magnetic (TM) modes. Since the thin TiN, HfO2, and ITO layers can be conformally deposited on the waveguide layer-by-layer using the well-developed atomic-layer-deposition (ALD) method at low temperature (e.g., less than 400°C) [21], the proposed modulator is CMOS backend-compatible.

2. Device structure

Figure 1 shows schematic of the proposed modulator integrated with a conventional stripe waveguide. The whole structure is embedded in a thick cladding SiO2 layer. The thin TiN layer is much wider than the Cu/ITO/HfO2 stack. One electrode is connected on Cu and the other is connected on the thin TiN layer. The dielectric core has width of W and height of h. The TiN, HfO2, and ITO layers are first set to have thickness of 5, 5, and 3 nm, respectively. The influence of these thicknesses on the modulator’s performance will be discussed in Section 5. The refractive indices of Si, Si3N4, SiO2, HfO2, TiN [22], and Cu [23] at 1.55-μm wavelength are set to 3.45, 2.0, 1.44, 1.87, 0.9 + 4.18j, and 0.282 + 11.048j, respectively. The permittivity of ITO is defined by the Drude mode as ε=εNITOe2ε0m*1ω2+iωΓ, where ε is the high-frequency permittivity, Γ is the electron damping factor, ω is the angular frequency of light, NITO is the electron concentration, m* is the effective mass, and e and ε0 are the electron charge and the permittivity of free space, respectively. We set ε = 3.9, Γ = 1.8 × 1014 s−1, and m* = 0.35me (where me is the electron mass) for ITO [17]. Figure 2 plots the calculated real part and imaginary part of the ITO’s permittivity as a function of NITO at 1.54, 1.55, and 1.56-μm wavelengths. One sees that the real part of the permittivity crosses zero at a certain NITO, which is defined as the ENZ point. The ENZ point is 6.5 × 1020 cm−3 at 1.55 μm and becomes smaller at longer wavelength. The ITO’s optical property is dielectric-like when its real part of permittivity is positive and is metal-like when its real part of permittivity is negative.

 figure: Fig. 1

Fig. 1 (a) 3D view and (b) cross-sectional view of the proposed EA modulator integrated with a stripe dielectric waveguide.

Download Full Size | PDF

 figure: Fig. 2

Fig. 2 The calculated real part and imaginary part of the ITO’s permittivity as a function of electron concentration in ITO, the real part of the permittivity crosses zero at a certain NITO, which is defined as the ENZ point.

Download Full Size | PDF

3. Electrical simulation

The stack forms a MOS capacitor where a voltage is applied between the TiN and Cu layers. For electrical simulation, the 3D structure shown in Fig. 1 can be simplified to the 2D structure shown in Fig. 3(a). A semiconductor device simulation software MEDICI [24] is used to calculate the 2D electron distribution NITO(x, y) in the thin ITO layer under different voltages. The electron concentration in the as-deposited ITO, which is defined as N0, is typically in the range of ~1020−1021 cm−3, depending on the deposition condition. Here, we assume N0 = 3.5 × 1020 cm−3. The mobility and work function of ITO are set to 50 cm2⋅V−1⋅s−1 and 0.5 eV, respectively, and the resistivity of TiN is set to ~300 μΩ⋅cm. Auger recombination, Shockley-Hall-Read recombination, surface recombination, Fermi-Dirac statics, and the modified local density approximation (MLDA) method are included. We find NITO(x, y) keeps almost unchanged along the x-coordinate (not shown here). Therefore, the 2D distribution NITO(x, y) can be further reduced to the 1D distribution NITO(y), as plotted in Fig. 3(b) under different voltages. One sees that electrons near the ITO/HfO2 interface are depleted under the negative voltages and are accumulated under the positive voltages. The concentration of accumulated electrons maximizes at the ITO/HfO2 interface and decreases to N0 quickly with the distance from the interface increasing. As a first approximation, the distribution can be approximated to a step function, i.e., NITO(y) = NAcL at (tITO – tAcL) < y < tITO and NITO(y) = N0 at 0 < y < (tITO – tAcL), where tITO is the ITO thickness and tAcL is the accumulation layer (AcL) thickness in the case of V > 0. In the case of V < 0, a depletion layer is formed, which is also represented as AcL for simplification. NAcL can be roughly estimated as NAcL=N0+ε0εHfO2etHfO2tAcL(VVFB), where εHfO2 ( = 25) is the dielectric constant of HfO2, tHfO2 is the HfO2 thickness, and VFB is the flat-band voltage. Based on Thomas-Fermi screening theory, tAcL is ~1 nm [18]. But in [12], tAcL is reported to be 5 ± 1 nm. Here we simply assume tAcL = tITO = 3 nm. NAcL is plotted in Fig. 3(c) as a function of V. One sees that the required voltage swing for NITO modification from 2 × 1020 to 7 × 1020 cm−3 is ~-2−4 V in the case of tAcL = 3 nm. The influence of tAcL and tITO on the modulator’s performance is discussed in Section 5.1.

 figure: Fig. 3

Fig. 3 (a) A simplified 2D structure for electrical simulation. (b) 1D election concentration distribution along y coordinate in the 3-nm ITO layer under different voltages, obtained from the MEDICI simulation. The dashed lines represent the average concentrations in the 3-nm ITO layer. (c) Average NITO in the accumulation layer versus voltage in the case of tAcL = 1, 2, or 3 nm.

Download Full Size | PDF

For the transient state simulation, the gate voltage is increased from −2 to 4 V with a ramp time of 10 fs. Figure 4 plots variation of the average NITO in the 3-nm-thick ITO layer with time. The sum of the rise time and the fall time (τs) for 10%-90% NITO variation is ~31 ps. The cut-off frequency estimated as 0.35τs is ~11 GHz. The device in accumulation has capacitance C=Aε0εHfO2tHfO2, where A is the area. For a 1-μm-long modulator with a 400-nm × 340-nm Si core, A is ~1.1 μm2 and C is ~48 fF. The resistance R mainly comes from the thin TiN layer. R = ~300 Ω when the lateral distance between the Cu/ITO/HfO2 stack and the Al contact on the TiN layer is 0.5 μm, as shown in Fig. 3(a). Then, the cutoff frequency defined by fmax=12π×RC is estimated to ~11 GHz, close to that read from Fig. 4. It indicates that the modulation speed is mainly limited by the RC delay. It is reported that the resistivity of TiN depends on the deposition conditions [25]. If the resistivity is improved to ~65 μΩ⋅cm, τs reduces to ~6.5 ps and the cutoff frequency increases to ~54 GHz, as shown by the red curve in Fig. 4. Other methods to reduce R include increasing the TiN layer thickness and/or shortening the lateral distance between the Cu/ITO/HfO2 stack and the Al contact on the TiN layer. However, the insertion loss may be degraded, as will be discussed in Section 5.2. The switching energy per bit can be estimated as E=12ConVon2+12CoffVoff2 [26]. For the above 1-μm-long modulator, Von = −2 V and Voff = 4 V, thus E = ~0.4 pJ/bit.

 figure: Fig. 4

Fig. 4 The transient response of the electron concentration in the 3-nm-thick ITO layer under a gate voltage variation between −2 and 4 V. The rise time and fall time are defined as the time period for 10% to 90% NITO variation. The red curve represents the device with a reduced R, which can be achieved by reducing TiN resistivity, increasing TiN layer thickness, and/or shortening the lateral distance of two electrodes.

Download Full Size | PDF

4. Optical simulation

4.1 Integration with high-index-contrast waveguides

The modulator integrated with a high-index-contrast stripe waveguide such as Si is first investigated. The mode properties of Cu/3-nm-ITO/5-nm-HfO2/5-nm-TiN/Si HPWs are calculated using the eigen-mode expansion (EME) method [27]. The HPW may contain many modes, depending on the size of the Si core. Among them two fundamental plasmonic modes can be distinguished: one is the quasi-TE mode whose dominate electric field component is along the x axis (Ex) and the other is the quasi-TM mode whose dominate electric field component is along the y axis (Ey). Figure 5 plots propagation loss (α), its differentiation (dα/dNITO), real part of the effective mode index (neff), and ratio of electric intensity in the ITO layer (which is defined as the electric field confinement factor) of these two fundamental plasmonic modes as a function of NITO for HPWs with different Si core sizes. One sees that the propagation loss and the ratio of electric field in the ITO layer increase with NITO increasing, reach maximum at a certain NITO, and then decrease with NITO further increasing. NITO for the maximum ratio of electric intensity in the ITO layer is very near the ENZ point (which is indicated by the dash lines in the figure), whereas NITO for the maximum α is slightly larger the ENZ point. Meanwhile, neff reaches minimum near the ENZ point and increases quickly with NITO further increasing. The value of this certain NITO is almost independent on the Si core size and the polarization. For EA application, the condition when the maximum α is reached can be defined as the “OFF”-state, and the condition when α is much smaller can be defined as the “ON”-state. Here we assume NITO = 7 × 1020 cm−3 for the “OFF”-state whereas NITO for the “ON”-state can be either >> 7 × 1020 cm−3 or << 7 × 1020 cm−3. Accordingly, the modulator works on either the accumulation mode, i.e., varying NITO from << 7 × 1020 cm−3 to 7 × 1020 cm−3, or the depletion mode, i.e., varying NITO from >> 7 × 1020 cm−3 to 7 × 1020 cm−3. From Fig. 5, one sees that the absolute value of dα/dNITO in the smaller NITO side is slightly larger than that in the larger NITO side. More importantly, α at (7 × 1020 cm−3ΔNITO) is slightly smaller than α at (7 × 1020 cm−3 + ΔNITO), where ΔNITO is the voltage induced modification of NITO. It indicates that the accumulation mode is more effective and also has smaller insertion loss than the depletion mode. Here we simply assume NITO = 2 × 1020 cm−3 for the “ON”-state.

 figure: Fig. 5

Fig. 5 The mode properties, i.e., the propagation loss α, the α differentiation dα/dNITO, the real part of the effective mode index neff, and the ratio of electric field intensity in the ITO layer, of the Cu/3-nm-ITO/5-nm-HfO2/5-nm-TiN/Si HPWs versus the electron concentration NITO in the 3-nm ITO layer for (a) quasi-TE mode and (b) quasi-TM mode. The dash lines represent the ENZ point.

Download Full Size | PDF

To understand the basis for this huge variation in the HPW’s optical property induced by NITO variation, 2D distributions of the dominant electric field component (i.e., Ex for the TE mode and Ey for the TM mode) are depicted in Figs. 6(a)6(d) for a HPW with 400-nm × 340-nm Si core at the “ON” and “OFF” states. One sees Ex of the TE mode is confined at the two sidewalls of the Si core and Ey of the TM mode is confined at the top of the Si core. This is because the plasmonic mode has its dominant electric field perpendicular to the metal/dielectric interface and the electric field intensity in a thin layer is inversely proportional to its absolute permittivity value due to the continuity of electric displacement normal to the interfaces. For both TE and TM modes, the electric field in the Si core at the “OFF”-state is much weaker than that at the “ON”-state. To see the electric field distributions more clearly, Fig. 6(e) plots normalized Ex taken along the a-a’ lines and Fig. 6(f) plots normalized Ey taken along the b-b’ lines, respectively. The corresponding electric field distribution in the 400-nm × 300-nm Si stripe waveguide is also shown for comparison. At the “ON”-state, ITO has positive real part of εITO and its |εITO| (2.7) is close to that of HfO2 (3.5), thus it plays a similar role as the dielectric HfO2. As a result, the HPW is similar to the previously reported Cu/insulator/Si/insulator/Cu waveguide for the TE mode [28] or the Cu/insulator/Si waveguide for the TM mode [29], which has a relatively small α. Here, the insulator layer is the sum of the ITO layer and the HfO2 layer. The thin TiN layer plays a minor role on the HPW’s optical property because the electric field intensity in the TiN layer is very small. With NITO increasing, |εITO| decreases and the imaginary part of εITO increases, thus both the electric field intensity in the ITO layer and α of HPW increase. At the “OFF”-state, ITO has εITO of −0.3 + 0.62j, the electric field in the ITO layer is dramatically enhanced due to its small |εITO| of 0.69. As a result, the HPW has very large α. With NITO further increasing, |εITO| increases again, making both the electric field intensity in the ITO layer and α of HPW decrease. The property of HPW approaches again to that of the Cu/insulator/Si/ insulator/Cu (for the TE mode) or Cu/insulator/Si waveguide (for the TM mode). Here, the insulator layer is only the HfO2 layer and the thin ITO layer behaves as the thin TiN layer.

 figure: Fig. 6

Fig. 6 2D electric field distribution in the Cu/3-nm-ITO/5-nm-HfO2/5-nm-TiN/400-nm × 340-nm-Si: (a) TE mode, Ex at “ON”-state, (b) TM mode, Ey at “ON”-state, (c) TE mode, Ex at “OFF”-state, and (d) TM mode, Ey at “OFF”-state. Normalized 1D distributions near the ITO interface: (e) along the a-a’ line for the TE mode and (f) along the b-b’ line for the TM mode, the inset are the distributions in the whole structure as well as the distributions in the 400-nm × 340-nm Si waveguide.

Download Full Size | PDF

Although NITO for the maximum α is almost independent on the Si core size, α at the “ON”-state (αon) and “OFF”- state (αoff) depend on the Si core size. To reveal the influence of the Si core size on the modulator’s performance, a figure of merit defined as FoM=αoffαonαon is plotted in Fig. 7 as a function of the Si core’s size, which reflects a tradeoff between the modulation depth and the insertion loss. For the TE mode, FoM depends on W weakly since αoff increases approximately FoM-times larger than αon with W decreasing, whereas FoM increases with h increasing since αon decreases while αoff keeps almost the same with h increasing. This is because with h increasing the effect of the top stack layers becomes smaller and the HPW approaches to the standard horizontal Cu/dielectric/Cu structure. For the TM mode, FoM depends on h weakly when h is larger than ~220 nm and FoM increases with W increasing. This is because with W increasing the effect of the sidewall stack layers becomes smaller and the HPW approaches to the standard vertical Cu/insulator/Si/insulator/Si structure. If h is too small, the fundamental TE mode is difficult to be excited, and if W is too small, the fundamental TM mode is difficult to be excited. Instead, modes whose electric field is not confined in the Cu/dielectric interface may be excited. One of such modes is shown by the dashed curves in Fig. 5(b) for example. The propagation loss of this mode depends on NITO weakly, thus it does not contribute to the EA modulation. It indicates that a larger Si core is benefit for our EA modulator, in contrast to the previously reported Cu/insulator/Si/insulator/ Si or Cu/insulator/Si waveguide where a small Si core is preferred to confine the optical mode tightly. Therefore, the Si core of the EA modulator can be set to be the same as that of the input/output Si waveguides. Here we set W = 400 nm and h = 340 nm, which is typical for a single-mode Si stripe waveguide.

 figure: Fig. 7

Fig. 7 FoM of HPWs with different Si core sizes for (a) quasi-TE mode and (b) quasi-TE mode.

Download Full Size | PDF

The 3D finite-difference time-domain (FDTD) method [27] is used to simulate a 1-μm-long EA modulator inserted in the 400-nm × 340-nm Si stripe waveguide. A 1.55-μm TE or TM light is launched into the left Si waveguide, propagating through the modulator, and then being monitored in the right Si waveguide. Figure 8(a) shows the absolute value of Poynting vector along the Y-cut at the center of the Si waveguide under the TE excitation at “ON” and “OFF”-states. The output power is 51% (−2.9 dB) at the “ON”-state and is 0.53% (−22.8 dB) at the “OFF”-state, indicating that the modulator has insertion loss of 2.9 dB and modulation depth of 19.9 dB. The mode profiles (i.e., the Ex distributions) in the input Si waveguide, the middle of the modulator, and the output waveguide at the “ON”-state are depicted in Fig. 8(b). The output light keeps the same TE polarization as the input light. The mode profile in the modulator shows a clear Ex enhancement in the ITO/HfO2 layer at both sides, indicating the excitation of the plasmonic TE mode. However, the Ex profile differs slightly from that shown in Fig. 6(a) probably because of excitation of other modes besides the fundamental TE mode in the modulator. At the “OFF”-state, the plasmonic mode excited in the modulator is quickly attenuated.

 figure: Fig. 8

Fig. 8 (a) The absolute value of Poynting vector along the Y-cut at the center of the Si waveguide under the TE excitation at the “OFF” and “ON” states, (b) Ex distributions in the input Si waveguide, the middle of modulator, and the output Si waveguide, (c) The absolute value of Poynting vector along the X-cut at the center of the Si waveguide under the TM excitation at the “OFF” and “ON” states, (d) Ey distributions in the input Si waveguide, the middle of modulator, and the output Si waveguide.

Download Full Size | PDF

The modulator under the TM excitation exhibits similar property. Figure 8(c) shows absolute value of Poynting vector along the X-cut at the center of the Si waveguide at “ON” and “OFF”-states. The output power is 48% (−3.2 dB) at the “ON”-state and is 1.5% (−18.4 dB) at the “OFF”-state, corresponding to insertion loss of 3.2 dB and modulation depth of 15.2 dB. Figure 8(d) shows mode profiles (i.e., the Ey distributions) in the input Si waveguide, the middle of the modulator, and the output waveguide at the “ON”-state. The output light keeps the same TM polarization as the input light. Again, the mode profile in the modulator differs slightly from that shown in Fig. 6(b) probably due to the excitation of other modes besides the fundamental plasmonic TM mode.

The transmission powers through the EA modulators with different lengths (L) at the “ON” and “OFF”-states are plotted in Fig. 9. The transmission at the “ON”-state, which is defined as the insertion loss, increases with L increasing. From linearly fitting, the propagation loss is extracted to ~1.9 (~1.3) dB/μm for the TE (TM) mode, close to that obtained from the EME calculation (Fig. 5). The coupling loss between the modulator and the Si waveguide is extracted to be ~0.8 (~1.1) dB/fact for the TE (TM) mode, which is determined by the mode-index difference and the mode overlap between the modulator and the input/out waveguide. The modulator also shows a weak Fabry-Perot resonance due to weak reflection at the waveguide/modulator facets, which deviates the insertion loss from the exactly linear dependence on L. At the “OFF”-state, the output power decreases quickly with L increasing and then depends on L weakly with L further increasing, which may be attributed to the excitation of low-loss modes besides the fundamental TE/TM plasmonic mode in the modulator, as observed in Figs. 8(b) and 8(d). As a result, the modulator has large modulation depth of ~15 dB/μm at the small L region while the modulation depth maximizes when L becomes longer. For the TE mode, the maximum modulation depth is ~32 dB when L ≥ ~2 μm, and for the TM mode, the maximum modulation depth is ~15.2 dB when L ≥ ~1 μm.

 figure: Fig. 9

Fig. 9 Transmitted powers of an EA modulator inserted in a 400-nm × 340-nm Si stripe waveguide at the “ON” and “OFF”-states as a function of the modulator’s length. The EA modulator has the same dimension of Si core as the input/output Si waveguide.

Download Full Size | PDF

Large modulation depth is also observed for other TCO-based EA modulators reported in literature. For example, a vertical Si/SiO2/TCO/Si EA modulator reported in [14] exhibits modulation depth of 3 dB when L = 0.25 μm. However, the vertical Au/SiO2/ITO/Si EA modulator reported in [15] has a relatively low modulation depth of ~1 dB/μm. The difference can be attributed to the variation range of the TCO property between the “ON” and “OFF” states. In both ours and [14], the TOC property is changed through the ENZ point between the “ON” and “OFF” states, whereas in [15], the TCO property at the “OFF”-state is still far from the ENZ point. Form Fig. 5, one sees that a large dα/dNITO is obtained only near the ENZ point.

4.2 Integration with low-index-contrast waveguides

As the above-mentioned, the Cu/ITO/HfO2/TiN stack can be deposited on a low-index-contrast waveguide such as Si3N4 to form an EA modulator. Passive Cu-Si3N4-Cu or Cu-SiO2-Si3N4-SiO2-Cu plasmonic waveguide components for CMOS backend integration has been demonstrated [30], but Si3N4 waveguide-based active devices are missing. The optical properties of Cu/3-nm-ITO/5-nm-HfO2/5-nm-TiN/Si3N4 HPWs are shown in Fig. 10 as a function of NITO. Similar to the Si-core counterparts, the propagation loss and the ratio of electric field intensity in the 3-nm ITO layer increases with NITO increasing, reaches maximum near the ENZ point, and then decreases with NITO further increasing. Meanwhile, neff shows large variation near the ENZ point. NITO for the maximum α and that for the maximum ratio of electric field in the ITO layer are similar to the Si-core counterparts. Therefore, we still define “ON” state at NITO = 2 × 1020 cm−3 and “OFF”-state at NITO = 7 × 1020 cm−3.

 figure: Fig. 10

Fig. 10 The mode properties (i.e., the propagation loss α, the real part of the effective mode index neff, and the ratio of electric field intensity in the ITO layer) of the Cu/3-nm-ITO/5-nm-HfO2/5-nm-TiN/Si3N4 HPWs as a function of average electron concentration NITO in the 3-nm ITO layer for (a) quasi-TE mode and (b) quasi-TM mode. The dash lines represent the ENZ point.

Download Full Size | PDF

However, αoff of the Si3N4-based HPWs is much smaller than that of the Si-based HPWs because of the low index contrast of Si3N4/SiO2. For the TE mode, αoff is 3.3 dB/μm for the 800-nm × 600-nm Si3N4 core HPW and is 5.1 dB/μm for the 400-nm × 600-nm Si3N4 core HPW. Thus, a narrower Si3N4 core is preferred to reach a larger modulation depth for the TE mode. For the TM mode, however, the Si3N4 core is required to be wide enough to excite the fundamental plasmonic TM mode. Otherwise, modes other than the fundamental plasmonic TM mode may be excited, as shown by the dashed curves in Fig. 10(b). It indicates a wider Si3N4 core is preferred for the TM mode. Here, the input/output Si3N4 stripe waveguide is set to W = 800 nm and h = 600 nm, which is typical for a single-mode Si3N4 stripe waveguide. We design the modular for the TE mode has the Si3N4 core of 400-nm × 600-nm, inserted in the 800-nm × 600-nm Si3N4 stripe waveguide through two 1-μm-long tapered couplers, as shown in Fig. 11(a). The tapered coupler has the same Cu/ITO/HfO2/TiN stack as the main body of the modulator, thus they can be regarded as a part of the modulator. For the TM mode, we design the modulator has the same size Si3N4 core of 800-nm × 600-nm as the input/output Si3N4 waveguide.

 figure: Fig. 11

Fig. 11 (a) The absolute value of Poynting vector along the Y-cut at the center of the Si3N4 waveguide under the TE excitation at the “OFF” and “ON” states, the modulator has a 400-nm × 600-nm Si3N4 core inserted in the 800-nm × 600-nm stripe Si3N4 waveguide through 1-μm-long tapered couplers. (b) Ex distributions in the input Si3N4 waveguide, the middle of modulator, and the output Si3N4 waveguide, (c) The absolute value of Poynting vector along the X-cut at the center of the Si3N4 waveguide under the TM excitation at the “OFF” and “ON” states, the modulator has the same 800-nm × 600-nm Si3N4 core as the input/output Si3N4 stripe waveguide. (d) Ey distributions in the input Si3N4 waveguide, the middle of modulator, and the output Si3N4 waveguide.

Download Full Size | PDF

3D FDTD simulation results for such 3-μm-long Si3N4 EA modulators are shown in Fig. 11. At the “ON”-state, the output power is −5.1 (−3.7) for the TE (TM) mode, and at the “OFF”-state, the output power is −13.0 (−18.5) dB for the TE (TM) mode. The mode profiles in the input Si3N4 waveguide, the middle of the modulator, and the output S3N4 waveguide at the “ON”-state are shown in Figs. 11(b) and 11(d) for the TE and TM modes, respectively. Similar to the Si-core counterpart, the output light keeps the same polarization as the input light for both TE and TM modes.

Figure 12 shows 3D FDTD simulation results for the Si3N4 EA modulators whose top views are shown in the inset schematically. The transmission at the “ON”-state decreases with L increasing. From linearly fitting, the propagation loss is extracted to 0.77 (0.32) dB/μm for the TE (TM) mode, close to that obtained from the EME calculation (Fig. 10). The coupling loss between the modulator and the Si3N4 waveguide is extracted to 1.5 (1.2) dB/facet for the TE (TM). At the “OFF”-state, the output power decreases quickly with L increasing first and then depends on L weakly with L further increasing, which can be attributed to the excitation of modes other than the fundamental plasmonic mode, as observed in Figs. 11(b) and 11(d) for the TE and TM modes, respectively. One sees from Fig. 12 that the optimal length for the TE modulator is ~4 μm (including two 1-μm-long coupler and 2-μm-long main body), which provides insertion loss of 6.3 dB and modulation depth of 16.5 dB, and the optimal length for the TM modulator is ~3 μm, which provides insertion loss of 3.7 dB and modulation depth of 14.8 dB. As compared with the Si counterpart, the Si3N4 modulator has poor performance in the viewpoint of the insertion loss, modulation depth, footprint, and the energy consumption (since the energy ∝ L), while, it has similar speed as the Si counterpart (since the speed is limited by the RC delay and CL, R1/L). Nevertheless, our design provides an approach, for the first time to our best knowledge, to directly modulate light propagating along the Si3N4 waveguide.

 figure: Fig. 12

Fig. 12 Transmitted powers of Si3N4 EA modulators inserted in the 800-nm × 600-nm Si3N4 stripe waveguide at the “ON” and “OFF”-states as a function of the modulator’s length. The TE modulator has 400-nm × 600-nm Si3N4 core and two 1-μm-long tapered couplers, and the TM modulator has 800-nm × 600-nm Si3N4 core, as shown schematically in the right.

Download Full Size | PDF

5. Influence of the stack’s parameters and processing tolerance

In the above analysis, the thicknesses and optical properties of the ITO, TiN and HfO2 layers are set to certain values without optimization. The layers’ thicknesses can be precisely controlled by ALD and their properties may be slightly tuned by the deposition conditions. The influences of these parameters on the modulator’s performance are discussed in this section, which can be used to guide the optimization route for the layers’ deposition conditions. Moreover, the influence of possible misalignment during fabrication is also discussed. For simplification, the beneath dielectric core is fixed to the 400-nm × 340-nm Si core. The results are applicable for the Si3N4 core EA modulators.

5.1 The ITO layer

The most important parameter for our EA modulator is the ITO layer and the accumulation layer in this layer. In the above analysis, we simply assume tAcL = tITO = 3 nm. Here we vary the ITO thickness from 3 to 10 nm and assume AcL has thickness of 1, 2, or 3 nm, respectively, i.e., NITO in the AcL region will be changed from 2 × 1020 cm−3 at the “ON”-state to 7 × 1020 cm−3 at the “OFF”-state while NITO out the AcL region keeps 3.5 × 1020 cm−3. Figure 13 plots αon and off – aon) as a function of tITO with different tAcL values. One sees that αon increases slightly and off – aon) decreases significantly with tITO increasing. It indicates that increasing tITO will degrade both the insertion loss and the modulation depth, which can be attributed to the fact that the ITO layer out the AcL region does not contribute to the EA modulation. Due to the same reason, off – aon) depends on tAcL more significantly. In the case of tITO = 3 nm, off – aon) decreases from ~22 dB/μm to ~7 dB/μm when tAcL decreases from 3 nm to 1 nm. By increasing the modulator length from ~1 μm (in the case of tAcL = 3 nm) to ~3 μm (in the case of tAcL = 1 nm), the similar modulation depth can be obtained, however, at a price of increasing the insertion loss from ~3.0 dB to ~7.4 dB. But, the required voltage swing is reduced to ~-0.6–1.2 V if tAcL = 1 nm, as read from Fig. 3(c). As a result, the switching energy per bit is reduced to ~0.1 pJ/bit. On the other hand, the RC-limited modulation speed keeps similar since CL and R1/L.

 figure: Fig. 13

Fig. 13 (a) αon and (b) the difference of αoff and αon of the Cu/ITO/5-nm-HfO2/5-nm-TiN/400-nm × 340-nm-Si HPWs at the “ON” and “OFF”-states as a function of tITO by assuming the AcL thickness tAcL of 1, 2, or 3 nm, respectively. NITO inside AcL is 2 × 1020 cm−3 at the “ON”-state and 7 × 1020 cm−3 at “OFF”-state whereas NITO outside AcL is 3.5 × 1020 cm−3.

Download Full Size | PDF

Other TCOs such as ZnO:Al can also be used as the active material. Their permittivities also obey the Drude model but having different values of ε, Γ, and N0 [31]. For example, ZnO:Al has ε of 3.5 and Γ of 2.3 × 1014 s−1 [32]. Here we still assume tITO = tAcL = 3 nm and NITO at “ON”-state = 2 × 1020 cm−3 while the values of ε and Γ are changed to investigate their effect on the modulator’s performance. In one case, ε varies from 3.1 to 4.1 while keeping Γ = 1.8 × 1014 s−1. In the other case, Γ varies from 1.2 × 1014 s−1 to 2.6 × 1014 s−1 while keeping ε = 3.9. The simulation results are plotted in Fig. 14. One sees from Fig. 14(a) that NITO for the maximum αoff (i.e., the required NITO for the “OFF”-state) decreases with ε decreasing. This is because the ENZ point decreases with ε decreasing according to the Drude mode. The reduction of the required NITO for the “OFF”-state can reduce the operating voltage, thus reduce the consumption energy per bit. Meanwhile, both αon and off – αon) increase slightly with ε decreasing, thus the modulator length can be slightly shortened to keep certain insertion loss and the modulation depth. For comparison, Si has large ε of ~11.7. Its ENZ point is ~1.2 × 1021 cm−2, much larger than that for ITO (~6.5 × 1020 cm−2). As a result, a large voltage is required to reach the sufficient modulation depth for the Si-based plasmonic modulator [10]. It indicates that Si is not a good active material for the plasmonic EA modulator due to its large ε value. On the other hand, Fig. 14(b) shows that the variation of Γ has no influence on NITO for the maximum α, whereas αon decreases and off – αon) increases with Γ decreasing, indicating that both the insertion loss and the modulation depth are improved with Γ decreasing. Overall, we can conclude that a TCO material for the EA modulator requires its ε and Γ values as smaller as better. The small ε value is more critical because it determines the required ΔNITO (i.e., the required voltage swing) for modulation between “ON” and “OFF” states.

 figure: Fig. 14

Fig. 14 αon (at NITO = 2 × 1020 cm−3), αon – αoff, and NITO for the maximum propagation loss of the Cu/ITO/5-nm-HfO2/5-nm-TiN/400-nm × 340-nm-Si HPWs as a function of ε and Γ of ITO: (a) ε varies from 3.1 to 4.1 while Γ = 1.8 × 1014 s−1 and (b) Γ varies from 1.2 to 2.6 while ε = 3.9.

Download Full Size | PDF

5.2 The TiN layer

The difference between our modulator and that reported in [15] is that our modulator has a thin TiN layer to serve as the bottom electrode. Figure 15(a) plots αon, off – αon) and FoM of Cu/3-nm-ITO/5-nm-HfO2/TiN/Si HPWs as a function of tTiN. Although both αon and off – αon) increase with tTiN increasing, FoM decreases significantly with tTiN increasing. It indicates that tTiN is as thinner as better in the viewpoint of the insertion loss. On the other hand, tTiN is as thicker as better in the viewpoint of the RC-limited modulation speed, as observed from Fig. 4. The insertion loss is contributed by the propagation loss ( = αon⋅L, where L is the modulator length) and the coupling loss between the modulator and the input/output waveguides. The coupling loss (including two facets) is ~1.6 (~2.2) dB for the TE (TM) mode and depends on tTiN weakly (not shown here). To balance the insertion loss and the resistance, an optimal tTiN may be set when the propagation loss equals the coupling loss because for the TiN layer thinner than this optimal value, the insertion loss depends on tTiN weakly as the coupling loss dominates. The optimal tTiN is ~5 nm for our 1-μm-long Si EA modulator, as read from Fig. 15(a). A method to reduce the resistance (hence increase the modulation speed) without degrading the insertion loss is to improve the mobility of the TiN layer, which may be achievable by optimization of the TiN deposition condition. Another method is to reduce the lateral distance between the Al contact on the TiN layer and the Cu/ITO/HfO2 stack. However, it may increase the fabrication difficulty.

 figure: Fig. 15

Fig. 15 The influence of the thin TiN layer on the modulator’s performance: (a) TiN has permittivity of −16.66 + 7.52j while its thickness ranges from 0 to 10 nm, (b) TiN has thickness of 5 nm while different permittivity values.

Download Full Size | PDF

The values of TiN permittivity reported in literature are diverse, for example, −16.66 + 7.52j in [22], −83.3 + 21.3j in [20], and −20 + 21j in [31]. We name them as TiN-1, TiN-2, and TiN-3, respectively. Moreover, we assume TiN-4 having the same permittivity as Ag (−86.64 + 8.74j). Figure 15(b) shows αon and off – αon) of HPWs with the above four kinds of TiN layers. One sees that the larger |εTiN| (e.g. the better plasmonic material) results in a slightly smaller αon. However, the dependence of αon and off – αon) on εTiN is relatively weak. Therefore, the optimization for the TiN layer can be focused on its conductivity rather than its permittivity.

5.3 The HfO2 layer

For conventional MOS transistors, the gate dielectric is preferred to have high dielectric constant and thin thickness to reduce the operating voltage. The influence of the HfO2 layer on the optical properties of our EA modulator is plotted in Fig. 16. In one case tHfO2 varies from 1 to 11 nm while its refractive index nHfO2 keeps 1.87. In the other case, nHfO2 varies from 1.44 to 2.4 while tHfO2 keeps 5 nm. One sees that both αon and off – αon) increase with tHfO2 decreasing or with nHfO2 increasing. This is because the optical power contained in the HfO2 layer, which does not contribute to the modulation, decreases with tHfO2 decreasing or with nHfO2 increasing. For the TE mode, FoM is almost independent on tHfO2 and nHfO2, whereas for the TM mode, FoM decreases slightly with tHfO2 decreasing or with nHfO2 increasing. It indicates that the influence of the HfO2 layer on the modulator’s optical property is weak. Therefore, the optimization of the HfO2 layer can be focused on its electrical property.

 figure: Fig. 16

Fig. 16 The influence of the HfO2 layer on the modulator’s performance: (a) HfO2 has refractive index of 1.87 while its thickness ranges from 1 to 12 nm. (b) The gate oxide has thickness of 5 nm while its refractive index ranges from 1.44 to 2.4.

Download Full Size | PDF

5.4. Fabrication tolerances

The proposed modulator can be fabricated using the standard CMOS backend technology, somewhat similar to that for the Si-based plasmonic modulators except ALD of the ultrathin layers [2830]. After the Si (or Si3N4) core formation, a thin TiN layer is deposited and patterned. Then, a thick SiO2 layer is deposited and a window is opened to define the plasmonic area. HfO2, ITO, and Cu are deposited sequentially, followed by chemical mechanical polishing (CMP) to remove the Cu/ITO/HfO2 stack outside the window. In this fabrication flow, a possible misalignment (ΔL) may exist between the TiN layer and the Cu/ITO/HfO2 stack, as shown in Fig. 17(a) schematically. The influence of such a misalignment is studied using the 3D FDTD method and the results are plotted in Fig. 17(b). One sees that both the insertion loss and the modulation depth depend on ΔL weakly. It indicates the device has large processing tolerance. The insertion loss even becomes slightly smaller when ΔL is negative. Therefore, the length of the TiN layer can be designed to be slightly smaller than that of the SiO2 window in which the Cu/ITO/HfO2 stack is filled.

 figure: Fig. 17

Fig. 17 (a) Schematic top view of the modulator, showing a possible misalignment of ΔL between the TiN layer and the Cu/ITO/HfO2 stack, (b) Insertion loss and modulation depth of the 1-μm-long Si EA modulator as a function of the misalignment ΔL.

Download Full Size | PDF

6. Conclusion

An EA modulator applicable for both TE and TM modes is proposed and investigated theoretically. The modulator offers various advantages including ultra-compact size (~1 μm2), high speed (>10 GHz), low energy consumption (~0.4 pJ/bit), broad optical bandwidth (originated from the nature of EA modulation), integration with all kinds of waveguides, large processing tolerance, and backend CMOS compatibility. It is almost ideal for seamless integration in the existing EPICs, especially the 3D integrated EPICs where multiple photonic layers are stacked above the electric layer. A key challenge to realize such an “ideal” modulator is optimization of the atomic layer deposited ITO layer. The electron concentration of the as-deposited ITO should be about half of the ENZ point (here, ~3.5 × 1020 cm−3) and its ε and Γ values are as smaller as better.

Acknowledgment

This work was supported by Singapore A*STAR-MINDEF Joint Grant 122-331-0074.

References and links

1. D. J. Lockwood and L. Pavesi, Silicon Photonics II: Components and Integration (Springer, 2011).

2. G. T. Reed, G. Mashanovich, F. Y. Gardes, and D. J. Thomson, “Silicon optical modulators,” Nat. Photon. 4(8), 518–526 (2010). [CrossRef]  

3. D. Marris-Morini, L. Vivien, G. Rasigade, J.-M. Fedeli, E. Cassan, X. Le Roux, P. Crozat, S. Maine, A. Lupu, P. Lyan, P. Rivallin, M. Halbwax, and S. Laval, “Recent progress in high-speed silicon-based optical modulators,” Proc. IEEE 97 (7), 1199–1215 (2009). [CrossRef]  

4. A. Biberman, K. Preston, G. Hendry, N. Sherwood-Droz, J. Chan, J. S. Levy, M. Lipson, and K. Bergman, “Photonic network-on-chip architectures using multilayer deposited silicon materials for high-performance chip multiprocessors,” ACM J. Emerging Technol. Comput. Syst. 7, 7 (2011).

5. N. Sherwood-Droz and M. Lipson, “Scalable 3D dense integration of photonics on bulk silicon,” Opt. Express 19(18), 17758–17765 (2011). [CrossRef]   [PubMed]  

6. J. Hu, L. Li, H. Lin, P. Zhang, W. Zhou, and Z. Ma, “Flexible integrated photonics: where materials mechanics and optics meet,” Opt. Mater. Express 3(9), 1313–1331 (2013). [CrossRef]  

7. D. K. Gramotnev and S. I. Bozhevolnyi, “Plasmonics beyond the diffraction limit,” Nat. Photon. 4(2), 83–91 (2010). [CrossRef]  

8. K. F. MacDonald and N. I. Zheludev, “Active plasmonics: current status,” Laser Photon. Rev. 4(4), 562–567 (2010). [CrossRef]  

9. J. A. Dionne, K. Diest, L. A. Sweatlock, and H. A. Atwater, “PlasMOStor: a metal-oxide-Si field effect plasmonic modulator,” Nano Lett. 9(2), 897–902 (2009). [CrossRef]   [PubMed]  

10. S. Zhu, G. Q. Lo, and D. L. Kwong, “Electro-absorption modulation in horizontal metal-insulator-silicon-insulator-metal nanoplasmonic slot waveguides,” Appl. Phys. Lett. 99(15), 151114 (2011). [CrossRef]  

11. S. Zhu, G. Q. Lo, and D. L. Kwong, “Phase modulation in horizontal metal-insulator-silicon-insulator-metal plasmonic waveguides,” Opt. Express 21(7), 8320–8330 (2013). [CrossRef]   [PubMed]  

12. E. Feigenbaum, K. Diest, and H. A. Atwater, “Unity-order index change in transparent conducting oxides at visible frequencies,” Nano Lett. 10(6), 2111–2116 (2010). [CrossRef]   [PubMed]  

13. M. A. Noginov, L. Gu, J. Livenere, G. Zhu, A. K. Pradhan, R. Mundle, M. Bahoura, Y. A. Barnakov, and V. A. Podolskiy, “Transparent conductive oxides: plasmonic materials for telecom wavelengths,” Appl. Phys. Lett. 99(2), 021101 (2011). [CrossRef]  

14. Z. Lu, W. Zhao, and K. Shi, “Ultracompact electroabsorption modulators based on tunable epsilon-near-zero-slot waveguides,” IEEE Photon. J. 4(3), 735–740 (2012). [CrossRef]  

15. V. J. Sorger, N. D. Lanzillotti-Kimura, R. M. Ma, and X. Zhang, “Ultra-compact silicon nanophotonic modulator with broadband response,” Nanophotonics 1(1), 17–22 (2012). [CrossRef]  

16. A. Melikyan, T. Vallaitis, N. Lindenmann, T. Schimmel, W. Freude, and J. Leuthold, “A surface plasmon polariton absorption modulator,” in Conf. on Lasers and Electro-optics (CLEO), Optical Society of America, San Jose, California (2010), p. JThE77.

17. A. Melikyan, N. Lindenmann, S. Walheim, P. M. Leufke, S. Ulrich, J. Ye, P. Vincze, H. Hahn, T. Schimmel, C. Koos, W. Freude, and J. Leuthold, “Surface plasmon polariton absorption modulator,” Opt. Express 19(9), 8855–8869 (2011). [CrossRef]   [PubMed]  

18. A. V. Krasavin and A. V. Zayats, “Photonic signal processing on electronic scales: electro-optical field-effect nanoplasmonic modulator,” Phys. Rev. Lett. 109(5), 053901 (2012). [CrossRef]   [PubMed]  

19. A. P. Vasudev, J. H. Kang, J. Park, X. Liu, and M. L. Brongersma, “Electro-optical modulation of a silicon waveguide with an “epsilon-near-zero” material,” Opt. Express 21(22), 26387–26397 (2013). [CrossRef]   [PubMed]  

20. V. E. Babicheva, N. Kinsey, G. V. Naik, M. Ferrera, A. V. Lavrinenko, V. M. Shalaev, and A. Boltasseva, “Towards CMOS-compatible nanophotonics: ultra-compact modulators using alternative plasmonic materials,” Opt. Express 21(22), 27326–27337 (2013). [CrossRef]   [PubMed]  

21. J. W. Elam, D. A. Baker, A. B. F. Martinson, M. J. Pellin, and J. T. Hupp, “Atomic layer deposition of indium tin oxide thin film using na halogenated precursors,” J. Phys. Chem. C 112(6), 1938–1945 (2008). [CrossRef]  

22. http://www.ioffe.ru/SVA/NSM/nk/Nitrides/Gif/tini.gif.

23. S. Roberts, “Optical properties of copper,” Phys. Rev. 118(6), 1509–1518 (1960). [CrossRef]  

24. https://www.synopsys.com/TOOLS/TCAD/DEVICESIMULATION/Pages/TaurusMedici.aspx.

25. H. K. Kim, J. Y. Kim, J. Y. Park, Y. Kim, Y. D. Kim, H. Jeon, and W. M. Kim, “Metalorganic atomic layer deposition of TiN thin films using TDMAT and NH3,” J. Korean Phys. Soc. 41, 739–744 (2002).

26. D. A. Miller, “Energy consumption in optical modulators for interconnects,” Opt. Express 20(S2), A293–A308 (2012). [CrossRef]   [PubMed]  

27. https://www.lumerical.com/.

28. S. Zhu, T. Y. Liow, G. Q. Lo, and D. L. Kwong, “Silicon-based horizontal nanoplasmonic slot waveguides for on-chip integration,” Opt. Express 19(9), 8888–8902 (2011). [CrossRef]   [PubMed]  

29. S. Zhu, G. Q. Lo, and D. L. Kwong, “Experimental demonstration of vertical Cu-SiO2-Si hybrid plasmonic waveguide components on an SOI platform,” IEEE Photon. Technol. Lett. 24(14), 1224–1226 (2012). [CrossRef]  

30. S. Zhu, G. Q. Lo, and D. L. Kwong, “Silicon nitride based plasmonic components for CMOS back-end-of-line integration,” Opt. Express 21(20), 23376–23390 (2013). [CrossRef]   [PubMed]  

31. P. R. West, S. Ishii, G. V. Naik, N. K. Emani, V. M. Shalaev, and A. Boltasseva, “Searching for better plasmonic materials,” Laser Photon. Rev. 4(6), 795–808 (2010). [CrossRef]  

32. F. Michelotti, L. Dominici, E. Descrovi, N. Danz, and F. Menchini, “Thickness dependence of surface plasmon polariton dispersion in transparent conducting oxide films at 1.55 microm,” Opt. Lett. 34(6), 839–841 (2009). [CrossRef]   [PubMed]  

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 (17)

Fig. 1
Fig. 1 (a) 3D view and (b) cross-sectional view of the proposed EA modulator integrated with a stripe dielectric waveguide.
Fig. 2
Fig. 2 The calculated real part and imaginary part of the ITO’s permittivity as a function of electron concentration in ITO, the real part of the permittivity crosses zero at a certain NITO, which is defined as the ENZ point.
Fig. 3
Fig. 3 (a) A simplified 2D structure for electrical simulation. (b) 1D election concentration distribution along y coordinate in the 3-nm ITO layer under different voltages, obtained from the MEDICI simulation. The dashed lines represent the average concentrations in the 3-nm ITO layer. (c) Average NITO in the accumulation layer versus voltage in the case of tAcL = 1, 2, or 3 nm.
Fig. 4
Fig. 4 The transient response of the electron concentration in the 3-nm-thick ITO layer under a gate voltage variation between −2 and 4 V. The rise time and fall time are defined as the time period for 10% to 90% NITO variation. The red curve represents the device with a reduced R, which can be achieved by reducing TiN resistivity, increasing TiN layer thickness, and/or shortening the lateral distance of two electrodes.
Fig. 5
Fig. 5 The mode properties, i.e., the propagation loss α, the α differentiation dα/dNITO, the real part of the effective mode index neff, and the ratio of electric field intensity in the ITO layer, of the Cu/3-nm-ITO/5-nm-HfO2/5-nm-TiN/Si HPWs versus the electron concentration NITO in the 3-nm ITO layer for (a) quasi-TE mode and (b) quasi-TM mode. The dash lines represent the ENZ point.
Fig. 6
Fig. 6 2D electric field distribution in the Cu/3-nm-ITO/5-nm-HfO2/5-nm-TiN/400-nm × 340-nm-Si: (a) TE mode, Ex at “ON”-state, (b) TM mode, Ey at “ON”-state, (c) TE mode, Ex at “OFF”-state, and (d) TM mode, Ey at “OFF”-state. Normalized 1D distributions near the ITO interface: (e) along the a-a’ line for the TE mode and (f) along the b-b’ line for the TM mode, the inset are the distributions in the whole structure as well as the distributions in the 400-nm × 340-nm Si waveguide.
Fig. 7
Fig. 7 FoM of HPWs with different Si core sizes for (a) quasi-TE mode and (b) quasi-TE mode.
Fig. 8
Fig. 8 (a) The absolute value of Poynting vector along the Y-cut at the center of the Si waveguide under the TE excitation at the “OFF” and “ON” states, (b) Ex distributions in the input Si waveguide, the middle of modulator, and the output Si waveguide, (c) The absolute value of Poynting vector along the X-cut at the center of the Si waveguide under the TM excitation at the “OFF” and “ON” states, (d) Ey distributions in the input Si waveguide, the middle of modulator, and the output Si waveguide.
Fig. 9
Fig. 9 Transmitted powers of an EA modulator inserted in a 400-nm × 340-nm Si stripe waveguide at the “ON” and “OFF”-states as a function of the modulator’s length. The EA modulator has the same dimension of Si core as the input/output Si waveguide.
Fig. 10
Fig. 10 The mode properties (i.e., the propagation loss α, the real part of the effective mode index neff, and the ratio of electric field intensity in the ITO layer) of the Cu/3-nm-ITO/5-nm-HfO2/5-nm-TiN/Si3N4 HPWs as a function of average electron concentration NITO in the 3-nm ITO layer for (a) quasi-TE mode and (b) quasi-TM mode. The dash lines represent the ENZ point.
Fig. 11
Fig. 11 (a) The absolute value of Poynting vector along the Y-cut at the center of the Si3N4 waveguide under the TE excitation at the “OFF” and “ON” states, the modulator has a 400-nm × 600-nm Si3N4 core inserted in the 800-nm × 600-nm stripe Si3N4 waveguide through 1-μm-long tapered couplers. (b) Ex distributions in the input Si3N4 waveguide, the middle of modulator, and the output Si3N4 waveguide, (c) The absolute value of Poynting vector along the X-cut at the center of the Si3N4 waveguide under the TM excitation at the “OFF” and “ON” states, the modulator has the same 800-nm × 600-nm Si3N4 core as the input/output Si3N4 stripe waveguide. (d) Ey distributions in the input Si3N4 waveguide, the middle of modulator, and the output Si3N4 waveguide.
Fig. 12
Fig. 12 Transmitted powers of Si3N4 EA modulators inserted in the 800-nm × 600-nm Si3N4 stripe waveguide at the “ON” and “OFF”-states as a function of the modulator’s length. The TE modulator has 400-nm × 600-nm Si3N4 core and two 1-μm-long tapered couplers, and the TM modulator has 800-nm × 600-nm Si3N4 core, as shown schematically in the right.
Fig. 13
Fig. 13 (a) αon and (b) the difference of αoff and αon of the Cu/ITO/5-nm-HfO2/5-nm-TiN/400-nm × 340-nm-Si HPWs at the “ON” and “OFF”-states as a function of tITO by assuming the AcL thickness tAcL of 1, 2, or 3 nm, respectively. NITO inside AcL is 2 × 1020 cm−3 at the “ON”-state and 7 × 1020 cm−3 at “OFF”-state whereas NITO outside AcL is 3.5 × 1020 cm−3.
Fig. 14
Fig. 14 αon (at NITO = 2 × 1020 cm−3), αon – αoff, and NITO for the maximum propagation loss of the Cu/ITO/5-nm-HfO2/5-nm-TiN/400-nm × 340-nm-Si HPWs as a function of ε and Γ of ITO: (a) ε varies from 3.1 to 4.1 while Γ = 1.8 × 1014 s−1 and (b) Γ varies from 1.2 to 2.6 while ε = 3.9.
Fig. 15
Fig. 15 The influence of the thin TiN layer on the modulator’s performance: (a) TiN has permittivity of −16.66 + 7.52j while its thickness ranges from 0 to 10 nm, (b) TiN has thickness of 5 nm while different permittivity values.
Fig. 16
Fig. 16 The influence of the HfO2 layer on the modulator’s performance: (a) HfO2 has refractive index of 1.87 while its thickness ranges from 1 to 12 nm. (b) The gate oxide has thickness of 5 nm while its refractive index ranges from 1.44 to 2.4.
Fig. 17
Fig. 17 (a) Schematic top view of the modulator, showing a possible misalignment of ΔL between the TiN layer and the Cu/ITO/HfO2 stack, (b) Insertion loss and modulation depth of the 1-μm-long Si EA modulator as a function of the misalignment ΔL.
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.