Electro-optic modulators are key elements in high-speed optical telecommunication links and preferably rely on materials with a linear electro-optic effect. Choosing adequate waveguide geometries is a key challenge in the design of electro-optic modulators. While all-dielectric geometries promise high-speed modulation with low propagation loss, their modulation efficiency suffers from low confinement and weak electrical fields, resulting in lengthy devices. Plasmonic geometries, on the other hand, allow for most compact devices featuring highest electro-optical bandwidths, but at the cost of higher losses. Alternatively, hybrid photonic–plasmonic solutions open a sweet spot for high-speed modulators with moderate loss. In this review, we discuss the three waveguide types by analyzing and comparing their performance and their sensitivity to variations in geometry with respect to a choice of the electro-optical Pockels-effect material.
I. INTRODUCTION
On-chip integrated electro-optic (EO) modulators are essential in communications and many other applications to encode electrical information onto an optical signal. They enable higher-quality and higher-speed encoding than would be possible by direct laser modulation.1–3 To form these external modulators, electro-refractive and electro-absorptive EO materials are used to activate integrated optical waveguides and to manipulate either the real or the imaginary part of the waveguide’s effective index.4 A variety of physical EO effects can be exploited in these devices.5 Popular EO effects in semiconductors are based on the free carrier dispersion effect,6 the quantum-confined Stark effect,7 or the Franz–Keldysh Effect.8 On the other hand, the linear electro-optic effect, also called Pockels effect, stands out as an effect that exclusively modulates the phase but not the amplitude of an optical field while offering a large optical bandwidth, a quasi-instantaneous EO response on the femtosecond timescale (for details on the frequency characteristics of the Pockels effect, refer to Wemple et al.9 and Dalton et al.10), and a linear response. In a proper configuration, it hence offers chirp-free operation.5,10–13 These benefits make modulators relying on the Pockels effect the first choice for high-quality and high-capacity optical links, as bulk lithium niobate modulators exemplify, which have been the workhorse in long-haul links for the past decades.
Given that the silicon photonic platform has become the mainstay in the industry, the integration of Pockels effect materials with different silicon-integrated optical waveguiding platforms currently receives strong attention. For instance, the silicon photonics platform has been functionalized with organic polymers,14–16 lithium niobate,17–19 and barium titanate.20–24 Alternative approaches have long relied on III-V materials,25 whereas currently, the lithium niobate on insulator (LNOI) platform has received strong attention.26–33 This is mainly because it relies on the well-accepted and reliable LiNbO3 material, and offers low loss and a relatively compact footprint.
The plasmonic platform extends the photonic platform with plasmonic building blocks, which overcome the classical refraction limit. Plasmonic devices are substrate and platform agnostic,34–37 and also compatible with silicon photonics,15,34,38,39 which makes them very flexible. Furthermore, the plasmonic approach is compatible with a wide variety of nonlinear materials such as organic Pockels-effect materials,15,39 BaTiO3,40 and LiNbO3.41 Even more spectacular, its substrate-agnostic nature means that devices can be fabricated by metal elements and nonlinear material alone.37,42 This allows a monolithic integration of plasmonics with electronic CMOS chips.43–45
Hybrid photonic–plasmonic modulators combine plasmonic and photonic waveguiding principles in a hybridized optical mode,44,46,47 and therefore, their properties form a compromise between those of purely photonic or purely plasmonic devices. The question then is, what can one expect, and how do these different approaches compare against each other?
In this work, we focus on an insightful presentation and comparison between the photonic, the plasmonic, and the photonic–plasmonic hybrid modulator concepts. We first introduce figures of merit for electro-refractive EO modulators. After introducing the most common photonic, plasmonic, and hybrid waveguide geometries, their fundamental performance metrics, and impact on the system level will be discussed and compared. This review has been written to complement and complete recent device-focused review articles by Kim et al.,48 Rahim et al.,49 Sinatkas et al.,4 and Tuniz.50 Note that the fundamental properties of waveguides and the design trade-offs that come with them and are described here, are not only valid for electro-optic modulators, but can be transferred to a wider scope, such as for second-harmonic generation51 or for Kerr nonlinear waveguides.52
A note on nomenclature: Seen from a higher level, the field of plasmonics is part of the field of photonics or nanophotonics. In the scope of this work, though, we focus on waveguiding mechanisms and use the word photonic for all-dielectric structures relying exclusively on photons, whereas we use plasmonic for structures relying on surface plasmon polaritons (SPPs).
A. Performance metrics
Various performance metrics can be used to compare electro-optic modulators to each other.4,25,49,53–57 A first class of metrics describes modulator characteristics in the scope of data transmission systems, whereas a second class of metrics is focused more on the underlying device and waveguide physics. We mostly follow the definitions by Witzens,57 refer to Miller56 for the energy consumption per bit, and to Alferness25 and Brosi et al.58 for the field interaction factor. Table I lists an overview. Effects that arise from scaling to more complex systems, potentially comprising several electro-optic modulators, such as crosstalk, are beyond the scope of this paper. The work of Winzer and Neilson55 and the references therein give a good introduction.
Overview on the performance metrics used to characterize electro-optical modulators. The first block comprises the metrics that are of interest on the system level, whereas the second block lists the metrics commonly used to describe modulator physics. Refer to Ref. 56 (Ebit), Refs. 25 and 58 (Γ), and Ref. 57 (rest).
Metric . | Symbol . | Unit . |
---|---|---|
System-level performance metrics | ||
Insertion loss | IL | dB |
EO bandwidth | f3dB | GHz |
Half-wave voltage | Vπ | V |
On-off voltage | Von/off | V |
Energy per bit | Ebit | fJ bit−1 |
Device-level performance metrics | ||
Propagation losses | α | dB cm−1, …, dB μm−1 |
Voltage–length product | VπL | V cm, …, V μm |
Voltage–length–loss product | VπLα | V dB |
Field interaction factor | Γ | 1; dimensionless |
Metric . | Symbol . | Unit . |
---|---|---|
System-level performance metrics | ||
Insertion loss | IL | dB |
EO bandwidth | f3dB | GHz |
Half-wave voltage | Vπ | V |
On-off voltage | Von/off | V |
Energy per bit | Ebit | fJ bit−1 |
Device-level performance metrics | ||
Propagation losses | α | dB cm−1, …, dB μm−1 |
Voltage–length product | VπL | V cm, …, V μm |
Voltage–length–loss product | VπLα | V dB |
Field interaction factor | Γ | 1; dimensionless |
The system-level metrics can be differentiated into optical, electro-optical, and electrical metrics:
The spectral window or optical bandwidth describes the spectral range in which the modulator can be operated. This can range from tens of nanometers for Mach–Zehnder modulators (MZMs) down to a few picometers for resonant devices. Note that the latter typically have multiple spectral windows, separated by their free spectral range.
The frequency response or electro-optical bandwidth is a measure of how fast the modulator can be driven. Usually, this characteristic is given in terms of the 3-dB bandwidth, f3dB: The 3-dB bandwidth refers to the frequency where the power of the modulation sideband would drop by 3 dB against its reference at low frequencies. This corresponds to a 6-dB drop of the power of the corresponding electrical frequency component after photodetection (). The electro-optic bandwidth can be limited by several factors including limited carrier velocity in semiconductors, velocity mismatch between the photonic and the RF wave in longer travelling wave modulators, or RC time constants determined by the capacitance of the modulators and the resistance of the feeding electrical lines.
Next, the half-wave voltage Vπ describes the required voltage a phase modulator needs to shift the phase of the transmitted light by Δφ = π. Depending on how phase modulators are integrated in intensity modulators, the on–off voltage Von/off of the intensity modulator might not be equal to the Vπ of the constituent phase modulators. This often leads to some ambiguity, most often with Mach–Zehnder modulators driven in the push–pull mode. There, opposite voltages are applied over each arm, resulting in Von/off = 1/2 · Vπ. When using plasmonic modulators, which can be treated as lumped, open RF elements, and feeding them through a common 50 Ω line, another factor 2 is gained through reflection of the voltage signal. Also, driving the lumped elements in a differential way adds another factor of 2,59 resulting in a final Von/off = 1/8 · Vπ of a plasmonic Mach–Zehnder modulator.
Finally, the energy consumption per bit, Ebit, gives the average electrical energy that the modulator requires to encode one bit. This quantity is typically given in units of fJ bit−1 and depends both on the device physics and the employed modulation scheme.55
The device-level metrics are of practical relevance and mostly linked to the waveguide cross section.
The propagation losses α give the optical attenuation per unit length. Depending on the modulator technology, this can range from dB⋅cm−1 to dB ⋅ μm−1.
The voltage–length product VπL is a length-independent measure of the electro-optic performance of linear electro-optic modulators. As the acquired phase shift of light propagating through a modulator is approximately proportional to the modulator’s length L, the necessary Vπ drops accordingly, so that their product VπL remains constant. As the voltage–length product is a property of the waveguide cross section, it can be used to compare different geometries.
Further, the combination of α and VπL can be an interesting figure of merit for the device designers: The so-called voltage–length–loss product VπLα in the somewhat eccentric unit of VdB gives the resulting losses that are incurred if a certain phase modulation Δφ is to be achieved for a given voltage swing. The VπLα accounts for the fact that the half-wave voltage Vπ and the propagation losses α are often subject to design trade-offs. Increasing the length of the modulator will reduce Vπ but increase αL, so that their product remains constant. Therefore, VπLα allows us to compare the performance of waveguide geometries on different technology platforms.
Finally, the field interaction factor Γ quantifies how strongly the externally applied electric field interacts with the optical mode in the modulator waveguide. In modulators relying on the Pockels effect, the field interaction factor is a measure of the sensitivity of the waveguide’s effective index change Δneff to a change of the active material’s refractive index Δnmat, Γ = Δneff/Δnmat.
II. THE POCKELS EFFECT IN ELECTRO-OPTIC MODULATORS
Also, Pockels materials offer virtually instantaneous electro-optic response (the EO bandwidth of the Pockels effect spans several hundreds of GHz up to hundreds of THz,12,27,60–63 corresponding to a femtosecond timescale). Consequently, Pockels modulators allow to encode information at highest quality and up to highest frequency with perfect control over chirp.11,13
A large variety of different Pockels materials is available, which we can categorize into two classes:
The first class comprises organic electro-optic (OEO) materials. They can be deposited by spin coating and have been used to activate silicon photonic or plasmonic slot waveguides,15 in the latter case even monolithically integrated on an electronic driver chip.43 In OEO materials, electrons oscillating in an asymmetric potential are responsible for the femtosecond-scaled EO response, leading to a flat frequency response of the Pockels effect from DC to hundreds of THz.10 However, organic materials have encountered reservations regarding their long-term thermal and photochemical stability.10,64 More recently, though, new OEO materials, which can be stabilized via a crosslinking process, have been introduced, which can withstand elevated temperatures for extended times.65–70
The second class comprises solid-state materials, such as lithium niobate, barium titanate, or lead zirconate titanate (PZT).71,72 These materials are more stable toward temperature influences,40,71 and forming electro-optic devices relies on epitaxial growth, wafer bonding, or a combination thereof.24,72,73 Thin films of lithium niobate (LiNbO3)—in the bulk form a long-time work horse for long-haul communications—have been used to activate the silicon and silicon nitride platforms with a Pockels material.74,75 More prominently, though, LiNbO3 is used on silicon oxide to form the lithium niobate on insulator (LNOI) platform.26,28,73,76 Modulators with barium titanate (BaTiO3) instead of LiNbO3 potentially offer a much stronger Pockels effect24 and have also yielded first results,21,40,77 while the performance of OEO- and LiNbO3-based modulators has not yet been matched.
In the following, the Pockels effect is introduced in a general form before we consider waveguide-based phase modulators.
A. The Pockels effect
B. Waveguide-based modulators with Pockels materials
Geometrical depiction of an electro-optic phase modulator. An electrical signal Vext is applied to two electrodes (distance ), which encompass a waveguide structure (length L) with an electro-optic material. The resulting electric field (EΩ ≈ Vext/d) induces a phase shift Δφ onto the light passing through the device. Figure adapted from Ref. 80.
Geometrical depiction of an electro-optic phase modulator. An electrical signal Vext is applied to two electrodes (distance ), which encompass a waveguide structure (length L) with an electro-optic material. The resulting electric field (EΩ ≈ Vext/d) induces a phase shift Δφ onto the light passing through the device. Figure adapted from Ref. 80.
Equation (10) tells us that there are three clear routes to increasing the modulation efficiency of a waveguide modulator: First, by selecting a material with a strong Pockels effect, i.e., large , second, by reducing the electrode distance d, and third, by designing the waveguide geometry such that the efficiency Γ = Δneff/Δnmat becomes large, i.e., that optical and RF fields interact strongly. The second and third options are of geometrical nature and influenced by the waveguide refractive index profile and electrode design, as will be discussed below. However, design compromises are often unavoidable. For example, increasing the field interaction factor Γ = Δneff/Δnmat might require a wider waveguide, which in turn necessitates an increase of the electrode spacing d and hence reduces the electric field strength EΩ.56
C. Field interaction factor and sensitivity
The previously introduced field interaction factor, Γ = Δneff/Δnmat, may become complex to determine, particularly when considering the vectorial nature of both optical and external electrical fields together with the full Pockels tensor: Among the three components of the external field and the three components of the optical field all components that are connected by a nonvanishing rijk will contribute toward Γ. In the general case, the reduction of dimensionality from a tensor to the scalars reff and Γ is not trivial and requires simplifying assumptions.
A more general route of calculating the field interaction factor could start from Eq. (7) and its formulation Γ = Δneff/Δnmat. Using a numerical mode simulation tool, one could simulate Δneff for both Δnmat = 0 and an arbitrary choice of Δnmat ≠ 0 and directly calculate Γ. However, this neglects the tensorial nature of the Pockels effect, the inhomogeneity and anisotropy of as well as the field distributions of and . This matters particularly for plasmonic modes, where spatial changes in the field components of and occur over very short distances. When anisotropic EO materials such as BaTiO3 are employed, a full tensorial description in simulation is necessary.
When Heni compared values of Γ obtained by a closed-form expression based on Eq. (12) with those obtained by simulation as described below, it was found that the approximation underestimates Γ consistently for a sweep of plasmonic slot widths and metal thicknesses84 (cf. p. 32).
D. Comparing waveguide geometries and materials
Ultimately, we are interested in minimizing the metrics VπL and VπLα. As discussed above, we can maximize the Pockels effect and the field confinement factor Γ and minimize the propagation losses α over the length L. The latter two parameters Γ and α depend on the waveguide geometry. An important question then is: Which waveguide geometry gives the highest modulation efficiency (with respect to Γ) and lowest losses α? The answer should be given for important nonlinear materials with refractive indices ranging from nmat = 1.7 to nmat = 2.3. The challenge now is that both the field interaction factor Γ and the Pockels effect strongly depend on the refractive index nmat. To clearly distinguish the effect of the waveguide geometry with the field interaction factor and the nonlinearity, we will maintain a constant nonlinearity of the materials by considering throughout our discussion of waveguides, corresponding to an organic electro-optic material with nmat ≈ 1.8 and reff ≈ 170 pm/V, as in Ref. 67. This results in the parameter combinations listed in Table II.
Overview of the active materials used for simulation purposes. The choice of the values results in a constant nonlinear effect , consistent with organic electro-optic material with nmat ≈ 1.8 and reff ≈ 170 pm/V.67 This allows us to compare geometries independent of the material. Readers can deduce the figures of merit for a certain material with a given nmat by a linear scaling. Example: To deduce VπL of a LiNbO3 modulator (nmat ≈ 2.2), scale the simulation value of VπL for nmat = 2.2 given throughout the paper by a factor of 3.
nmat . | . | Remarks . |
---|---|---|
1.7 | 204 | Approximately organic EO material, e.g., Ref. 85 |
1.8 | 171 | Approximately organic EO material, e.g., Ref. 67 |
1.9 | 146 | Approximately organic EO material, e.g., Ref. 67 |
2.0 | 125 | Included for completeness |
2.1 | 108 | Included for completeness |
2.2 | 94 | ∼3× stronger than LiNbO3’s r33 (, Ref. 86) |
2.3 | 82 | ∼4× weaker than BaTiO3’s r33 (, Ref. 24) |
nmat . | . | Remarks . |
---|---|---|
1.7 | 204 | Approximately organic EO material, e.g., Ref. 85 |
1.8 | 171 | Approximately organic EO material, e.g., Ref. 67 |
1.9 | 146 | Approximately organic EO material, e.g., Ref. 67 |
2.0 | 125 | Included for completeness |
2.1 | 108 | Included for completeness |
2.2 | 94 | ∼3× stronger than LiNbO3’s r33 (, Ref. 86) |
2.3 | 82 | ∼4× weaker than BaTiO3’s r33 (, Ref. 24) |
It should also be noted, that in this work, we assume the Pockels coefficient reff to be constant for any modulation frequency , although this is not the case for all materials: Ferroelectrics, for instance, have an additional strain-optic contribution linked to their piezoelectricity at lower frequencies.12 However, taking all these effects into account would go beyond the scope of this work, and the frequency response can be considered by inverse-linearly scaling VπL and VπLα. Be aware that the scaling holds only if all tensorial contributions to reff scale alike with frequency, which might not be the case for low-frequency contributions in mechanically clamped materials with a piezo-electric response.
In the figures illustrating the simulation results per waveguide geometry, we always plot the field interaction factor Γ, the propagation losses α, the voltage–length product VπL and the voltage–length–loss product VπL · α. In many geometries, we will find a trade-off between the modulator’s VπL or Γ and its propagation loss α, and we hope that the consistent depiction helps to illustrate this effect. The method used to perform the simulations for the result diagrams is described in the supplementary material.
III. PLASMONIC, PHOTONIC, AND HYBRID MODULATORS
Subsequently, we compare three fundamental waveguiding concepts that are currently of high interest in electro-optic waveguide modulators.
First, photonic waveguide modulators, where the refractive index profile of the waveguide is formed such that light is guided either in the material with the highest refractive index, or in a narrow, low-index gap between higher-index materials. Photonic waveguides can have very low propagation losses, but the diffraction limit sets a lower boundary to the mode size and confinement.
Second, and on the other extreme, plasmonic waveguide modulators utilizing metal–dielectric interfaces to tightly confine the optical energy. Metal–insulator–metal (MIM) gap waveguides exhibit no cutoff, so that modes can be confined to nanometer-sized gaps.87,88 The big challenge of plasmonic geometries is their fundamental optical attenuation, which comes from Ohmic losses of electrons in the metal, which oscillate as part of the plasmonic electromagnetic wave.
Third, hybridizing a photonic and a plasmonic mode, e.g., by introducing a thin lower-index layer between a high-index material and a metal electrode, potentially offers a compromise between highest-speed plasmonic and low-loss photonic concepts.89–91
Figure 2 exemplifies some waveguide geometries, with photonic concepts in the first row, plasmonic in the second, and hybrid photonic–plasmonic in the third. We have tried to list typical geometries, although literature knows much more variety and the usage of other materials, such as transparent conductive oxides could even blur the boundaries between the categories. For photonics, these are silicon–organic hybrid (SOH) waveguides,15 lithium niobate on insulator,29,73 and barium titanate on insulator.21,77 For the plasmonic geometries, we have chosen the established plasmonic–organic hybrid waveguide geometry,15 its ferroelectric-filled variant with similarities to a rib waveguide,40 and a vertically stacked metal–insulator–metal waveguide.37,92 For the hybrid case, we have chosen the most basic geometry, one width additional lateral contact electrodes as needed for electro-optic modulation.44,47,93 Based on this second geometry, we propose a variant comprising a partially etched section improving the modal confinement. More references to literature on photonic, plasmonic, and their hybrid waveguide geometries are given in the respective sections below.
Cross sections of exemplary [(a)–(c)] photonic, [(d)–(f)] plasmonic, and [(g)–(i)] hybrid photonic–plasmonic waveguides for electro-optic modulator applications. The marks for the electrode distance d indicate the distance over which the external voltage Vext drops. (a) Silicon slot waveguide filled with an organic EO material. The silicon rails are doped to induce electrical conductivity. (b) A LiNbO3 slab waveguide with lateral metal electrodes. (c) A BaTiO3 modulator, where the optical confinement is provided by a deposited Si strip. (d) A plasmonic slot waveguide filled with an organic EO material. (e) Plasmonic slot waveguide formed with a BaTiO3 slab. (f) A vertically arranged plasmonic slot waveguide. (g) Archetype of a hybrid photonic–plasmonic waveguide where the organic EO material is sandwiched between a doped Si strip and a metal top electrode. (h) For operation, the bottom Si electrode must be electrically connected. (i) A partial etch, or sub-wavelength grating can improve modal confinement.
Cross sections of exemplary [(a)–(c)] photonic, [(d)–(f)] plasmonic, and [(g)–(i)] hybrid photonic–plasmonic waveguides for electro-optic modulator applications. The marks for the electrode distance d indicate the distance over which the external voltage Vext drops. (a) Silicon slot waveguide filled with an organic EO material. The silicon rails are doped to induce electrical conductivity. (b) A LiNbO3 slab waveguide with lateral metal electrodes. (c) A BaTiO3 modulator, where the optical confinement is provided by a deposited Si strip. (d) A plasmonic slot waveguide filled with an organic EO material. (e) Plasmonic slot waveguide formed with a BaTiO3 slab. (f) A vertically arranged plasmonic slot waveguide. (g) Archetype of a hybrid photonic–plasmonic waveguide where the organic EO material is sandwiched between a doped Si strip and a metal top electrode. (h) For operation, the bottom Si electrode must be electrically connected. (i) A partial etch, or sub-wavelength grating can improve modal confinement.
A. Photonic
Photonic waveguide-based electro-optic modulators are widely available. Metal-diffused waveguides in lithium niobate crystals94 have paved the way for generations of low-loss, high-performance electro-optic modulators.76 In parallel, optical platforms based on semiconductors have been investigated for their strong electro-optic effects and the semiconductor industry’s fabrication expertise. Indium phosphide (InP) and gallium arsenide (GaAs) have been explored first for their linear electro-optic effect25 and later incorporated in multi-quantum well heterostructures to form electro-refractive95 and electro-absorptive96–98 modulators based on the quantum-confined Stark effect.99,100 InGaAs/InP now forms some of the fastest and most compact modulator technologies, benefiting from the availability of light sources on the same platform.101 Silicon (Si) photonics benefits from established CMOS fabrication processes and techniques,6 now driving the successful and commercial development of compact, fast, and cost-efficient modulators.57 Although Si benefits from the high-yield and high-volume capabilities of the CMOS industry, its carrier-based modulation is limited to bandwidths up to ∼60 GHz in millimeter-long devices.53,102 Functionalization of Si waveguides with other EO materials is investigated to form potentially faster and more compact modulators. Prominent examples are silicon slot waveguides filled with electro-optic organic material, which provide a strong Pockels effect.14,15 Ferroelectric materials might alleviate some of the thermal and chemical concerns faced by organics, and lithium niobate and barium titanate have already been successfully integrated with silicon photonics.21,23,24,81,103 The lithium niobate on insulator (LNOI) technology questions the necessity of a silicon photonic platform, as waveguides and modulators are directly formed out of the electro-optic material, yielding promising high-speed modulators with low optical insertion losses.28
One of the greatest advantages of photonic modulator designs is their potentially low optical insertion loss. State-of-the-art LNOI modulators can reach fiber-to-fiber insertion losses below 5 dB,28,104 followed by InP modulators with ∼9 dB losses.105,106
One of the greatest challenges is the modulation efficiency and the required length of photonic modulators: To avoid metal-induced losses, the metallic electrodes must be kept relatively far from the waveguide. This leads to an increased electrode gap d, which increases the voltage-length product VπL [see Eq. (10)]. An exception are electrodes formed from semi-transparent materials, such as doped Si,14,15 or capacitively coupled electrodes.83
Figure 3 shows numerical simulations of the photonic waveguide geometries from Fig. 2, where the optical electric field is color coded and field lines illustrate the external RF field. In the following, the silicon–organic hybrid (SOH) from Figs. 2(a) and 3(a), as well as the LNOI device [see Figs. 2(b) and 3(b)], are discussed in detail.
Optical field distributions (, colored) and electrical RF fields (, black field lines) for three examples of the photonic waveguide modulator structures plotted in Fig. 2. (a) Silicon–organic hybrid waveguide. The slot between two doped silicon rails is filled with an organic EO material. The RF field is applied over the silicon rails. (b) Lithium niobate on insulator. The active material itself forms the waveguide; the RF field is applied between two metal electrodes. (c) Silicon on barium titanate. The silicon strip provides lateral confinement to the optical mode. The RF field is applied between two metal electrodes.
Optical field distributions (, colored) and electrical RF fields (, black field lines) for three examples of the photonic waveguide modulator structures plotted in Fig. 2. (a) Silicon–organic hybrid waveguide. The slot between two doped silicon rails is filled with an organic EO material. The RF field is applied over the silicon rails. (b) Lithium niobate on insulator. The active material itself forms the waveguide; the RF field is applied between two metal electrodes. (c) Silicon on barium titanate. The silicon strip provides lateral confinement to the optical mode. The RF field is applied between two metal electrodes.
1. Example: Silicon–organic hybrid
Silicon–organic hybrid (SOH) modulators rely on a silicon slot waveguide filled with a nonlinear organic material. Strongly doped, and hence electrically conductive, silicon slabs connect the slot waveguide to metallic electrodes further away from the optical mode. The slot widths d in such configurations ranges from 100 to 200 nm,107 and the organic electro-optic materials used can exhibit nonlinearities up to (n0 ≈ 1.81, in-device r33 = 390 pm/V107), exceeding lithium niobate’s properties by a factor of 7. However, the field interaction factor Γ is modest because Eω and EΩ have only little overlap within the EO material. Figure 4(a) illustrates the dependence of Γ on the slot width d. Note that the effect of doping and carriers in the silicon rails make the modeling not trivial.83 Here, we assume a doping to ρ = 5.5 · 10−4 Ω m, which corresponds to a silicon absorption coefficient of 0.1 mm−1,108 or an imaginary part of the refractive index of . Note that a higher doping concentration typically improves the EO bandwidth at the cost of higher propagation losses. A study on the effect is given in the discussion of hybrid photonic–plasmonic waveguides below. The electrical field is modeled to drop entirely within the slot.83 The Si thickness is 220 nm, and the slab thickness is 70 nm.16
Simulation results for the SOH geometry. nmat is color coded, from nmat = 1.7 (light) to nmat = 2.3 (dark). It is always . Narrow slots and low-index materials yield the best field interaction (a), with best voltage–length product (b) for narrow slots, too. Higher losses in narrow slots (c) are out-weighed by the stronger field, so that the voltage–length–loss product becomes best for narrow slots, too.
Simulation results for the SOH geometry. nmat is color coded, from nmat = 1.7 (light) to nmat = 2.3 (dark). It is always . Narrow slots and low-index materials yield the best field interaction (a), with best voltage–length product (b) for narrow slots, too. Higher losses in narrow slots (c) are out-weighed by the stronger field, so that the voltage–length–loss product becomes best for narrow slots, too.
Populating Eq. (10) with the simulation d of Γ = 0.2, geometrical (d = 150 nm) and material () parameters yields VπL = 840 V µm. This is in good agreement with the measurement value VπL = 640 V µm in Ref. 107.
Although the modulation efficiency VπL and the voltage–length–loss product VπLα are best for small slot widths d, one could think that, rationally, only the narrowest slots should be fabricated. However, at least two findings contradict this conclusion: First, wider slots make the electric field poling of the OEO more efficient,15 yielding higher (the record of r33 = 390 pm/V has been measured in a 190-nm-wide slot107). Second, narrower slots are more difficult to fabricate, and their optical propagation losses tend to be higher than those of wider slots due to higher sensitivity to scattering effects.83
Doping levels do not influence the losses of practically available Si slot waveguides, which typically exceed 0.5 dB/mm108 and can reach up to 2.5 dB/mm in recent realizations and are likely to be dominated by fabrication imperfections.16 To mitigate optical losses induced by doped silicon, the metal electrodes can be capacitively coupled to the Si rail via a high-k material.83
2. Example: Lithium niobate on insulator
Figure 5 visualizes the influence of the electrode separation on a state-of-the-art LNOI modulator geometry, see Fig. 3(b), with an 800-nm-wide, 600-nm-high lithium niobate slab waveguide formed by thinning the surrounding lithium niobate to 300 nm and adding 800 nm SiO2 cladding. The gold electrodes are 1.1 µm thick and their distance d is swept from 2 to 5 µm. The field interaction factor Γ [Fig. 5(a)] stays nearly unchanged by the sweep, with high-index materials yielding the best confinement, Γ ≈ 0.6 for nLNO = 2.2. At the same time, the modulation efficiency increases linearly with decreasing electrode distance, a consequence of the increase of the external RF field (EΩ = Vext/d). On the other hand, the optical propagation losses sharply increase from 0.05 to ∼200 dB/cm when the electrode distance is decreased from 5 to 2 µm. At the same time, the RF propagation losses (not shown) increase with the smaller the electrode distance, amounting already to 7 dB cm−1 for a 5 µm gap and being the main limitation to state-of-the-art LNOI modulators.31 Although the modulation efficiency is best for the smallest electrode distance, this regime is not accessible, in practice, because the RF and optical losses forbid modulators that are sufficiently long to provide a low Vπ. State-of-the-art LNOI modulators as used in high-data-rate experiments are 10 mm (5 mm) long and have a 6-dB (3-dB) bandwidth roll-off of up to 100 GHz.28,29,109 A recent breakthrough using segmented electrodes is discussed below, potentially offering 250 GHz 3-dB bandwidth in a 20-mm-long MZM with VπL = 1.7 V cm and a DC π-voltage of Vπ,DC = 1 V.31
Design trade-offs in LNOI modulators. (a) Decreasing the electrode distance does not increase the field interaction factor, but (b) brings an increase of the electrical field strengths and improves the modulation efficiency, resulting in smaller VπL. (c) On the other hand, electrodes closer to the waveguide contribute to more photonic propagation loss, resulting in larger VπLα (d). However, shorter devices can have a higher electro-optic bandwidth. Normalized to . The design trade-offs in BTO-based photonic modulators are similar.
Design trade-offs in LNOI modulators. (a) Decreasing the electrode distance does not increase the field interaction factor, but (b) brings an increase of the electrical field strengths and improves the modulation efficiency, resulting in smaller VπL. (c) On the other hand, electrodes closer to the waveguide contribute to more photonic propagation loss, resulting in larger VπLα (d). However, shorter devices can have a higher electro-optic bandwidth. Normalized to . The design trade-offs in BTO-based photonic modulators are similar.
3. Frequency response and the challenge of traveling wave electrodes
Electro-optic modulators with traveling wave electrodes show impressive performance and electro-optical bandwidths exceeding 100 GHz.17,18,27,29,31,109 However, designing such electrodes is difficult, as their characteristic impedance needs to be matched to the driving electronics to avoid reflections. Furthermore, three factors largely determine the traveling wave modulators’ frequency response: First, the RF attenuation, second, the RF-optical velocity mismatch, and third, impedance matching between driver electronics and on-chip electrodes.31,57
In a real modulator the frequency response is lower as it may be limited by any or all of the above-mentioned limitations, i.e., by attenuation f3dB,α, or by a walk-off f3dB,τ.
Impedance matching of the traveling wave electrode to the drive electronics is crucial to minimize the RF reflections and make use of the full RF signal. The transferred power fraction to the device is , with the reflection factor , where Z0 is the source impedance and ZL is the load impedance of the travelling wave electrode. The source impedance is often 50 Ω, which is particularly challenging to reach for small electrode separations as desired for efficient modulation: Slowly tapering down the electrode gap from a matched, wider, section, to the active, narrower section can add hundreds of micrometers to the electrode length, with the according RF attenuation.27 Furthermore, ZL is frequency dependent, making the impedance matching even more challenging.
Additionally, the electrodes must be terminated with a matching load to avoid reflections. Reflections may result in standing waves within the modulator, which would lead to resonant enhancement and suppression of certain frequency ranges, resulting in inter-symbol interference in data modulation setups. In laboratory conditions, the termination can be realized with a second microwave probe,16,29,74,111 whereas integrated solutions use on-chip doped silicon112 or NiCr resistors.113 Notably, if the termination can be omitted one might benefit from a voltage doubling as a result of the reflection.59
4. Advantages, challenges, and recent advances
An advantage of photonic waveguides and modulators is the low optical propagation loss. This is particularly true for geometries, such as the LNOI waveguides, which feature on-chip device losses of <0.5 dB.114,115
Challenges faced by photonic modulators are the large size of the photonic modes and/or the large distances between modes and RF electrodes. The large dimensions make large voltages necessary or result in low modulation efficiencies. To compensate for the low efficiency, one may increase the length of the modulator. Yet, this leads to a new design challenge: The trade-off between efficiency and bandwidth: Efficient modulators tend to be long—but come with a reduced bandwidth. High bandwidth modulators are small but require high drive voltages. Depending on the geometry, photonic modulator lengths range from 280 µm16 for high-performance SOH devices to 1–5 mm for LNOI modulators.28,109 Devices larger than 1/10th of the guided RF wavelength (free space wavelength divided by the RF index nRF) cannot be considered lumped anymore.57 For Ω = 2π · 100 GHz and nRF ≈ 2,31, λRF = 1.5 mm, and devices longer than therefore require traveling wave electrodes. In traveling wave modulators, not only the walk-off between optical and electrical signals poses a limit to the modulator’s electro-optic bandwidth but also the attenuation of the traveling RF wave. To reduce the required length and keep up a high bandwidth one could design RF waveguides with a smaller gap. Yet, this comes at a price of higher RF losses (and for very small electrode distance d with high optical losses, too).31
Hence, research focuses on advances in the current photonic modulator designs. In LNOI devices, where RF propagation losses have limited the EO bandwidth, a new segmented electrode design has recently been introduced, which lowers the RF propagation losses at 50 GHz from 7 to 2 dB/cm.31 When a quartz substrate is used instead of the Si substrate, 3-dB bandwidths of 250 GHz have been achieved with 2 cm-long devices, offering a π-voltage at DC of Vπ,DC = 1 V.31
In SOH devices, replacing the resistive coupling between metal electrodes and Si slot by a capacitive coupling via a high-k dielectric has recently been studied.83 However, this approach reduces the field interaction factor as compared to standard SOH modulators and attempts to increase Γ by weakly doping the Si slot waveguide have resulted in higher RF propagation losses.83 For a 50 Ω-matched traveling wave electrode, numerical simulations yield Γ ≈ 0.14 for undoped and Γ ≈ 0.32 for doped silicon, with RF propagation losses increasing from 1.7 to 2.3 dB/mm.83 Measurements show RF propagation losses exceeding 5 dB/mm, yielding similar results as modeled for more traditional, resistively coupled devices.108
Furthermore, efforts are on the way to develop photonic modulators that use new electro-optical materials with a higher nonlinearity. Notably, improved organic linear-electro-optical materials promise electro-optical coefficients in the order of r33 = 1000 pm/V.116 Thin-film barium titanate (BaTiO3) has already been shown to exhibit extremely large Pockels coefficients with r42 = 923 pm/V at nmat ≈ 2.3. According to Eq. (10) and Ref. 40, this should lead to reff = 750 pm/V and . The waveguides of these BaTiO3 modulators are formed by loading the film with silicon nitride103 or silicon.20,21,24 Electro-optic bandwidths of these devices are still below 30 GHz24 or 5 GHz.21,103 However, the use of BaTiO3 is accompanied by at least two major challenges. First, its microwave permittivity is very high, with estimates and measurements ranging around ϵr ≈ 1000.24,117–119 This property is detrimental to the design of travelling wave electrodes, as the microwave signal is slowed down significantly.120 Second, BaTiO3 has a relatively low Curie temperature of ∼120 °C,121,122 leading to intrinsic instability of the spontaneous polarization and requiring the material to be DC-biased during operation. In photonic devices with an electrode gap of 4.75 µm, a bias of ∼10 V was necessary to ensure strong EO modulation.24 Such high DC voltages complicate RF biasing circuits and are not compatible with highest-speed driving electronics. In later experiments, the bias was reduced to ∼2 V for a gap of 2.6 µm, however, with no indication if the bias voltage ensured full RF modulation.120
Apart from BaTiO3, lead zirconate titanate (PZT) has been co-integrated with a SiN waveguide and a Pockels coefficient of reff = 70 pm/V has been measured, which yields .49,123
B. Plasmonic metal–insulator–metal
Plasmonic metal–insulator–metal (MIM) modulators offer a way to overcome the challenges of size and speed, which photonic modulators are bearing. Figures 2(d)–2(f) show typical examples of plasmonic MIM slot waveguides, where the two electrodes can be horizontally separated in-plane [Figs. 2(d) and 2(e)] or vertically separated in an out-of-plane configuration [Fig. 2(f)]. The optical energy is guided in a thin dielectric layer, d ≈ 100 nm thick, consisting of an electro-optic Pockels material and separating the RF electrodes.
Plasmonic modulators are very compact and efficient. They are small because the active dielectric layer and the supported plasmonic modes have sub-wavelength dimensions. Therefore, EΩ = V/d is very large, and the field interaction factor Γ can even become larger than unity due to the slow-down that plasmonic structures bring with them.124 This results in most compact modulators with lengths in the order of ∼10 µm. The compact size brings two more major advantages: First, the parasitic resistance and capacitance are very small, so that the electro-optic bandwidth of these modulators exceeds 500 GHz.62 Furthermore, the devices can be truly treated as lumped, capacitive loads. Hence, the driving voltage supplied from a 50 Ω line is reflected, leading to its effective doubling. All these factors together result in modulators with voltage–length products as small as ∼60 V μm.124 A second benefit arising from the lumped character of the devices is that they can be driven differentially,59 effectively reducing the voltage–length product by another factor of 2. Conversely, the strong confinement and modulation efficiency come at the cost of propagation losses, which are about ∼0.5 dBμm−1.125 Furthermore, the high mode intensity in plasmonic waveguides may lead to the photodegradation of organic electro-optic material107 and hence limit the maximal power that the modulator can withstand. As this effect is linked to the presence of oxygen in the material,126,127 hermetic packaging can mitigate this risk.
Particularly the horizontal configurations of MIM waveguides, Figs. 2(d) and 2(e) have been exploited in experiments to form the most compact, high-speed plasmonic modulators.39,40,128–132
The coupling of a photonic mode to a plasmonic mode needs special attention. Photonic–plasmonic conversion has been successfully solved by, e.g., deploying silicon bus waveguides,39,133,134 tapered silicon fiber-to-chip grating couplers,35 or via mode conversion from a dielectrically loaded long-range surface plasmon polariton (SPP).34
MIM waveguides [see Figs. 6(a) and 6(b)] rely on very small lithographic features (slot widths d ∼ 100 nm), which can be difficult to realize. This results in maximal aspect ratios of ∼5 (e.g., a 30-nm-wide gap in a 150-nm-high gold film). However, increasing the aspect ratio would be desirable. A large aspect ratio helps to increase the field interaction factor Γ and leakage to the substrate may be better suppressed.135 Furthermore, the modal intensity is reduced, potentially reducing the rate of photodegradation.
Three examples of plasmonic metal–insulator–metal waveguides. The plasmonic mode is color-coded, the external RF field EΩ is displayed by the field lines. The metals are 150 nm thick, and the plasmonic slots are d = 100 nm.15 (a) Horizontal metal slot on glass as used in plasmonic–organic hybrid devices. (b) Ferroelectric plasmonic waveguide with a barium titanate slab.40 (c) Vertical MIM slot waveguide with width w = 1000 nm.37
Three examples of plasmonic metal–insulator–metal waveguides. The plasmonic mode is color-coded, the external RF field EΩ is displayed by the field lines. The metals are 150 nm thick, and the plasmonic slots are d = 100 nm.15 (a) Horizontal metal slot on glass as used in plasmonic–organic hybrid devices. (b) Ferroelectric plasmonic waveguide with a barium titanate slab.40 (c) Vertical MIM slot waveguide with width w = 1000 nm.37
A way to fabricate MIM waveguides with higher aspect ratios of the gap can be realized by transitioning to a vertical layer stack,37,92,136–140 see Fig. 6(c). This allows plasmonic slots with a very large aspect ratio, which can help to mitigate detrimental boundary effects at the metal–insulator interfaces.24,139 Additionally, the top electrode might protect the EO material from detrimental environmental influences, such as oxygen. Furthermore, with horizontal layers, it is easier to obtain very smooth metal surfaces that form the plasmonic waveguide141 and offer lower propagation losses142 as these surfaces are not defined by etching. Finally, in a vertical stack, the slot thickness d can be defined by well-controllable layered deposition techniques, which increases fabrication tolerance and lowers cost in contrast to highest-resolution lithography methods.37
Figure 6 shows modal simulations of the different MIM modulator configurations from Fig. 2. The slot dimension is chosen to be d = 100 nm, the metal is 150 nm thick, and the upper electrode of the vertical configuration is w = 1000 nm wide. The optical mode is always strongly confined to the gap region and is well aligned with the RF field. From the simulations, one can also see that in the horizontal configurations [Figs. 6(a) and 6(b)] more of the optical field leaks into the cladding and substrate, leading to a slightly reduced field interaction factor Γ and slightly higher propagation losses α. This can also be seen from the plots in Fig. 7, which displays the dependence of the modulator properties on the electrode separation d.
Influence of the plasmonic gap width d on the performance metrics for horizontal slot (dashed) [see Fig. 6(a)] and vertical slot (solid) waveguides [see Fig. 6(c)]. The higher aspect ratio of the vertical slot reduces leakage into non-EO material, so that modulation efficiency and losses are superior.
Influence of the plasmonic gap width d on the performance metrics for horizontal slot (dashed) [see Fig. 6(a)] and vertical slot (solid) waveguides [see Fig. 6(c)]. The higher aspect ratio of the vertical slot reduces leakage into non-EO material, so that modulation efficiency and losses are superior.
Increasing the plasmonic gap width d has multiple implications. The field interaction factor and the voltage–length product deteriorate for wider gaps. Conversely, the losses are reduced. Whereas simulations reveal that a minimal gap width d leads to an optimal voltage–length–loss product VπLα, modulators with a ∼70 nm-wide-gap perform best in experiment.143 When comparing horizontal (dashed) and vertical (solid), the vertical configuration outperforms the horizontal in all simulated parameters: Its greater aspect ratio reduces leakage into the surrounding non-EO materials. Therefore, the stronger confinement Γ leads to a smaller voltage–length product, and to less energy guided in plasmonic edge modes, which tend to be more lossy, and more prone to scattering. Simulations show that a larger aspect ratio (determined by a large waveguide width in the vertical structure) improves the modulators’ performance characteristics (see Fig. 8). Note that the larger aspect ratio increases the capacitance of the vertical device, resulting in an EO bandwidth lower than that that of devices with a horizontal slot, but numerical 3D simulations suggest that the frequency response still is larger than 300 GHz.
Influence of the waveguide width w on the performance of the vertical slot waveguide with slot width d = 100 nm. As the confinement increases for wider waveguides and the propagation losses decrease, it is advantageous to select wider waveguides. Note that the device capacitance increases, and the waveguide is not single moded anymore.
Influence of the waveguide width w on the performance of the vertical slot waveguide with slot width d = 100 nm. As the confinement increases for wider waveguides and the propagation losses decrease, it is advantageous to select wider waveguides. Note that the device capacitance increases, and the waveguide is not single moded anymore.
The simulations shown above are in good agreement with experiment. For instance, working with a device having a plasmonic slot of width d = 130 nm, a gold thickness tAu = 150 nm, and , as in Ref. 125, simulations and Eq. (14) lead to Γ = 0.82, so that VπL ≈ 400 V µm with Eq. (10). Within a Mach–Zehnder modulator and if driven by RF frequencies, the simulations would suggest a VπL ≈ 100 V µm for the structure from the experiment. This is only slightly overestimating the measured value of 130 V µm.125
1. Frequency response
Plasmonic devices outperform the efficiency-bandwidth products of photonic devices. Due to their compactness, they show only low RF parasitics and offer small device capacitances. As a result, plasmonic modulators can be treated as a capacitive termination of their drive line, resulting in the full reflection of the impeding RF wave and an effective voltage doubling. This regime of operation holds to at least 500 GHz.62 Furthermore, the metal electrodes are in direct contact with the active section such that there is hardly any resistance. In consequence, the RC-constant is small and accordingly the frequency response of plasmonic organic hybrid modulators is high and flat from DC to at least 500 GHz.62 Furthermore, electro-optic modulation in plasmonic gaps at 1.2563 and 2.4 THz144 has been demonstrated. Based on measurements of R and C, the EO bandwidth of these modulators might be as high as 80 THz.128 The extremely large EO bandwidth also makes plasmonic MIM modulators an excellent choice for microwave photonics and radio over fiber applications.62,145
2. Advantages, challenges, and recent advances
Plasmonic waveguide modulators’ most unique feature is their electro-optic bandwidth exceeding 500 GHz,62 with operation up to 2.4 THz144—unmatched by any other modulator technology. On the other hand, plasmonic propagation losses are fundamental and pose an important challenge to the technology. Nevertheless, state-of-the-art plasmonic devices combined with foundry-quality silicon photonics can result in plasmonic MZMs with less than 12 dB fiber-to-fiber losses and 5 dB on-chip losses.146 More recently, plasmonic race-track modulators with fiber-to-fiber losses of 6.5 dB and on-chip plasmonic losses of 1.2 dB have been shown.147 Furthermore, a change from horizontal to vertical plasmonic waveguides can help to further minimize the impact of losses (see Fig. 7).37,136,137 The vertical geometry also facilitates the use of better plasmonic metals141 or performance-boosting interfacial layers.139
Generally, materials with a stronger electro-optic effect can help to keep devices short and losses low. The use of barium titanate has resulted in high-speed ferroelectric plasmonic modulators.40 Although the voltage–length product of first-generation phase shifter plasmonic ferroelectric modulators is ∼400 Vμm, it is estimated that this number can decrease by a factor of 10 to ∼40 Vμm,40 an efficiency where devices shorter than 5 µm could be deployed.
C. Hybrid photonic–plasmonic waveguides
Hybrid photonic–plasmonic waveguide modulators are a compromise between pure photonic and pure plasmonic modulators. As such, they have the potential to address the confinement-loss trade-off of plasmonic MIM modulators. Typically, these hybrid photonic–plasmonic structures comprise a thin dielectric layer incorporated between a transparent conductor with high refractive index, such as silicon and a metal [see Figs. 2(g)–2(i)]. While the propagation loss of these hybrid waveguides is lower than those of plasmonic geometries, their field confinement Γ (see field simulations in Fig. 9) is higher than those of purely photonic solutions. This makes hybrid photonic–plasmonic waveguides interesting for technological applications, such as lasers,148 frequency converters,149 or electro-optic switches150 and modulators.90,93,132 Indeed, the hybrid waveguide structure offers a route to easily functionalize well-established photonic platforms, such as silicon or silicon nitride, with the advantages of sub-wavelength confinement in plasmonics.
Cross section and modal simulation of a hybrid photonic–plasmonic modulator. (a) Simplified geometry of a plain hybrid photonic–plasmonic waveguide. (b) Lateral ground contacts have been added to provide a low-resistive path between ground and signal. (c) An additional partial etch helps to improve the confinement and field interaction factor Γ.
Cross section and modal simulation of a hybrid photonic–plasmonic modulator. (a) Simplified geometry of a plain hybrid photonic–plasmonic waveguide. (b) Lateral ground contacts have been added to provide a low-resistive path between ground and signal. (c) An additional partial etch helps to improve the confinement and field interaction factor Γ.
Figure 10 shows the field interaction factor Γ, the voltage–length–loss product VπLα, and the voltage–length product VπL of a hybrid photonic–plasmonic modulator as depicted in Fig. 9(b). The field interaction factor Γ in Fig. 10(a) is smaller than that of MIM modulators, because energy is transported not only in the slot but also in the silicon waveguide (see Fig. 9). Still, Γ can reach values close to unity, although values between ∼0.4 and ∼0.7 are more realistic. Figure 10(b) shows how the voltage–length product increases with the dielectric thickness (gap dimension d) —this is because the external field Eext drops with d−1. The same holds true for the voltage–length–loss product in Fig. 10(d), which is also optimal at the smallest gap dimension d even though losses are highest there [Fig. 10(c)]. As for plasmonic MIM devices, the experimental best case is likely to be determined by a trade-off between the theoretical optimum (very small d), fabrication constraints (requiring slightly larger d) , and potential surface effects at the interfaces of the EO active dielectric with the metal electrode and the silicon waveguide.143,151,152
Performance parameter of a hybrid geometry modulator shown for changing slot dimensions d. Solid lines: 340 nm-thick silicon; dashed lines: 220 nm silicon. (a) Change of the field confinement Γ. The thinner silicon layer consistently provides better confinement. However, this comes at the price of higher losses, resulting in a comparable VπLα for both silicon thickness. (a) For thicker silicon (solid lines), the field interaction Γ has an optimum when the dielectric is thick enough to confine the energy strongly, and thin enough so that the energy is not yet transported in the silicon waveguide. (b) The voltage–length product increases steadily, as the strength of the external modulating field Eext = Vextd−1 drops. (c) Losses decrease for larger d. (d) The voltage–length–loss product is best for small d (low VπL, but large α), or very large d (large VπL, but low α).
Performance parameter of a hybrid geometry modulator shown for changing slot dimensions d. Solid lines: 340 nm-thick silicon; dashed lines: 220 nm silicon. (a) Change of the field confinement Γ. The thinner silicon layer consistently provides better confinement. However, this comes at the price of higher losses, resulting in a comparable VπLα for both silicon thickness. (a) For thicker silicon (solid lines), the field interaction Γ has an optimum when the dielectric is thick enough to confine the energy strongly, and thin enough so that the energy is not yet transported in the silicon waveguide. (b) The voltage–length product increases steadily, as the strength of the external modulating field Eext = Vextd−1 drops. (c) Losses decrease for larger d. (d) The voltage–length–loss product is best for small d (low VπL, but large α), or very large d (large VπL, but low α).
When designing hybrid structures for practical use, additional points require consideration: First, ground electrodes need to be added laterally to the simplified geometry of Figs. 2(g) and 9(a), resulting in a structure as depicted in Fig. 9(b). This requires the conductive dielectric to be wider than the plasmonic metal strip, which comes at the expense of modal confinement. The confinement loss can be mitigated by introducing a partially etched or sub-wavelength-grated section between the waveguide core region and the lateral contacts [see Figs. 9(c) and 2(i), and particularly Fig. 11].
Detailed schematic of a hybrid modulator with lateral contacts and a partial etch between ground contacts and hybrid photonic-plasmonic section. The partial etch contributes to the lateral confinement of the plasmonic mode and hence to a high confinement factor Γ. The equivalent circuit diagram takes the contact resistance Rcontact and channel resistance Rchannel as well as modulator capacitance Cmodulator into account and assumes the active material to be a perfect insulator.
Detailed schematic of a hybrid modulator with lateral contacts and a partial etch between ground contacts and hybrid photonic-plasmonic section. The partial etch contributes to the lateral confinement of the plasmonic mode and hence to a high confinement factor Γ. The equivalent circuit diagram takes the contact resistance Rcontact and channel resistance Rchannel as well as modulator capacitance Cmodulator into account and assumes the active material to be a perfect insulator.
The partial etch can also help to avoid optical absorption losses from contact dopings or silicidation, which could be employed to reduce the metal–silicon contact resistance. On the other hand, the partial etch leads to a higher device resistance and hence reduces its electro-optic bandwidth. Further relations, which are subject to design trade-offs, are detailed in Table III. They can be optimized iteratively to fulfill the system requirements, which usually are given by insertion losses, bandwidth, and Vπ.
Design trade-offs inherent to hybrid plasmonic modulators.
Increasing parameter . | Acts positively on . | Acts negatively on . |
---|---|---|
Depth of partial etch hPE, or filling factor of sub-wavelength-grated contact | Resistance, field interaction | Bandwidth VπL |
Doping concentration | Bandwidth, α, VπLα | Resistance |
Waveguide width wwg | Capacitance | VπL, α, bandwidth |
Electrode separation wPE | Resistance | α, bandwidth |
Gap d | Bandwidth | Capacitance, field interaction, Eext |
Increasing parameter . | Acts positively on . | Acts negatively on . |
---|---|---|
Depth of partial etch hPE, or filling factor of sub-wavelength-grated contact | Resistance, field interaction | Bandwidth VπL |
Doping concentration | Bandwidth, α, VπLα | Resistance |
Waveguide width wwg | Capacitance | VπL, α, bandwidth |
Electrode separation wPE | Resistance | α, bandwidth |
Gap d | Bandwidth | Capacitance, field interaction, Eext |
To help understand the design complexity and the intertwined performance trade-offs, we analyze an exemplary structure based on a recently published geometry.44
The device is based on a 30-nm-thick BaTiO3 film on 220 nm silicon on insulator.44 In the waveguide area, the silicon is n-doped to a concentration of 1 · 10−18 cm−3. In this study, we consider an additional partial etch between the waveguide and the electrodes, which can increase Γ and VπL. The schematic is shown in Fig. 11. The waveguide is 200 nm wide; the gap dimension is d = 30 nm, the partial etch is 100 nm deep, and the distance between signal and ground electrodes is 200 nm. To be consistent with the simulations above, we have assumed BaTiO3’s refractive index to be nmat = 2.3, with .
Figure 12 displays the impact of waveguide width, doping concentration, depth of the partial etch, and distance between ground electrodes on the EO bandwidth (details on its calculation in Sec. III C 1) and the modulation efficiency in terms of VπLα (the influence of the gap dimension d was already elaborated in Fig. 10). Only n-type doping was considered here because p-type doping results in higher optical loss for the same achieved conductivity.153
Modulation efficiency VπLα (dark blue, left y-axis) and EO bandwidth fRC (bright blue, right y-axis) when changing (a) doping concentration, (b) partial etch depth, (c) waveguide width, and (d) electrode spacing. A doping concentration up to 1019 cm−3 increases the bandwidth without causing additional optical losses. In all other dimensions, an improvement of VπLα to small values goes together with a degradation of the cut-off frequency.
Modulation efficiency VπLα (dark blue, left y-axis) and EO bandwidth fRC (bright blue, right y-axis) when changing (a) doping concentration, (b) partial etch depth, (c) waveguide width, and (d) electrode spacing. A doping concentration up to 1019 cm−3 increases the bandwidth without causing additional optical losses. In all other dimensions, an improvement of VπLα to small values goes together with a degradation of the cut-off frequency.
1. Frequency response
The plasmonic nature of the device usually allows it to be sufficiently short to allow modeling as lumped elements. In Fig. 11, one can also see an equivalent circuit model, similar to typical models of plasmonic modulators.39 The model consists of a capacitor Cmod for the active region, resistors Rchannel for the conductive path through silicon, and contact resistors Rcontact between the ground contacts and silicon.
2. Advantages, challenges and recent advances
Hybrid plasmonic waveguide modulators can surpass photonic modulators in bandwidth (see Fig. 12) while maintaining low insertion loss compared to pure plasmonic modulators (compare Figs. 8 and 10). Using a vertical stack to define the active region brings larger fabrication tolerances37 and allows integration of epitaxial materials.44 Further investigation is needed to optimize the combination of conductive dielectric, nonlinear material and their electrical contact simultaneously140,143,151,156 and interface effects need to be considered, as bulk behavior is usually dissimilar to materials in contact with the metal.143,151
Recent exciting experiments demonstrate the potential of hybrid photonic–plasmonic devices. Phase modulation in a BaTiO3-based hybrid modulator has recently been demonstrated,44 as well as an organic EO material-based device relying on graphene instead of silicon for the electric path.93,157 A recent implementation of a plasmonically enhanced graphene organic hybrid electro-optic phase modulator of 10 µm length with low plasmonic losses of 2.5 dB has shown a bandwidth of 270 GHz. It has already been verified for high-speed on-off-keying data modulation at a line rate of 140 Gbit/s. More complex hybrid photonic–plasmonic layer stacks relying on ultra-thin metal layers promise even lower optical losses.47
IV. COMPARISON OF PHOTONIC, PLASMONIC, AND HYBRID GEOMETRIES
A qualitative comparison of photonic, plasmonic and hybrid plasmonic–photonic waveguide geometries in their key performance metrics is illustrated in Fig. 13. Photonic modulators are unmatched in their low losses, but lose modal confinement and strength of the modulating, external electric field. This results in an increased voltage–length product that requires long devices. These, at the same time, require travelling wave electrodes, which typically limit the available electro-optic bandwidth to <100 GHz.16,27,29,108,109,158 Recent advances on the structuring of the electrodes promise to bring that number up to 250 GHz for a 2 cm-long device.31
Illustrative comparison between photonic, plasmonic, and photonic–plasmonic hybrid modulators based on experimental key performance metrics. A larger radius stands for better performance. BW: bandwidth; VπL: voltage–length product; loss: device loss; α: loss per unit length; and Γ: field confinement.
Illustrative comparison between photonic, plasmonic, and photonic–plasmonic hybrid modulators based on experimental key performance metrics. A larger radius stands for better performance. BW: bandwidth; VπL: voltage–length product; loss: device loss; α: loss per unit length; and Γ: field confinement.
Plasmonic electro-optic modulators span the other extreme: Their voltage–length product is extremely small, and so is their size. This allows them to be driven as a lumped element, with electro-optic bandwidths exceeding 500 GHz and demonstrated operation up to 2.4 THz. The plasmonic mode confinement that enables these devices comes at the cost of fundamental propagation losses, currently at ∼0.5 dB/μm.35 However, as devices benefit from stronger electro-optic materials,67,159 they can become shorter and it is expected that fiber-to-fiber losses can drop: Once fiber-to-plasmonic coupling challenges are overcome, we will likely see losses of plasmonic devices below 5 dB. Most recently, plasmonic ring modulators featuring on-chip-losses with as little as 1.2 dB with a bandwidth in excess of 176 GHz have been demonstrated.147
Hybrid photonic–plasmonic modulator geometries represent a less explored field of ongoing research and their basic idea to give up some of the plasmonic confinement to reduce the device loss must still be proven in experiment. If successful, these hybrid devices could boast relatively low propagation losses together with a low voltage–length product and fill in the gap between plasmonic and photonic devices. Latest results with plasmonically enhanced graphene–organic hybrid phase modulator show that plasmonic hybrid devices with losses as low as 2.5 dB, bandwidths in excess of 270 GHz, and a length of 10 µm can be realized.93
Table IV lists numerical results for the geometries discussed in this work. For the sake of a fair comparison, all geometries are simulated with a material with nmat = 2.0 and reff = 125 pm/V, so that again . The results corroborate the qualitative findings.
Numerical comparison between photonic, plasmonic, and photonic–plasmonic hybrid waveguiding concepts. Simulations are carried out with nmat = 2 and (r33 = 125 pm/V). Dimensions for SOH: 150 nm slot width d, 220 nm Si thickness, 70 nm slab thickness, electrode distance. LNOI: 5 µm electrode distance. Plasmonic: 100 nm gap d, 150 nm gold thickness, in the vertical configuration waveguide width wwg: 1000 nm. Hybrid: 30 nm gap, 340 nm Si thickness, and 500 nm width. The EO bandwidth is estimated from literature or simulation result.
Geometry . | VπL . | VπLα (V dB) . | α . | Γ . | Bandwidth . |
---|---|---|---|---|---|
Photonic (SOH) | 0.07 V cm | 0.014 | 0.21 dB/cm | 0.23 | <100 GHz16,108,109 |
Photonic (LNOI) | 1.37 V cm | 0.3 | 0.2 dB/cm | 0.57 | 50–100 GHz29 |
250 GHz (sim.)31 | |||||
Plasmonic (POH), horizontal | 170 V μm | 85 | 0.49 dB/μm | 0.9 | >500 GHz62 |
>1 THz63 | |||||
Plasmonic (POH), vertical | 127 V μm | 55 | 0.43 dB/μm | 1.2 | >70 GHz37 |
300 GHz (sim.)a | |||||
Hybrid | 87 V μm | 25 | 0.13 dB/μm | 1.0 | ∼50–100 GHzb,93 |
Geometry . | VπL . | VπLα (V dB) . | α . | Γ . | Bandwidth . |
---|---|---|---|---|---|
Photonic (SOH) | 0.07 V cm | 0.014 | 0.21 dB/cm | 0.23 | <100 GHz16,108,109 |
Photonic (LNOI) | 1.37 V cm | 0.3 | 0.2 dB/cm | 0.57 | 50–100 GHz29 |
250 GHz (sim.)31 | |||||
Plasmonic (POH), horizontal | 170 V μm | 85 | 0.49 dB/μm | 0.9 | >500 GHz62 |
>1 THz63 | |||||
Plasmonic (POH), vertical | 127 V μm | 55 | 0.43 dB/μm | 1.2 | >70 GHz37 |
300 GHz (sim.)a | |||||
Hybrid | 87 V μm | 25 | 0.13 dB/μm | 1.0 | ∼50–100 GHzb,93 |
When taking RF lines into consideration, based on simulations.
Based on equivalent circuit calculations, similar as presented in Ref. 80.
The experimental situation is more intricate, with results listed in Table V. Whereas the experimental performance of LNOI devices is close to the expectations from simulation, deviations exist for SOH devices, where the losses in experiment exceed those expected from simulation. Part of this discrepancy might be explained by the importance of fabrication process and its effect on the device performance.
Comparison between photonic, plasmonic, and photonic–plasmonic hybrid modulators. Values have been experimentally extracted from Mach–Zehnder modulators in push–pull configuration, if not otherwise labeled.
Geometry . | VπL . | VπLα (V dB) . | L . | α . | Bandwidth . |
---|---|---|---|---|---|
Photonic (SOH) | 0.4 V mm16 | 1.016 | 280 µm16 | 2.5 dB/cm16 | ∼40 GHz16 |
∼100 GHz109 | |||||
Photonic (LNOI) | ∼2 V cm31 | <1.5 | ∼2 cm | <0.75 dB/cm31 | 100 GHz29 |
250 GHza,31 | |||||
Photonic (BaTiO3) | 0.23 V cmb | 1.3 | 2 mm | 5.8 dB/cm | ∼2 GHz (MZM)c |
∼20 GHz (racetrack)d | |||||
Plasmonic (POH), horizontal | 130 V μm | 75 | 15 µm | 0.6 dB/μm | >500 GHz62 |
>THz63 | |||||
Plasmonic (POH), verticale | 100 V μm (1550 nm), 30 V μm (1260 nm) | 50 | 11 µm | 0.5 dB/μm | >70 GHz |
15 | ∼300 GHzf | ||||
Hybrid | 350 V μm | 87 | 10 µm | 0.25 dB/μm | ∼50–100 GHzg,93 |
>270 GHzh,93 |
Geometry . | VπL . | VπLα (V dB) . | L . | α . | Bandwidth . |
---|---|---|---|---|---|
Photonic (SOH) | 0.4 V mm16 | 1.016 | 280 µm16 | 2.5 dB/cm16 | ∼40 GHz16 |
∼100 GHz109 | |||||
Photonic (LNOI) | ∼2 V cm31 | <1.5 | ∼2 cm | <0.75 dB/cm31 | 100 GHz29 |
250 GHza,31 | |||||
Photonic (BaTiO3) | 0.23 V cmb | 1.3 | 2 mm | 5.8 dB/cm | ∼2 GHz (MZM)c |
∼20 GHz (racetrack)d | |||||
Plasmonic (POH), horizontal | 130 V μm | 75 | 15 µm | 0.6 dB/μm | >500 GHz62 |
>THz63 | |||||
Plasmonic (POH), verticale | 100 V μm (1550 nm), 30 V μm (1260 nm) | 50 | 11 µm | 0.5 dB/μm | >70 GHz |
15 | ∼300 GHzf | ||||
Hybrid | 350 V μm | 87 | 10 µm | 0.25 dB/μm | ∼50–100 GHzg,93 |
>270 GHzh,93 |
With segmented electrode on quartz substrate, simulation only.
DC value, might overestimate the actual performance.12,40,60,117
Limited by velocity mismatch between RF and optical wave.
Limited by photon lifetime in the resonator.
Values for VπL from Ref. 37 have been measured for phase modulators, and are divided by a factor of 2 in this table, to match the measurements reported from Mach–Zehnder modulators in push–pull configuration.
When taking RF lines into consideration, based on simulations.
Based on a Si-OEO-Au stack and equivalent circuit calculations, similar as presented in Ref. 80.
Based on a Si-graphene-OEO-Au stack.
To conclude, integrated electro-optic modulators relying on the Pockels have seen great advances in technological readiness in recent years and can overcome the bandwidth limitations and nonlinearities of silicon–photonic plasma-dispersion modulators. Photonic, plasmonic, and photonic–plasmonic hybrid waveguide structures can all be employed to form modulators, each technology with its own drawbacks and benefits: Photonic silicon–organic hybrid thin-film lithium niobate modulators promise to be low loss and fast, but have a considerable footprint. They might be well-suited for classical optical communication tasks.28,160 Plasmonic metal–insulator–metal waveguides are extremely compact and exhibit unmatched electro-optical bandwidths but need clever engineering to reduce losses. Their unique properties may make them a good fit for microwave photonics,62,146,161 THz spectroscopy,144,162 or close integration with electronics for (short-haul) optical communications.43,147,163 Photonic–plasmonic hybrid structures allow to address a middle ground between the photonic and plasmonic extremes, but they require more research attention to find their ideal application.
The presented technologies and waveguide structures cover an extremely wide field of possibilities, and researchers as well as engineers will have great flexibility for choosing the best fit for the application at their hands.
SUPPLEMENTARY MATERIAL
The supplementary material comprises a derivation of the anisotropic refractive index change Δnij induced by the Pockels effect, a derivation of the field interaction factor Γ, and an explanation of the method used for the numerical studies of Γ, α, VπL, and VπL · α.
ACKNOWLEDGMENTS
The authors gratefully acknowledge partial support by ETH, aCryComm FETOPEN (Grant No. 899558), NEBULA EU-ICT (Grant No. 871658), PlasmoniAC EU-ICT (Grant No. 871391), plaCMOS EU-ICT (Project No. 980997), PLASILOR ERC (Grant No. 670478), FLEX-SCALE EU JU-SNS (Grant No. 101096909), and ALLEGRO EU CL4 (Grant No. 101092766). ETH would like to acknowledge support by the Binning Rohrer Nano Center (BRNC) cleanroom operated by IBM Rüschlikon and ETH Zurich, Switzerland.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Andreas Messner: Conceptualization (equal); Data curation (lead); Funding acquisition (supporting); Investigation (equal); Methodology (equal); Visualization (lead); Writing – original draft (lead); Writing – review & editing (lead). David Moor: Conceptualization (supporting); Funding acquisition (supporting); Investigation (equal); Methodology (equal); Visualization (supporting); Writing – original draft (equal); Writing – review & editing (equal). Daniel Chelladurai: Conceptualization (supporting); Investigation (equal); Methodology (equal); Visualization (supporting); Writing – original draft (equal); Writing – review & editing (equal). Roman Svoboda: Formal analysis (supporting); Investigation (supporting); Methodology (supporting); Writing – original draft (supporting). Jasmin Smajic: Funding acquisition (supporting); Project administration (supporting); Supervision (supporting); Writing – review & editing (supporting). Juerg Leuthold: Conceptualization (equal); Funding acquisition (lead); Project administration (lead); Investigation (equal); Methodology (equal); Supervision (lead); Visualization (supporting); Writing – original draft (equal); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available within the article.