Next Article in Journal
Characterizing Acoustic Behavior of Silicon Microchannels Separated by a Porous Wall
Previous Article in Journal
Morse Code Recognition Based on a Flexible Tactile Sensor with Carbon Nanotube/Polyurethane Sponge Material by the Long Short-Term Memory Model
Previous Article in Special Issue
Sidewall Corrugation-Modulated Phase-Apodized Silicon Grating Filter
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Towards High-Performance Pockels Effect-Based Modulators: Review and Projections

by
Yu Li
1,2,
Muhan Sun
1,
Ting Miao
1 and
Jianping Chen
1,2,*
1
State Key Laboratory of Advanced Optical Communication Systems and Networks, Department of Electronic Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
2
SJTU-Pinghu Institute of Intelligent Optoelectronics, Pinghu 314200, China
*
Author to whom correspondence should be addressed.
Micromachines 2024, 15(7), 865; https://doi.org/10.3390/mi15070865
Submission received: 3 June 2024 / Revised: 26 June 2024 / Accepted: 27 June 2024 / Published: 30 June 2024
(This article belongs to the Special Issue Silicon Photonic Devices and Integration)

Abstract

:
The ever-increasing demand for high-speed data transmission in telecommunications and data centers has driven the development of advanced on-chip integrated electro-optic modulators. Silicon modulators, constrained by the relatively weak carrier dispersion effect, face challenges in meeting the stringent requirements of next-generation photonic integrated circuits. Consequently, there has been a growing interest in Pockels effect-based electro-optic modulators, leveraging ferroelectric materials like LiNbO3, BaTiO3, PZT, and LaTiO3. Attributed to the large first-order electro-optic coefficient, researchers have delved into developing modulators with expansive bandwidth, low power consumption, compact size, and linear response. This paper reviews the working principles, fabrication techniques, integration schemes, and recent highlights in Pockels effect-based modulators.

1. Introduction

Thanks to the development of artificial intelligence, big data, media flow, and other applications, expanding data demands have led to an exponential growth in new hyperscale data centers [1,2,3], resulting in ever-increasing bandwidth requirements for optical transceiver modules. Nowadays, Microsoft, Google, et al., have widely adopted 100 G and partially updated 200 G optical transceiver modules in their data centers. In addition, 400 G [4,5,6] and 800 G [7,8] transceiver modules have already been prototyped and are being produced in small volumes by suppliers, such as Intel [8] and Marvell [3]. Electro-optic (EO) effect-based modulators are commonly employed for facilitating electric-to-optic conversion in an optical transmitter for its fast response time compared to other modulation mechanisms, such as the thermo-optic effect [9,10,11]. Specifically, silicon-based Mach–Zehnder modulators (MZMs) are widely adopted and commercialized in optical transceiver modules. However, the lattice structure of silicon is centrosymmetric and thus has no first-order EO effect (also named the Pockels effect [12]). The modulation in silicon is realized through second-order non-linear free-carrier dispersion (FCD) effect [13]. Due to the use of relatively inexpensive silicon materials, the application of the FCD effect earns the benefits of lower costs and can seamlessly integrate with existing CMOS processes, making it suitable for large-scale production. However, the FCD effect in silicon is relatively weak; typically, a carrier concentration variation of 1017 cm−3 could only induce a refractive index change of 10−4, and this significantly limits further improvement in bandwidth and modulation efficiency in silicon-based EO modulators for beyond 100 G/lane transmission. Up to now, the demonstrated silicon-based modulators with a bandwidth exceeding 60 GHz have involved either (i) co-designing MZM with radio frequency integrated circuit (RFIC) drivers [14] or peripheral resistor–capacitor equalization [15] to generate bandwidth gain; (ii) introducing the slow-light effect through phase-shift grating couplers which sacrificed the driving voltages [16,17]; or (iii) replacing MZM structures with compact microrings [18,19], which requires a complicated wavelength-locking scheme. These methodologies still fundamentally rely on FCD effects; hence, they are unable to completely address the challenges associated with the relationship between carriers and the refractive index.
In order to overcome this issue, ferroelectric-material-based modulators have been studied in recent years. Inspired by the traditional bulk lithium niobate (LiNbO3) modulators in fiber communications, integrated thin-film lithium niobate modulators based on the Pockels effect were initially investigated. With an EO coefficient of 30 pm/V [20,21], modulators with a bandwidth exceeding 100 GHz have been demonstrated [22,23,24,25]. This breakthrough has sparked research interests in Pockels effect-based modulators, leading to the successive exploration of other ferroelectric materials with a higher EO coefficient, including barium titanate (BaTiO3) [26,27,28], lithium tantalite (LaTiO3) [29], lead zirconate titanate (PZT) [30], et al. The primary advantage of the Pockels effect lies in its high electro-optic coefficient, which enables significant modulation depth at low voltages. This characteristic allows ferroelectric-material-based modulators to achieve efficient modulation under low power consumption conditions. Additionally, ferroelectric materials typically exhibit extremely fast response times, reaching picosecond or even sub-picosecond levels, making them suitable for ultra-high-speed optical communication and modulators.
Research on other materials such as phase change material (PCM) [31,32] and indium tin oxide (ITO) [33,34,35] has also been reported; however, they suffered from bandwidth limitations due to a long electro-optic response time and thus are not favored in high-speed modulation. Two-dimensional (2D) material-assisted modulation has also been developed for its ultra-fast response times and compact size [36,37,38]. Specifically, the graphene-assisted electro-absorption (EA)-based modulator has achieved a bandwidth of 39 GHz with a footprint of 27 μm2 [36]. However, limited by the weak graphene/dielectric combination and the limited quality of graphene, EA graphene modulators encounter difficulties in achieving high speed and high modulation efficiencies simultaneously. Two-dimensional material-based modulators also face challenges with material integration, stability, and scalability in manufacturing processes.
Considering that the next-generation modulators for photonic transceivers and high-speed computing emphasize large bandwidth, low power consumption, and compact size, this review focuses on Pockels effect-based modulators, which have the potential to achieve a wide bandwidth and high modulation efficiency, as well as compact size. In this review, Section 2 presents the fundamentals of the Pockels effect and the derivation of electro-optic coefficients on various ferroelectric material platforms. Section 3 provides a comprehensive overview of the state of the art in Pockels effect-based modulators. Finally, Section 4 provides conclusions and perspectives.

2. Fundamentals of Pockels Effect-Based Modulators

The Pockels effect, also known as the linear EO effect, is a phenomenon in non-centrosymmetric materials where the refractive index changes linearly in response to an applied electric field. This effect was first described by Friedrich Pockels in 1893 [39]. It provides a mechanism for the manipulation of light using electric fields with an ultra-fast response time. Based on a lattice structure and physical properties, typical materials that exhibit the Pockels effect can be classified into four categories: (i) crystal compounds in the form of ABO3, such as LiNbO3, BaTiO3, and LiTaO3, and other ferroelectric materials; (ii) compounds with a cubic or hexagonal lattice structure in the form of AB, such as gallium arsenide (GaAs), cadmium telluride (CdTe), and other III-V materials; (iii) the MH2XO4 crystal family (where M = K, Na, NH4+; X = As, P), such as potassium dihydrogen phosphate (KH2PO4), ammonium dihydrogen phosphate (NH4H2PO4), and other inorganic compounds; and (iv) other materials with non-centrosymmetric crystal structures.
For an anisotropic material, the surface of constant energy density in the field vector space is an ellipsoid in the principal axis coordinate system, as shown in Figure 1, leading to an associated principal refractive index in each (orthogonal) principal direction. The equation for the index ellipse can be expressed as follows:
1 n 2 1 x 2 + 1 n 2 2 y 2 + 1 n 2 3 z 2 + 2 1 n 2 4 y z + 2 1 n 2 5 x z + 2 1 n 2 6 x y = 1
where (1/n2)i is the optical indicatrix. Upon the employment of an electric field, the optical indicatrix changes as follows:
1 n 2 i = j r i j k E j
where E j is the electric field, and r i j k is the linear EO tensor. Given that an ellipse coordinate system is defined by three axes, the r i j k is a 3 × 3 × 3 matrix with 27 elements. Identical tensors arising from the physical symmetry can be removed to reduce the matrix complexity, leaving 18 independent elements that can be rearranged into a 3 × 6 matrix, as shown below:
r i j = r 11 r 12 r 13 r 21 r 22 r 23 r 31 r 32 r 33 r 41 r 42 r 43 r 51 r 52 r 53 r 51 r 62 r 63
For particular Pockels crystals with specific point group symmetry structures, the number of non-zero tensors could be further reduced.

2.1. Pockels Effect for Materials in Point Group 3 m

Point group 3 m [40] is one of the 32 crystallographic point groups, which is characterized by a three-fold rotational axis (120-degree rotation symmetry) along one direction and three mirror planes intersecting this axis. The “3” denotes the three-fold rotational symmetry, and the “m” represents the mirror planes. Using LiNbO3 as an example, with its rhombohedral lattice system and trigonal crystal structure, it belongs to point group 3 m. Here, r 12 = r 61 = − r 22 , r 23 = r 13 , r 42 = r 51 ; thus, the tensor matrix could be simplified in the following form:
r i j = 0 r 22 r 13 0 r 22 r 13 0 0 r 33 0 r 42 0 r 42 0 0 r 22 0 0
As shown in Table 1, LiNbO3 has a maximum EO coefficient of r 33 . Therefore, aligning the electric field E with the Z-axis, assuming light travels in the same direction for waveguide-based modulators, leads to an optimized effective EO coefficient. Then, the corresponding index ellipsoid in the presence of E z could be streamlined to the following:
1 n x 2 + r 13 E z x 2 + 1 n y 2 + r 13 E z y 2 + 1 n z 2 r 33 E z z 2 = 1
Notice that for the index ellipse of LiNbO3, the refractive index along the x-axis and y-axis is equal, typically denoted as ordinary index n o , and the refractive index along the z-axis is extraordinary index n e . By considering minor variations in the refractive index upon the applied electric field, the formulations for the refractive index along the three axes can be simplified as follows:
n x = n y = n o + Δ n x n o 1 2 n o 3 r 13 E z
n z = n e + Δ n z n e 1 2 n e 3 r 33 E z
where Δ n x and Δ n z are the refractive changes upon the applied electric field E z . For waveguide-based MZMs, the refractive index in Equations (6) and (7) needs to be supplanted by the waveguide effective index n e f f .
Upon a voltage V applied to the MZM waveguide phase shifter, the phase difference along the two arms of the LiNbO3 MZM is as follows:
= 2 π λ n e f f L
where L is the length of the phase shifter in MZM, and λ is the input wavelength. Furthermore, taking into consideration the overlapping factor Γ between the electric field and the optical mode in the waveguide, could be expressed using the effective EO coefficient r and E z as follows:
= Γ π V n e f f 3 r L d λ
where d is the distance between the parallel electrodes. The modulation efficiency of the MZM is reflected by the half-wave voltage–length product:
V π L = d λ n e f f 3 r Γ
For the LiNbO3-based MZM, r is related to r 33 . From the equation, we can see that for a given material, the way to increase the modulation efficiency is either by varying the n e f f through waveguide dimension optimization or by enhancing the overlapping factor through electrode design.

2.2. Pockels Effect for Materials in Point Group 4 mm

An alternative method to enhance the modulation efficiency is to use other ferroelectric materials, such as LiTaO3, which is in the same point group as LiNbO3. However, for materials with a lattice structure in a distinct point group, the effective EO tensor matrix is varied. For BaTiO3, above its Curie temperature of 120 °C, the lattice structure is in cubic phase with a point group of m3m. In this phase, the BaTiO3 enters paraelectric mode and is centrosymmetric. At room temperature, the crystal structure is in tetragonal phase with a point group of 4 mm. Thus, this review mainly focuses on the tetragonal phase. Specifically, point group 4 mm features a four-fold rotational axis along the principal axis with 90-degree rotational symmetry [48]. The first m represents mirror planes that are repeated by the rotational symmetry, and the second m denotes mirror planes that bisect the first set. The tensor matrix for the 4 mm point group could be streamlined in the following form:
r i j = 0 0 r 13 0 0 r 13 0 0 r 33 0 r 51 0 r 51 0 0 r 22 0 0
The corresponding index ellipse is then simplified as follows:
1 n o 2 + r 13 E z x 2 + 1 n o 2 + r 13 E z y 2 + 1 n e 2 + r 33 E z z 2 + r 51 E y 2 y z + r 51 E x 2 z x = 1
BaTiO3-based modulators could either work in the elongated axis, the c-axis (dominated by r 51 ), or in the other two axes (dominated by the combined effect of r 51 , r 13 , and r 33 ). In the c-axis configuration, the optic axis is perpendicular to the BaTiO3 thin film. The resulting waveguide n e f f change correlates with the square of E x , which requires a relatively large bias voltage and is not favored for low-power consumption applications.
In the a-axis scenario, the optic axis lies within the plane of the BaTiO3 thin film, allowing for the alignment of the waveguide either parallel to the optic axis or at a specified angle, as shown in Figure 2a. Meanwhile, only the TE mode can be modulated in the a-axis scenario. The equivalent refractive index n T E and EO coefficient r T E for the TE mode of the waveguide, which is tilted at a ϕ-degree to the optic axis, can then be expressed as follows:
n e f f = n T E = n z = n o n e n e 2 sin 2 ϕ + n o 2 cos 2 ϕ
r = r T E = r 33 cos 3 ϕ + ( r 13 + 2 r 51 ) sin 2 ϕ cos ϕ
Calculated using the EO coefficients in Table 1, the equivalent EO coefficient as a function of the tilted angle ϕ is shown in Figure 2b. When ϕ = 55°, the highest equivalent r T E = 624 pm/V is obtained with an equivalent n T E of 2.40. Employing the same analysis as in Section 2.1, the modulation efficiency for BaTiO3-based MZM can be calculated.
Another Pockels effect material showcased for EO modulation is PZT, an inorganic compound denoted by the chemical formula PbZrxTi1−xO3. The crystal structure of the traditional PZT is similar to BaTiO3, featuring a lower EO coefficient and a higher Curie temperature, as outlined in Table 1. Consequently, the aforementioned analysis is also applicable to the case of PZT.

3. Ferroelectric-Material-Based Modulators

3.1. LiNbO3-Based Modulators

Leveraging the relatively mature crystal fabrication industry and its traditional applications in fiber optics, LiNbO3 stands out as the most prevalent and extensively researched material for Pockels effect-based integrated modulators. Traditional bulk LiNbO3 EO modulators, based on proton exchange or titanium diffusion, have a low refractive index contrast between the waveguide and the cladding, leading to a large optical mode. This requires large spacing between the modulation electrodes to avoid optical crosstalk and signal degradation, which results in high driving voltages and large power consumption. To address these limitations, recent exploration of an integrated LiNbO3 modulator has employed new fabrication techniques, including hybrid integration and lithium niobate on insulator (LNOI), to realize high-index-contrast LiNbO3 waveguides.
The key step towards high-index-contrast LiNbO3 waveguides is to produce single-crystalline LiNbO3 thin films on low-index dielectric substrates. Over the years, numerous techniques, such as molecular beam epitaxy (MBE), pulsed laser deposition (PLD), and chemical vapor deposition (CVD), have been investigated for the growth of LiNbO3 thin films. However, none of these have demonstrated mature single-crystalline LiNbO3-thin-film growth capability on low-refractive-index substrates. Thus, nowadays, a more widely used technique is the wafer bonding technique, either by bonding the LiNbO3 thin film to a substrate, where the thin film is directly obtained from the bulk LiNbO3, or by bonding Si/SiN device layers to LiNbO3 substrates. Figure 3 summarizes the commonly used integration schemes of LiNbO3-based modulators.

3.1.1. LNOI-Based Modulators

The most adopted approach to obtaining thin-film LiNbO3 on low-refractive-index materials is using the smart-cut technique. This technique, analogous to silicon-on-insulator (SOI) fabrication [53], involves the He ion implantation of an LiNbO3 wafer, followed by flip-wafer bonding to a carrier wafer. Both the carrier wafer and the bonding wafer are oxidized to enable atomic bonding. Subsequently, the bonded wafer is annealed in order to activate micro-explosions of the implanted ions, causing the LiNbO3 wafer to split and leaving a thin film on the carrier wafer. This bonded wafer is then polished and annealed to repair ion-implantation-induced damages. The method was first developed back in the early 2000s at a die scale. Recently, 8-inch LNOI wafers have been commercialized, underscoring the potential for cost-effective LNOI-based modulators and revitalizing the development of LNOI-based devices and systems. Instead of atomic oxide bonding, some researchers [54,55] chose to use polymer adhesive, namely benzocyclobutene (BCB), to bond the wafers due to its low-temperature operation and versatility. However, the limited temperature tolerance of the bonded wafer prevents the employment of high-temperature annealing processes for repairing crystal lattice damages, resulting in high waveguide loss and a low EO coefficient.
One of the challenges for LNOI technology is to achieve effective mode confinement in waveguides. Lithium niobate is relatively chemically inert, making it difficult to precisely etch LiNbO3 waveguide profiles with low roughness. Various techniques, including both wet etching and dry etching, have been explored to reduce the waveguide propagation loss.
Wet etching is one of the techniques that were first developed for LiNbO3 waveguide formation. Back in 2007, H. Hu et al. [56] demonstrated a 6.5 μm width and 8 μm height waveguide with a TE transmission loss of 0.3 dB/cm. The etching solution was composed of hydrofluoric acid (HF) and nitric acid (HNO3), with the addition of ethanol for surface smoothness. Over time, researchers have focused on enhancing etching capabilities through solvent refinement. As of 2023, significant progress has been made by R. Zhuang et al. [57], with a propagation loss down to 0.04 dB/cm for high-quality factor micro-resonators. This wet etching method comprised a solvent of mixed H2O2 and NH4OH at an escalated temperature of 85 °C, followed by 2 h of annealing at 250 °C in order to repair the potential damage induced by ion slicing and wet etching.
Alternatively, leveraging the advantages of high precision, anisotropy, and smooth sidewalls, dry etching using inductively coupled plasma (ICP) reactive-ion etching [58,59,60] is extensively preferred. The fluorine-gas-based ICP etch has the benefit of a fast etching speed but results in the formation of the byproduct LiF, which accumulates on sidewalls and reduces the slope angle. To mitigate the problem, I. Krasnokutska et al. [59] mixed Ar ions with the etching gas, achieving a propagation loss of 0.4 dB/cm and an enhanced slope angle of 75°. Chlorine-based gases, such Cl2 and BCl3, are also used for etching LiNbO3. Compared to fluorine gases, chlorine-based etching produces the byproduct LiCl, which has a lower melting point and thus is able to achieve an improved sidewall angle. M. Bahadori et al. [61] reported fully etched LiNbO3 waveguides with a height of 560 nm, using a mixture of Cl2, BCl3, and Ar, achieving a sidewall angle of 83°. Additionally, C. Shen et al. [62] employed a SiO2 mask and a mixture of Ar, Cl2, and BCl3 as the etching gas, obtaining an etching sidewall angle of close to 80°. There is also research focused on Ar plasma-based physical etching. F. Kaufmann et al. [63] conducted ICP etching using Ar ions alone. When the chamber pressure was reduced from 5 mTorr to 1 mTorr under a 600 V DC bias, the sidewall angle increased from ~52° to 62°, and the optimized propagation loss was 1.55 dB/cm. Furthermore, G. Chen et al. [58] achieved a propagation loss of 0.15 dB/cm for LiNbO3 waveguides with a thin-film thickness of 400 nm and a ridge height of 200 nm.
Thanks to the development of LNOI wafers and etching technology, over the past five years, there have been significant advancements in the development of LNOI-based modulators. These devices have seen improvements in both bandwidth and modulation efficiencies and integration with other photonic components. However, there remains a trade-off between the bandwidth and modulation efficiency in LiNbO3 modulators due to the intrinsic EO coefficient of the material. Increasing the bandwidth requires minimizing the overlap between the optic and electric fields in order to reduce the capacitance, which consequently lowers the modulation efficiency.
In order to mitigate these constraints, F. Juneghani et al. [64] achieved an extrapolated bandwidth of 170 GHz with a modulation efficiency of 3.3 V·cm by positioning the optical waveguides asymmetrically relative to the metal electrodes and introducing a thin SiO2 dielectric buffer layer beneath the electrodes. As shown in Figure 4a, the asymmetrical waveguide design allowed a strong electric field near the signal electrode while maintaining the distance between the electrodes. The buffer layer mitigated the optical loss and reduced the RF effective index for velocity matching. G. Chen et al. [58] integrated a periodic capacitively loaded travelling-wave electrode to reduce the RF loss and an undercut in the silicon substrate to decrease the RF refractive index, attaining a bandwidth beyond 67 GHz and a tuning efficiency of 2.2 V·cm, as shown in Figure 4b. Alternately, N. Chen et al. [60] enhanced the overlap between the optic and electric fields through introducing isolation trenches around the waveguide to enlarge the refractive index contrast. They further improved the mode confinement through implementing high-refractive-index material as the waveguide cladding [65], achieving a modulation efficiency of 1.4 V·cm and a bandwidth of over 67 GHz, as shown in Figure 4c.
On the other hand, advanced modulation formats [23,66,67] have been employed to further increase the data transmission capability of the modulators. F. Yang et al. [67] achieved a 250 Gb/s data transmission upon six-level pulse amplitude modulation (PAM) using an LNOI-based modulator with a bandwidth of 110 GHz. M. Xu et al. [23] demonstrated a high-performance dual-polarization in-phase quadrature (IQ) modulator for coherent transmission, enabling a single-wavelength 1.96 Tb/s data transmission, as shown in Figure 5a. X. Wang et al. [66] demonstrated a dual-polarization IQ modulator supporting 1.6 Tb/s data transmission under 256 quadrature amplitude modulation (QAM), as shown in Figure 5b.

3.1.2. Modulators with Hybrid-Integrated LiNbO3 on Si/SiN Platforms

To avoid the complexity associated with etching LiNbO3 for LNOI-based modulators, an alternative approach involves bonding Si or SiN [68,69] device layers to LiNbO3 substrates. In this approach, LiNbO3 functions as the slab, and together with the bonded Si or SiN waveguide layers, hybrid Si/SiN-LiNbO3 ridge waveguides are created. Through the optimization of the Si/SiN waveguide dimensions and LiNbO3 film thickness, the majority of light can be confined in the LiNbO3 film. Additionally, the hybrid integration scheme presents an easy and scalable way to integrate modulators with other maturely developed passive components on the Si/SiN platform. Figure 6 depicts the recent development in modulators based on LiNbO3-Si/SiN hybrid integration schemes.
The bonding of LiNbO3 to Si or SiN layers can be realized at the wafer-to-wafer [70,71,72], die-to-wafer [50,68], or die-to-die [73,74] level. The mainstream bonding technique is hydrophilic bonding. The process begins with surface cleaning of both the LiNbO3 and Si/SiN substrates using solvents such as acetone and isopropanol, followed by a rinse in deionized water. Surface activation is then carried out, typically through oxygen plasma treatment, to enhance surface energy and promote strong bonding. In direct bonding, the activated surfaces are then brought into contact under pressure for uniform adhesion. Subsequently, thermal annealing is conducted to enhance the bond by promoting atomic interdiffusion. In 2023, M. Churaev et al. [72] successfully bonded a 4-inch LiNbO3 wafer on a 100 mm diameter silicon wafer with Si3N4 waveguides on top and SiO2 as the bonding interface. S. Ghosh et al. [71] demonstrated an Al2O3-assisted bonding procedure for 100 mm LiNbO3 wafers on 200 mm Si3N4-integrated photonic integrated circuit (PIC) wafers.
Leveraging the bonding technology, P. Weigel et al. [22] demonstrated a 5 mm long hybrid-integrated MZM with a bandwidth of 106 GHz and a VπL product of 6.7 V·cm, with 81% light confined in the LiNbO3 layer, achieving a large effective EO coefficient and 5% light in the Si waveguide. Furthermore, the same research group reduced the VπL product to 3.1 V·cm [73] through the implementation of slow-wave electrodes for improved velocity matching.
Figure 6. Recent developments in modulators based on LiNbO3-Si/SiN hybrid integration schemes. (a) Schematic of a LiNbO3-on-Si MZM with slow-wave electrodes [73]. (b) Schematic of a folded MZM based on LiNbO3-on-SiN platform [75]. (c) Schematic of an MZM with dual-tapers design on a BCB-assisted LiNbO3-on-SiN platform [68] (adapted from [68] with permission: Copyright © 2024 Wiley-VCH GmbH). (d) Schematic of a bias-drift-free BCB-assisted LiNbO3-on-Si MZM [76].
Figure 6. Recent developments in modulators based on LiNbO3-Si/SiN hybrid integration schemes. (a) Schematic of a LiNbO3-on-Si MZM with slow-wave electrodes [73]. (b) Schematic of a folded MZM based on LiNbO3-on-SiN platform [75]. (c) Schematic of an MZM with dual-tapers design on a BCB-assisted LiNbO3-on-SiN platform [68] (adapted from [68] with permission: Copyright © 2024 Wiley-VCH GmbH). (d) Schematic of a bias-drift-free BCB-assisted LiNbO3-on-Si MZM [76].
Micromachines 15 00865 g006
Besides LiNbO3-Si hybrid integration, the LiNbO3-on-SiN platform has garnered research interest due to its wide transmission window, low propagation loss, and cost-effectiveness. A. Ahmed et al. [77] demonstrated a hybrid LiNbO3-SiN MZM with a Vπ below 1 V by extending the EO interaction region to 2.4 cm and reducing the electrode gap between the ground and signal. Taking advantage of a Michelson interferometer (MI) with a doubled EO interaction length, X. Huang et al. [78] achieved exceptional modulation efficiency with a low VπL of 1.06 V·cm for a 0.6 mm long MI modulator (MIM). However, the counterpropagation of light in MIMs has limited the device bandwidth to 40 GHz. Alternately, S. Nelan et al. [75] increased the EO interaction length from 6 mm (physical) to 10 mm (effective) by employing a folded waveguide structure. A 180° bend was introduced to overcome the polarity inverse between the optic mode and electric field. This design realized a modulation efficiency of 4 V·cm and a bandwidth of 37.5 GHz.
On the contrary, there is also research on Si [51] or SiN [79] layers bonded to LiNbO3 substrates. P. Zhang et al. [79] demonstrated an EO MZM on a SiN-loaded LNOI platform with a coplanar waveguide design for push–pull modulation. The achieved bandwidth was 30 GHz, and the VπL was 2.24 V·cm.
In addition to hydrophilic bonding, BCB-assisted LiNbO3 hybrid-integrated modulators have also been investigated [25,54,76] due to their ease of fabrication and flexibility in device design. X. Liu et al. [54] proposed a 10 mm long modulator with an estimated bandwidth beyond 300 GHz and a modulation efficiency of 1.2 V·cm. The LiNbO3 thin film was bonded to a quartz substrate using BCB, with air-filled undercut regions to enhance the waveguide mode confinement. M. He et al. [25] designed a Si-LiNbO3 double-waveguide-layer structure through BCB adhesive die-to-wafer bonding and LiNbO3 dry etching, in order to mitigate the constraints of LNOI-only structures and maintain a great optical mode confinement at the same time. Under a single-drive push–pull operation, an EO bandwidth of 70 GHz and a modulation efficiency of 2.2 V·cm was achieved for a 3 mm length modulator.

3.1.3. Summary of LiNbO3-Based Modulators

In conclusion, LiNbO3-based modulators have undergone rapid development, with various integration methods and fabrication technologies emerging. In order to improve key performance metrics including bandwidth, modulation efficiency, and insertion loss, studies have focused on novel designs, such as electrode engineering, hybrid waveguide configurations, and increasing interaction length schemes. Table 2 summarizes the state of the art of LiNbO3-based modulators in the last five years.
Table 2. State of the art of LiNbO3-based modulators in the last five years.
Table 2. State of the art of LiNbO3-based modulators in the last five years.
YearSchemeStructureLength
(mm)
Vπ
(V)
VπL
(V·cm)
Bandwidth
(GHz)
Insertion Loss (dB)Ref.
2023LNOIMZM56.63.3170NA *[64]
2023LNOIMZM43.521.41>670.5[65]
2023LNOIMZM431.2>402.43[60]
2022LNOIMZM54.742.37>110NA[67]
2021LNOIMZM53.51.75>40NA[80]
2021LNOIMZM132.363.068602[81]
2021LNOIMZM41.60.64>3NA[82]
2020LNOIIQM131.92.4>481.8[83]
2023SiN + LNMZM74.33371[68]
2022Si + LNMZM5NA3.11101.8[73]
2022Si + LNMZM102.22.2>670.2[58]
2022SiN + LNMZM64437.5NA[75]
2021SiN + LNMZM7.82.82.1830NA[79]
2021SiN + LNMIMNA17.81.06>40NA[78]
2021TFLNMZM10NA1.2>300<1[54]
2020SiN + LNMZM240.8752.11NA5.4[77]
* NA: Not available.

3.2. BaTiO3-Based Modulators

Recently, BaTiO3 emerged as a highly promising material for EO modulators in integrated photonics due to its high EO coefficient. As listed in Table 1, the reported EO coefficient for BaTiO3 exceeds 800 pm/V, positioning it as a strong candidate for realizing high-performance modulators.

3.2.1. Heterogenous Growth of BaTiO3 Thin Films

The integration of BaTiO3 on PICs typically involves the monolithic growth of BaTiO3 thin films on substrates. The heterogeneous growth of BaTiO3 thin films on substrates has been developed through diverse techniques, including metal–organic chemical vapor deposition (MOCVD) [84], PLD [85,86,87], MBE [88,89], and RF sputtering [90,91]. The thin film is typically grown on magnesium oxide (MgO) substrate or silicon substrate.
Using MgO as a substrate for BaTiO3-thin-film growth has been well studied due to its similar lattice structure to BaTiO3 and its broad spectral transparency [26,92]. MgO with a cubic lattice structure has a larger lattice constant (4.21 Å) compared to tetragonal-phase BaTiO3, causing a lattice mismatch of 4–5% (depending on the axis) [93]. The resulting BaTiO3 thin film could be either c-oriented or a-oriented, depending on the deposition methodology, growth conditions, and buffer layers.
Silicon is another commonly used substrate, considering its widespread availability and compatibility with semiconductor manufacturing techniques [94,95,96,97]. However, the lattice constant of silicon is 5.43 Å, leading to a severe lattice mismatch with BaTiO3. To address this issue, a buffer layer of SrTiO3 is implemented, with a thickness of a few nanometers. The SrTiO3 buffer layer helps to alleviate the strain caused by the lattice mismatch between the silicon substrate and the BaTiO3 thin film, facilitating the growth of high-quality epitaxial films.
In addition, chemical solution deposition (CDS) using La2O2CO3 template film has also been developed for BaTiO3-thin-film growth. E. Picavet et al. [98] successfully fabricated a BaTiO3 film of 190 nm thickness. The thin layer featured a c-axis aligned in plane and a Pockels coefficient of 139 pm/V.

3.2.2. State of the Art of BaTiO3-Based Modulators

In this section, the state of the art of BaTiO3-based modulators is presented. Compared to that of LiNbO3 material, the development of BaTiO3-based modulators is still in an early stage. Modulators in various configurations and integration schemes have been explored. Figure 7 summarizes the typical integration schemes of BaTiO3-based modulators.
Notice that BaTiO3 also exhibits chemical inertness. In heterogeneous integration approaches, typically, waveguide formations need to rely on either depositing amorphous-Si or SiN device layers or bonding to additional PIC chips. The nanoscale gap between the silicon or MgO substrate and BaTiO3 influences the distribution of the optical field within the BaTiO3 layer, requiring precise device design to attain optimal modulation efficiency.
A. Posadas et al. [90] reported an a-axis BaTiO3-based MZM with an effective EO coefficient of 157 pm/V and a VπL of 0.42 V·cm. The BaTiO3 thin film was deposited on SOI wafers using RF sputtering, followed by a polycrystalline silicon-rich SiN layer deposition to form the hybrid waveguides. Instead of SiN, C. Xiong et al. [97] formed the hybrid waveguide using a deposited amorphous-Si layer, with crystalline BaTiO3 thin film directly epitaxial on the SOI, achieving an effective Pockels coefficient of 213 pm/V, as shown in Figure 8a. P. Girouard et al. [26] demonstrated a photonic crystal-based modulator with a device length of 1 mm. The measured EO bandwidth was over 40 GHz, and the VπL was ~0.66 V·cm.
Another method to create waveguide morphology is through an ion-milling dry-etch technique [91,100], as shown in Figure 8b,c. Z. Dong et al. [91] achieved an etching depth of 175 nm using Ar ions to physically bombard the BaTiO3 layer. With an RF-sputtered BaTiO3 thin film on the SOI wafer and fully etched BaTiO3 waveguides, an MZM with an effective EO coefficient of 89 pm/V was demonstrated, with a corresponding VπL of 2.3 V·cm.
In order to use crystalline Si waveguides for potential low propagation loss, F. Eltes et al. [102] proposed to flip-chip bond the BaTiO3 films that were grown on SOI substrate to another PIC chip. This process allowed the exposure of the Si device layer of SOI after removing the Si and SiO2 layers beneath and thus enabled waveguide patterning on the Si device layer. An MZM supporting a 25 Gb/s data transmission was demonstrated with a waveguide propagation loss of 3 dB/cm and a low VπL of 0.2 V·cm.
Alternatively, the heterogeneous-grown BaTiO3 thin films could be integrated with Si or SiN device layers using a bonding technique. Lumiphase, for instance, developed a hybrid bonding platform in 2022 [103] in order to attach BaTiO3 thin films on damascene SiN waveguides. Leveraging this platform, they reported a series of works on various modulator schemes. A microring modulator with a low propagation loss of 0.7 dB/cm and a VπL of below 0.56 V·cm was demonstrated to validate the potential of the platform. In 2023, they reported a BaTiO3-based MZM with a footprint of 0.4 mm × 1.6 mm [27]. The device featured an insertion loss of 2 dB and a VπL of 4.8 V·mm, with an operating data rate up to 200 Gb/s under linear receiver equalization. Their recent work also involved plasmonic modulators [101,104], as shown in Figure 8d. The light was coupled from SiN waveguides to the plasmonic section through two transition tapers, functioning to realize light transition between SiN-to-BaTiO3 and BaTiO3-to-plasmonic mode. A modulator with a VπL of below 0.1 V·mm, a total insertion loss of 20.3 dB, and a bandwidth of 110 GHz was demonstrated with measured eye diagrams up to 256 GBd. Table 3 summarizes the recent state of the art of BaTiO3-integrated modulators.

3.3. Pockels Effect Modulators Based on Other Ferroelectric Materials

At present, Pockels effect-based modulators predominantly rely on LiNbO3 and BaTiO3. However, there are other ferroelectric materials worth noticing that have already shown high-performance capability, such as LiTaO3 and PZT.
PZT material has been explored for high-performance modulators for its low cost, lower phase transition temperature, and high EO coefficient. In 2021, D. Ban et al. [108] demonstrated PZT-thin-film growth on sapphire substrates using chemical solution deposition, showcasing an EO coefficient of 133 pm/V. The resulting MZM achieved a VπL of 3.6 V at 1550 nm and a verified modulation performance within the 6–12 GHz signal frequency range. In 2022, they further presented a modulator with PZT thin films grown on SiO2/Si substrates, featuring an enhanced VπL of 1.4 V·cm, a propagation loss of 1.8 dB/cm, and a bandwidth of 12 GHz [109]. Recently, S. Yokoyama et al. [110] demonstrated an MZM supporting 200 Gb/s PAM4 transmission based on PZT and lead lanthanum zirconate titanate (PLZT)-on-insulator platforms. By adding lanthanum to PZT to form PLZT, the EO coefficient is increased. The PZT/PLZT ridge waveguide had a ridge height of 100 nm and width of 1.6 μm; the estimated effective EO coefficient was 86 pm/V and 198 pm/V for PZT and PLZT modulators, respectively. The corresponding VπL for the 4 mm long MZM was 1.3 V·cm and 0.88 V·cm at 1550 nm.
LiTaO3 on insulator (LTOI) has already been commercialized in 5G radio frequency devices, making it a mature platform for integration applications. Compared to LNOI, LTOI exhibits a similar EO coefficient but has an optical birefringence that is over one order of magnitude lower. This is specifically helpful towards realizing waveguide bends with a reduced radius. For a proof-of-concept demonstration, J. Shen et al. [29] reported a microring modulator on an LTOI platform, using a deposited amorphous-Si layer for waveguide profiling. The achieved wavelength tuning efficiency was 12.8 pm/V with a bandwidth of over 20 GHz. Earlier this year, C. Wang et al. [47] achieved an MZM with a VπL of 1.9 V·cm and a bandwidth of 40 GHz on a 4-inch LTOI wafer, as shown in Figure 9b,c. The wafer was manufactured using the smart-cut technique, followed by a stepper-based wafer-scale etching process to create low-loss waveguides.

3.4. Challenges and Comparison between Ferroelectric-Material-Based Platforms

Leveraging the Pockels effect, ferroelectric-material-based modulators exhibit the capability of achieving high modulation efficiency. Among these materials, BaTiO3 exhibits the highest EO coefficient, typically achieving a VπL below 0.5 pm/V for EO modulators. On the contrary, both LiNbO3 and LaTiO3 have relatively lower EO coefficients, requiring specific waveguides or electrode design to lower the VπL. Consequently, modulators based on these two materials typically require a phase-shifter length exceeding 5 mm. The optical birefringence of LaTiO3 is an order of magnitude lower than that of LiNbO3, enabling smaller waveguide bending radii and compact footprints.
In principle, ferroelectric-material-based devices are able to possess ultra-fast response times, reaching the picosecond or even sub-picosecond range. However, to date, only LiNbO3-based modulators have demonstrated bandwidths of over 100 GHz. One of the main limitations is the physical properties of fabricated thin films. There is a discrepancy between the actual EO coefficient and the refractive index of grown thin films compared to the values reported for bulk materials, leading to a mismatch between the optical and the microwave velocities that degrade the EO bandwidth, particularly for BaTiO3, which has a relatively high dielectric constant at RF frequencies.
Nonetheless, the high production costs and the challenges associated with scaling up to 8-inch wafer manufacturing remain significant barriers for the extensive application of ferroelectric-material-based platforms. Among the four platforms, only LNOI has achieved a commercial product at the 8-inch wafer scale, albeit at a relatively high cost. For LTOI and PZT, researchers have reported wafer-scale manufacturing. BaTiO3 has yet to achieve mature wafer-level thin-film manufacturing technology on a PIC-compatible platform. The high chemical inertness of LiNbO3 and BaTiO3 also contributes to increased costs related to etching processes in production.
In addition, limited by the large bandgap, ferroelectric materials are not ideal candidates for photodetection at communication wavelengths. Therefore, it is critical to investigate photodetection schemes on ferroelectric platforms. This is essential not only for receivers but also for the potential on-chip power monitoring of transmitters, such as bias point monitoring for MZMs and wavelength-locking for microring modulators.

4. Conclusions and Perspectives

In the past decade, outstanding performance and innovative functionalities have been reported for Pockels effect-based modulators. In this review, we present a comprehensive summary of the working principle, fabrication techniques, integration schemes, and latest breakthroughs for integrated modulators based on four ferroelectric materials, namely, LiNbO3, BaTiO3, PZT, and LaTiO3.
Leveraging their exceptional electro-optic properties, Pockels effect-based modulators represent a promising frontier for achieving potentially outstanding bandwidth, power efficiency, and compactness in next-generation communication systems. The matured LNOI platform has enabled the extensive development of LiNbO3-based modulators with demonstrated ultra-high bandwidth and diverse schemes to optimize the modulation efficiency and footprints. With a compelling EO coefficient, BaTiO3-based modulators have emerged in the last few years, achieving superior modulation efficiency. Additionally, PZT and LiTaO3 are also promising candidates due to their ease of fabrication and low optical birefringence, respectively.
Furthermore, the advantages of the Pockels effect-based modulators extend their application beyond data communication. There is already research evaluating the performance of Pockels effect-based modulators in neuromorphic photonic processors [111] and quantum photonics [112,113]. In addition, they also hold promising potential in diverse technology fields requiring high modulator functionalities, including sensing [114], satellite data links [115,116], and beyond. In conclusion, as research continues to push the boundaries of material science and manufacturing technology, ferroelectric-material-based modulators are poised to emerge as competitive candidates for future advanced photonics integration systems.

Author Contributions

Writing—original draft preparation, Y.L.; writing—review and editing, M.S. and T.M.; supervision, J.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (NSFC), grant number 62305212 and 62120106010.

Data Availability Statement

Data will be available based on request.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of this study; in the collection, analyses, or interpretation of data; in the writing of this manuscript; or in the decision to publish the results.

References

  1. Goda, K.; Jalali, B.; Lei, C.; Situ, G.; Westbrook, P. AI boosts photonics and vice versa. APL Photon. 2020, 5, 070401. [Google Scholar] [CrossRef]
  2. Alagappan, G.; Ong, J.R.; Yang, Z.; Ang, T.Y.L.; Zhao, W.; Jiang, Y.; Zhang, W.; Png, C.E. Leveraging AI in Photonics and Beyond. Photonics 2022, 9, 75. [Google Scholar] [CrossRef]
  3. Nagarajan, R.; Ding, L.; Coccioli, R.; Kato, M.; Tan, R.; Tumne, P.; Patterson, M.; Liu, L. 2.5D Heterogeneous Integration for Silicon Photonics Engines in Optical Transceivers. IEEE J. Sel. Top. Quantum Electron. 2023, 29, 1–10. [Google Scholar] [CrossRef]
  4. Pincemin, E.; Loussouarn, Y.; Sotomayor, A.; Losio, G.; McCarthy, M.; Nelson, L.; Malik, A.; Riggs, I.; Nielsen, T.; Williams, T.; et al. End-to-End Interoperable 400-GbE Optical Communications through 2-km 400GBASE-FR4, 8 × 100-km 400G-OpenROADM and 125-km 400-ZR Fiber Lines. J. Light. Technol. 2023, 41, 1250–1257. [Google Scholar] [CrossRef]
  5. Yu, H.; Doussiere, P.; Patel, D.; Lin, W.; Al-Hemyari, K.; Park, J.; Jan, C.; Herrick, R.; Hoshino, I.; Busselle, L.; et al. 400Gbps Fully Integrated DR4 Silicon Photonics Transmitter for Data Center Applications. In Proceedings of the 2020 Optical Fiber Communications Conference and Exhibition (OFC), San Diego, CA, USA, 8–12 March 2020; pp. 1–3. [Google Scholar]
  6. Pincemin, E.; Loussouarn, Y.; Sotomayor, A.; Losio, G.; McCarthy, M.; Nelson, L.; Malik, A.; Riggs, I.; Nielsen, T.; Williams, T.; et al. 927-km End-to-End Interoperable 400-GbEthernet Optical Communications through 2-km 400GBASE-FR4, 8 × 100-km 400G-OpenROADM and 125-km 400-ZR Fiber Lines. In Proceedings of the Optical Fiber Communication Conference (OFC) 2022, San Diego, CA, USA, 6–10 March 2022; p. Th4A.3. [Google Scholar]
  7. Isono, H. Recent standardization activities for client and networking optical transceivers and its future directions. In Next-Generation Optical Networks for Data Centers and Short-Reach Links IV; SPIE: Bellingham, WA, USA, 2017; Volume 10131. [Google Scholar]
  8. Fathololoumi, S.; Hui, D.; Jadhav, S.; Chen, J.; Nguyen, K.; Sakib, M.N.; Li, Z.; Mahalingam, H.; Amiralizadeh, S.; Tang, N.N.; et al. 1.6 Tbps Silicon Photonics Integrated Circuit and 800 Gbps Photonic Engine for Switch Co-Packaging Demonstration. J. Light. Technol. 2021, 39, 1155–1161. [Google Scholar] [CrossRef]
  9. Chung, S.; Nakai, M.; Hashemi, H. Low-power thermo-optic silicon modulator for large-scale photonic integrated systems. Opt. Express 2019, 27, 13430–13459. [Google Scholar] [CrossRef]
  10. Saha, N.; Brunetti, G.; Armenise, M.N.; Ciminelli, C. Tunable narrow band add-drop filter design based on apodized long period waveguide grating assisted co-directional coupler. Opt. Express 2022, 30, 28632–28646. [Google Scholar] [CrossRef]
  11. Harris, N.C.; Ma, Y.; Mower, J.; Baehr-Jones, T.; Englund, D.; Hochberg, M.; Galland, C. Efficient, compact and low loss thermo-optic phase shifter in silicon. Opt. Express 2014, 22, 10487–10493. [Google Scholar] [CrossRef]
  12. Pockels, F. Effects of the Electrical and Magnetic Fields. In Lehrbuch der Kristalloptik; Teubner: Leipzig, Germany, 1906; Part IV, Chapter III; pp. 492–510. [Google Scholar]
  13. Soref, R.; Bennett, B. Electrooptical effects in silicon. IEEE J. Quantum Electron. 1987, 23, 123–129. [Google Scholar] [CrossRef]
  14. Zhang, H.; Li, M.; Zhang, Y.; Zhang, D.; Liao, Q.; He, J.; Hu, S.; Zhang, B.; Wang, L.; Xiao, X.; et al. 800 Gbit/s transmission over 1 km single-mode fiber using a four-channel silicon photonic transmitter. Photon. Res. 2020, 8, 1776–1782. [Google Scholar] [CrossRef]
  15. Alam, M.S.; Li, X.; Jacques, M.; Xing, Z.; Samani, A.; El-Fiky, E.; Koh, P.-C.; Plant, D.V. Net 220 Gbps/λ IM/DD Transmssion in O-Band and C-Band With Silicon Photonic Traveling-Wave MZM. J. Light. Technol. 2021, 39, 4270–4278. [Google Scholar] [CrossRef]
  16. Jafari, O.; Shi, W.; LaRochelle, S. Efficiency-Speed Tradeoff in Slow-Light Silicon Photonic Modulators. IEEE J. Sel. Top. Quantum Electron. 2021, 27, 1–11. [Google Scholar] [CrossRef]
  17. Han, C.; Jin, M.; Tao, Y.; Shen, B.; Shu, H.; Wang, X. Ultra-compact silicon modulator with 110 GHz bandwidth. In Proceedings of the Optical Fiber Communication Conference (OFC), San Diego, CA, USA, 6–10 March 2022; p. Th4C.5. [Google Scholar]
  18. Ran, S.; Zhou, G.; Li, Y.; Lu, L.; Tang, N.; Liu, F.; Chen, J.; Zhou, L. Micro-Ring Modulators With Integrated Inductor to Mitigate Bandwidth and Extinction Ratio Trade-Off. IEEE Photon. Technol. Lett. 2024, 36, 231–234. [Google Scholar] [CrossRef]
  19. Chan, D.W.U.; Wu, X.; Zhang, Z.; Lu, C.; Lau, A.P.T.; Tsang, H.K. Ultra-Wide Free-Spectral-Range Silicon Microring Modulator for High Capacity WDM. J. Light. Technol. 2022, 40, 7848–7855. [Google Scholar] [CrossRef]
  20. Weis, R.S.; Gaylord, T.K. Lithium niobate: Summary of physical properties and crystal structure. Appl. Phys. A 1985, 37, 191–203. [Google Scholar] [CrossRef]
  21. Akazawa, H.; Fukuda, H. Epitaxial ZnO/LiNbO3/ZnO stacked layer waveguide for application to thin-film Pockels sensors. AIP Adv. 2015, 5, 057163. [Google Scholar] [CrossRef]
  22. Weigel, P.O.; Zhao, J.; Fang, K.; Al-Rubaye, H.; Trotter, D.; Hood, D.; Mudrick, J.; Dallo, C.; Pomerene, A.T.; Starbuck, A.L.; et al. Bonded thin film lithium niobate modulator on a silicon photonics platform exceeding 100 GHz 3-dB electrical modulation bandwidth. Opt. Express 2018, 26, 23728–23739. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, M.; Zhu, Y.; Pittalà, F.; Tang, J.; He, M.; Ng, W.C.; Wang, J.; Ruan, Z.; Tang, X.; Kuschnerov, M.; et al. Dual-polarization thin-film lithium niobate in-phase quadrature modulators for terabit-per-second transmission. Optica 2022, 9, 61–62. [Google Scholar] [CrossRef]
  24. Kharel, P.; Reimer, C.; Luke, K.; He, L.; Zhang, M. Breaking voltage–bandwidth limits in integrated lithium niobate modulators using micro-structured electrodes. Optica 2021, 8, 357–363. [Google Scholar] [CrossRef]
  25. He, M.; Xu, M.; Ren, Y.; Jian, J.; Ruan, Z.; Xu, Y.; Gao, S.; Sun, S.; Wen, X.; Zhou, L.; et al. High-performance hybrid silicon and lithium niobate Mach–Zehnder modulators for 100 Gbit s−1 and beyond. Nat. Photon. 2019, 13, 359–364. [Google Scholar] [CrossRef]
  26. Girouard, P.; Chen, P.; Jeong, Y.K.; Liu, Z.; Ho, S.T.; Wessels, B.W. X2 Modulator with 40-GHz Modulation Utilizing BaTiO3 Photonic Crystal Waveguides. IEEE J. Quantum Electron. 2017, 53, 1–10. [Google Scholar] [CrossRef]
  27. Eltes, F.; Li, W.; Berikaa, E.; Alam, M.S.; Bernal, S.; Minkenberg, C.; Plant, D.V.; Abel, S. Thin-film BTO-based modulators enabling 200 Gb/s data rates with sub 1 Vpp drive signal. In Proceedings of the Optical Fiber Communication Conference (OFC), San Diego, CA, USA, 5–9 March 2023; p. Th4A.2. [Google Scholar]
  28. Chelladurai, D.; Kohli, M.; Horst, Y.; Eppenberger, M.; Kulmer, L.; Blatter, T.; Winiger, J.; Moor, D.; Messner, A.; Convertino, C.; et al. Electro-Optic Frequency Response of Thin-Film Barium Titanate (BTO) from 20 to 270 GHz. In Proceedings of the European Conference on Optical Communication (ECOC), Basel, Switzerland, 18–22 September 2022; p. We5.72. [Google Scholar]
  29. Shen, J.; Zhang, Y.; Feng, C.; Xu, Z.; Zhang, L.; Su, Y. Hybrid lithium tantalite-silicon integrated photonics platform for electro-optic modulation. Opt. Lett. 2023, 48, 6176–6179. [Google Scholar] [CrossRef] [PubMed]
  30. Alexander, K.; George, J.P.; Verbist, J.; Neyts, K.; Kuyken, B.; Van Thourhout, D.; Beeckman, J. Nanophotonic Pockels modulators on a silicon nitride platform. Nat. Commun. 2018, 9, 3444. [Google Scholar] [CrossRef] [PubMed]
  31. Liang, H.; Soref, R.; Mu, J.; Majumdar, A.; Li, X.; Huang, W.P. Simulations of Silicon-on-Insulator Channel-Waveguide Electrooptical 2 × 2 Switches and 1 × 1 Modulators Using a Ge2Sb2Te5 Self-Holding Layer. J. Light. Technol. 2015, 33, 1805–1813. [Google Scholar] [CrossRef]
  32. Delaney, M.; Zeimpekis, I.; Lawson, D.; Hewak, D.W.; Muskens, O.L. A New Family of Ultralow Loss Reversible Phase-Change Materials for Photonic Integrated Circuits: Sb2S3 and Sb2Se3. Adv. Funct. Mater. 2020, 30, 2002447. [Google Scholar] [CrossRef]
  33. Wood, M.G.; Campione, S.; Parameswaran, S.; Luk, T.S.; Wendt, J.R.; Serkland, D.K.; Keeler, G.A. Gigahertz speed operation of epsilon-near-zero silicon photonic modulators. Optica 2018, 5, 233–236. [Google Scholar] [CrossRef]
  34. Hsu, W.-C.; Nujhat, N.; Kupp, B.; Conley, J.F.; Rong, H.; Kumar, R.; Wang, A.X. Sub-volt high-speed silicon MOSCAP microring modulator driven by high-mobility conductive oxide. Nat. Commun. 2024, 15, 826. [Google Scholar] [CrossRef] [PubMed]
  35. Jaffray, W.; Saha, S.; Shalaev, V.M.; Boltasseva, A.; Ferrera, M. Transparent conducting oxides: From all-dielectric plasmonics to a new paradigm in integrated photonics. Adv. Opt. Photon. 2022, 14, 148–208. [Google Scholar] [CrossRef]
  36. Agarwal, H.; Terrés, B.; Orsini, L.; Montanaro, A.; Sorianello, V.; Pantouvaki, M.; Watanabe, K.; Taniguchi, T.; Thourhout, D.V.; Romagnoli, M.; et al. 2D-3D integration of hexagonal boron nitride and a high-κ dielectric for ultrafast graphene-based electro-absorption modulators. Nat. Commun. 2021, 12, 1070. [Google Scholar] [CrossRef]
  37. Jin, M.; Wei, Z.; Meng, Y.; Shu, H.; Tao, Y.; Bai, B.; Wang, X. Silicon-Based Graphene Electro-Optical Modulators. Photonics 2022, 9, 82. [Google Scholar] [CrossRef]
  38. Liu, M.; Yin, X.; Ulin-Avila, E.; Geng, B.; Zentgraf, T.; Ju, L.; Wang, F.; Zhang, X. A graphene-based broadband optical modulator. Nature 2011, 474, 64–67. [Google Scholar] [CrossRef] [PubMed]
  39. Kaminow, I.P.; Turner, E.H. Linear Electrooptical Materials. In An Introduction to Electrooptic Devices; Kaminow, I.P., Ed.; Academic Press: Cambridge, MA, USA, 1974; pp. 108–120. [Google Scholar]
  40. Hahn, T.; Klapper, H.; Müller, U.; Aroyo, M. Point Groups and Crystal Classes. In International Tables for Crystallography; International Union of Crystallography: Chester, UK, 2016; Volume A, pp. 720–776. Available online: https://onlinelibrary.wiley.com/iucr/itc/Ac/ch3o2v0001/ch3o2.pdf (accessed on 1 June 2024).
  41. Servizzi, A.J.; Boyd, J.T.; Sriram, S.; Kingsley, S.A. Extraordinary-mode refractive-index change produced by the linear electro-optic effect in LiNbO3 and reverse-poled LiNbO3. Appl. Opt. 1995, 34, 4248–4255. [Google Scholar] [CrossRef]
  42. Smolenskii, G.A.; Krainik, N.N.; Khuchua, N.P.; Zhdanova, V.V.; Mylnikova, I.E. The Curie Temperature of LiNbO3. Phys. Status Solidi 1966, 13, 309–314. [Google Scholar] [CrossRef]
  43. Castera, P.; Tulli, D.; Gutierrez, A.M.; Sanchis, P. Influence of BaTiO3 ferroelectric orientation for electro-optic modulation on silicon. Opt. Express 2015, 23, 15332–15342. [Google Scholar] [CrossRef] [PubMed]
  44. Gubinyi, Z.; Batur, C.; Sayir, A.; Dynys, F. Electrical properties of PZT piezoelectric ceramic at high temperatures. J. Electroceramics 2008, 20, 95–105. [Google Scholar] [CrossRef]
  45. Kondo, M.; Sato, K.; Ishii, M.; Wakiya, N.; Shinozaki, K.; Kurihara, K. Electrooptic Properties of Lead Zirconate Titanate Films Prepared on Silicon Substrate. Jpn. J. Appl. Phys. 2006, 45, 7516. [Google Scholar] [CrossRef]
  46. Zhu, M.; Du, Z.; Jing, L.; Yoong Tok, A.I.; Tong Teo, E.H. Optical and electro-optic anisotropy of epitaxial PZT thin films. Appl. Phys. Lett. 2015, 107, 031907. [Google Scholar] [CrossRef]
  47. Wang, C.; Li, Z.; Riemensberger, J.; Lihachev, G.; Churaev, M.; Kao, W.; Ji, X.; Zhang, J.; Blesin, T.; Davydova, A.; et al. Lithium tantalate photonic integrated circuits for volume manufacturing. Nature 2024, 629, 784–790. [Google Scholar] [CrossRef]
  48. Hefferan, K.; O’Brien, J. Earth Materials; John Wiley & Sons: Hoboken, NJ, USA, 2022. [Google Scholar]
  49. Yu, L.; Shang, J.; Luo, K.; Lin, Q.; Chen, H.; Qiu, W.; Guan, H.; Lu, H. Design of High-Speed Mid-Infrared Electro-Optic Modulator Based on Thin Film Lithium Niobate. IEEE Photon. J. 2022, 14, 1–6. [Google Scholar] [CrossRef]
  50. Weigel, P.O.; Valdez, F.; Zhao, J.; Li, H.; Mookherjea, S. Design of high-bandwidth, low-voltage and low-loss hybrid lithium niobate electro-optic modulators. J. Phys. Photon. 2021, 3, 012001. [Google Scholar] [CrossRef]
  51. Wang, J.; Xu, S.; Chen, J.; Zou, W. A Heterogeneous Silicon on Lithium Niobate Modulator for Ultra-Compact and High-Performance Photonic Integrated Circuits. IEEE Photon. J. 2021, 13, 4900112. [Google Scholar] [CrossRef]
  52. Xu, M.; Chen, W.; He, M.; Wen, X.; Ruan, Z.; Xu, J.; Chen, L.; Liu, L.; Yu, S.; Cai, X. Michelson interferometer modulator based on hybrid silicon and lithium niobate platform. APL Photon. 2019, 4, 100802. [Google Scholar] [CrossRef]
  53. Bruel, M.; Aspar, B.; Charlet, B.; Maleville, C.; Poumeyrol, T.; Soubie, A.; Auberton-Herve, A.J.; Lamure, J.M.; Barge, T.; Metral, F.; et al. “Smart cut”: A promising new SOI material technology. In Proceedings of the IEEE International SOI Conference Proceedings, Tucson, AZ, USA, 3–5 October 1995; pp. 178–179. [Google Scholar]
  54. Liu, X.; Xiong, B.; Sun, C.; Wang, J.; Hao, Z.; Wang, L.; Han, Y.; Li, H.; Luo, Y. Sub-terahertz bandwidth capactively-loaded thin-film lithium niobate electro-optic modulators based on an undercut structure. Opt. Express 2021, 29, 41798–41807. [Google Scholar] [CrossRef]
  55. Guarino, A.; Poberaj, G.; Rezzonico, D.; Degl’Innocenti, R.; Günter, P. Electro–optically tunable microring resonators in lithium niobate. Nat. Photon. 2007, 1, 407–410. [Google Scholar] [CrossRef]
  56. Hu, H.; Ricken, R.; Sohler, W.; Wehrspohn, R.B. Lithium Niobate Ridge Waveguides Fabricated by Wet Etching. IEEE Photon. Technol. Lett. 2007, 19, 417–419. [Google Scholar] [CrossRef]
  57. Zhuang, R.; He, J.; Qi, Y.; Li, Y. High-Q Thin-Film Lithium Niobate Microrings Fabricated with Wet Etching. Adv. Mater. 2023, 35, 2208113. [Google Scholar] [CrossRef] [PubMed]
  58. Chen, G.; Chen, K.; Gan, R.; Ruan, Z.; Wang, Z.; Huang, P.; Lu, C.; Lau, A.P.T.; Dai, D.; Guo, C.; et al. High performance thin-film lithium niobate modulator on a silicon substrate using periodic capacitively loaded traveling-wave electrode. APL Photon. 2022, 7, 026103. [Google Scholar] [CrossRef]
  59. Krasnokutska, I.; Tambasco, J.-L.J.; Li, X.; Peruzzo, A. Ultra-low loss photonic circuits in lithium niobate on insulator. Opt. Express 2018, 26, 897–904. [Google Scholar] [CrossRef]
  60. Chen, N.; Yu, Y.; Lou, K.; Mi, Q.; Chu, T. High-efficiency thin-film lithium niobate modulator with highly confined optical modes. Opt. Lett. 2023, 48, 1602–1605. [Google Scholar] [CrossRef] [PubMed]
  61. Bahadori, M.; Yang, Y.; Goddard, L.L.; Gong, S. High performance fully etched isotropic microring resonators in thin-film lithium niobate on insulator platform. Opt. Express 2019, 27, 22025–22039. [Google Scholar] [CrossRef]
  62. Shen, C.; Wang, C.; Zhu, Y.; Wu, J.; Chen, Y.; Li, Z.; Huang, K.; Zhao, X.; Song, S.; Zhang, J.; et al. A comparative study of dry-etching nanophotonic devices on a LiNbO3-on-insulator material platform. In Proceedings of the Optics Frontier: Optics Young Scientist Summit, Ningbo, China, 4–7 December 2020. [Google Scholar]
  63. Kaufmann, F.; Finco, G.; Maeder, A.; Grange, R. Redeposition-free inductively-coupled plasma etching of lithium niobate for integrated photonics. Nanophotonics 2023, 12, 1601–1611. [Google Scholar] [CrossRef]
  64. Arab Juneghani, F.; Gholipour Vazimali, M.; Zhao, J.; Chen, X.; Le, S.T.; Chen, H.; Ordouie, E.; Fontaine, N.K.; Fathpour, S. Thin-Film Lithium Niobate Optical Modulators with an Extrapolated Bandwidth of 170 GHz. Adv. Photon. Res. 2023, 4, 2200216. [Google Scholar] [CrossRef]
  65. Chen, N.; Lou, K.; Yu, Y.; He, X.; Chu, T. High-Efficiency Electro-Optic Modulator on Thin-Film Lithium Niobate with High-Permittivity Cladding. Laser Photon. Rev. 2023, 17, 2200927. [Google Scholar] [CrossRef]
  66. Wang, X.; Shang, C.; Pan, A.; Cheng, X.; Gui, T.; Yuan, S.; Gui, C.; Zheng, K.; Zhang, P.; Song, X.; et al. Thin-film lithium niobate dual-polarization IQ modulator on a silicon substrate for single-carrier 1.6 Tb/s transmission. APL Photon. 2022, 7, 076101. [Google Scholar] [CrossRef]
  67. Yang, F.; Fang, X.; Chen, X.; Zhu, L.; Zhang, F.; Chen, Z.; Li, Y. Monolithic thin film lithium niobate electro-optic modulator with over 110 GHz bandwidth. Chin. Opt. Lett. 2022, 20, 022502. [Google Scholar] [CrossRef]
  68. Ruan, Z.; Chen, K.; Wang, Z.; Fan, X.; Gan, R.; Qi, L.; Xie, Y.; Guo, C.; Yang, Z.; Cui, N.; et al. High-Performance Electro-Optic Modulator on Silicon Nitride Platform with Heterogeneous Integration of Lithium Niobate. Laser Photon. Rev. 2023, 17, 2200327. [Google Scholar] [CrossRef]
  69. Chang, L.; Pfeiffer, M.H.P.; Volet, N.; Zervas, M.; Peters, J.D.; Manganelli, C.L.; Stanton, E.J.; Li, Y.; Kippenberg, T.J.; Bowers, J.E. Heterogeneous integration of lithium niobate and silicon nitride waveguides for wafer-scale photonic integrated circuits on silicon. Opt. Lett. 2017, 42, 803–806. [Google Scholar] [CrossRef] [PubMed]
  70. Huang, R.; Tang, M.; Kan, W.; Xu, H.; Wu, K.; Wang, Z.; Li, H. Single-crystalline LiNbO3 integrated onto Si-based substrates via Ar plasma-activated low-temperature direct bonding. J. Phys. D Appl. Phys. 2024, 57, 015102. [Google Scholar] [CrossRef]
  71. Ghosh, S.; Yegnanarayanan, S.; Kharas, D.; Ricci, M.; Plant, J.J.; Juodawlkis, P.W. Wafer-scale heterogeneous integration of thin film lithium niobate on silicon-nitride photonic integrated circuits with low loss bonding interfaces. Opt. Express 2023, 31, 12005–12015. [Google Scholar] [CrossRef] [PubMed]
  72. Churaev, M.; Wang, R.N.; Riedhauser, A.; Snigirev, V.; Blésin, T.; Möhl, C.; Anderson, M.H.; Siddharth, A.; Popoff, Y.; Drechsler, U.; et al. A heterogeneously integrated lithium niobate-on-silicon nitride photonic platform. Nat. Commun. 2023, 14, 3499. [Google Scholar] [CrossRef]
  73. Valdez, F.; Mere, V.; Wang, X.; Boynton, N.; Friedmann, T.A.; Arterburn, S.; Dallo, C.; Pomerene, A.T.; Starbuck, A.L.; Trotter, D.C.; et al. 110 GHz, 110 mW hybrid silicon-lithium niobate Mach-Zehnder modulator. Sci. Rep. 2022, 12, 18611. [Google Scholar] [CrossRef]
  74. Boynton, N.; Cai, H.; Gehl, M.; Arterburn, S.; Dallo, C.; Pomerene, A.; Starbuck, A.; Hood, D.; Trotter, D.C.; Friedmann, T.; et al. A heterogeneously integrated silicon photonic/lithium niobate travelling wave electro-optic modulator. Opt. Express 2020, 28, 1868–1884. [Google Scholar] [CrossRef]
  75. Nelan, S.; Mercante, A.; Hurley, C.; Shi, S.; Yao, P.; Shopp, B.; Prather, D.W. Compact thin film lithium niobate folded intensity modulator using a waveguide crossing. Opt. Express 2022, 30, 9193–9207. [Google Scholar] [CrossRef]
  76. Sun, S.; He, M.; Xu, M.; Gao, S.; Chen, Z.; Zhang, X.; Ruan, Z.; Wu, X.; Zhou, L.; Liu, L.; et al. Bias-drift-free Mach–Zehnder modulators based on a heterogeneous silicon and lithium niobate platform. Photon. Res. 2020, 8, 1958–1963. [Google Scholar] [CrossRef]
  77. Ahmed, A.N.R.; Nelan, S.; Shi, S.; Yao, P.; Mercante, A.; Prather, D.W. Subvolt electro-optical modulator on thin-film lithium niobate and silicon nitride hybrid platform. Opt. Lett. 2020, 45, 1112–1115. [Google Scholar] [CrossRef] [PubMed]
  78. Huang, X.; Liu, Y.; Li, Z.; Guan, H.; Wei, Q.; Tan, M.; Li, Z. 40 GHz high-efficiency Michelson interferometer modulator on a silicon-rich nitride and thin-film lithium niobate hybrid platform. Opt. Lett. 2021, 46, 2811–2814. [Google Scholar] [CrossRef] [PubMed]
  79. Zhang, P.; Huang, H.; Jiang, Y.; Han, X.; Xiao, H.; Frigg, A.; Nguyen, T.G.; Boes, A.; Ren, G.; Su, Y.; et al. High-speed electro-optic modulator based on silicon nitride loaded lithium niobate on an insulator platform. Opt. Lett. 2021, 46, 5986–5989. [Google Scholar] [CrossRef] [PubMed]
  80. Liu, Y.; Li, H.; Liu, J.; Tan, S.; Lu, Q.; Guo, W. Low Vpi; thin-film lithium niobate modulator fabricated with photolithography. Opt. Express 2021, 29, 6320–6329. [Google Scholar] [CrossRef] [PubMed]
  81. Ying, P.; Tan, H.; Zhang, J.; He, M.; Xu, M.; Liu, X.; Ge, R.; Zhu, Y.; Liu, C.; Cai, X. Low-loss edge-coupling thin-film lithium niobate modulator with an efficient phase shifter. Opt. Lett. 2021, 46, 1478–1481. [Google Scholar] [CrossRef]
  82. Jin, M.; Chen, J.; Sua, Y.; Kumar, P.; Huang, Y. Efficient electro-optical modulation on thin-film lithium niobate. Opt. Lett. 2021, 46, 1884–1887. [Google Scholar] [CrossRef]
  83. Xu, M.; He, M.; Zhang, H.; Jian, J.; Pan, Y.; Liu, X.; Chen, L.; Meng, X.; Chen, H.; Li, Z.; et al. High-performance coherent optical modulators based on thin-film lithium niobate platform. Nat. Commun. 2020, 11, 3911. [Google Scholar] [CrossRef] [PubMed]
  84. Towner, D.J.; Ni, J.; Marks, T.J.; Wessels, B.W. Effects of two-stage deposition on the structure and properties of heteroepitaxial BaTiO3 thin films. J. Cryst. Growth 2003, 255, 107–113. [Google Scholar] [CrossRef]
  85. Winiger, J.; Keller, K.; Moor, D.; Baumann, M.; Kim, D.; Chelladurai, D.; Kohli, M.; Blatter, T.; Dénervaud, E.; Fedoryshyn, Y.; et al. PLD Epitaxial Thin-Film BaTiO3 on MgO—Dielectric and Electro-Optic Properties. Adv. Mater. Interfaces 2024, 11, 2300665. [Google Scholar] [CrossRef]
  86. Buchal, C.; Beckers, L.; Eckau, A.; Schubert, J.; Zander, W. Epitaxial BaTiO3 thin films on MgO. Mater. Sci. Eng. B 1998, 56, 234–238. [Google Scholar] [CrossRef]
  87. Niu, G.; Yin, S.; Saint-Girons, G.; Gautier, B.; Lecoeur, P.; Pillard, V.; Hollinger, G.; Vilquin, B. Epitaxy of BaTiO3 thin film on Si(001) using a SrTiO3 buffer layer for non-volatile memory application. Microelectron. Eng. 2011, 88, 1232–1235. [Google Scholar] [CrossRef]
  88. Abel, S.; Stöferle, T.; Marchiori, C.; Rossel, C.; Rossell, M.D.; Erni, R.; Caimi, D.; Sousa, M.; Chelnokov, A.; Offrein, B.J. A strong electro-optically active lead-free ferroelectric integrated on silicon. Nat. Commun. 2013, 4, 1671. [Google Scholar] [CrossRef] [PubMed]
  89. Kormondy, K.J.; Popoff, Y.; Sousa, M.; Eltes, F.; Caimi, D.; Rossell, M.D.; Fiebig, M.; Hoffmann, P.; Marchiori, C.; Reinke, M. Microstructure and ferroelectricity of BaTiO3 thin films on Si for integrated photonics. Nanotechnology 2017, 28, 075706. [Google Scholar] [CrossRef] [PubMed]
  90. Posadas, A.B.; Park, H.; Reynaud, M.; Cao, W.; Reynolds, J.D.; Guo, W.; Jeyaselvan, V.; Beskin, I.; Mashanovich, G.Z.; Warner, J.H.; et al. Thick BaTiO3 Epitaxial Films Integrated on Si by RF Sputtering for Electro-Optic Modulators in Si Photonics. ACS Appl. Mater. Interfaces 2021, 13, 51230–51244. [Google Scholar] [CrossRef]
  91. Dong, Z.; Raju, A.; Posadas, A.B.; Reynaud, M.; Demkov, A.A.; Wasserman, D.M. Monolithic Barium Titanate Modulators on Silicon-on-Insulator Substrates. ACS Photon. 2023, 10, 4367–4376. [Google Scholar] [CrossRef]
  92. Tang, P.; Towner, D.J.; Hamano, T.; Meier, A.L.; Wessels, B.W. Electrooptic modulation up to 40 GHz in a barium titanate thin film waveguide modulator. Opt. Express 2004, 12, 5962–5967. [Google Scholar] [CrossRef]
  93. Touloukian, Y.; Kirby, R.; Taylor, E.; Lee, T. Thermophysical properties of matter-the tprc data series. In Thermal Expansion-Nonmetallic Solids; Thermophysical and Electronic Properties Information Analysis Center: Lafayette, IN, USA, 1977; Volume 13. [Google Scholar]
  94. Abel, S.; Eltes, F.; Ortmann, J.E.; Messner, A.; Castera, P.; Wagner, T.; Urbonas, D.; Rosa, A.; Gutierrez, A.M.; Tulli, D.; et al. Large Pockels effect in micro- and nanostructured barium titanate integrated on silicon. Nat. Mater. 2019, 18, 42–47. [Google Scholar] [CrossRef] [PubMed]
  95. Pernice, W.H.P.; Xiong, C.; Walker, F.J.; Tang, H.X. Design of a Silicon Integrated Electro-Optic Modulator Using Ferroelectric BaTiO3 Films. IEEE Photon. Technol. Lett. 2014, 26, 1344–1347. [Google Scholar] [CrossRef]
  96. Rosa, A.; Tulli, D.; Castera, P.; Gutierrez, A.M.; Griol, A.; Baquero, M.; Vilquin, B.; Eltes, F.; Abel, S.; Fompeyrine, J.; et al. Barium titanate (BaTiO3) RF characterization for application in electro-optic modulators. Opt. Mater. Express 2017, 7, 4328–4336. [Google Scholar] [CrossRef]
  97. Xiong, C.; Pernice, W.H.P.; Ngai, J.H.; Reiner, J.W.; Kumah, D.; Walker, F.J.; Ahn, C.H.; Tang, H.X. Active Silicon Integrated Nanophotonics: Ferroelectric BaTiO3 Devices. Nano Lett. 2014, 14, 1419–1425. [Google Scholar] [CrossRef] [PubMed]
  98. Picavet, E.; Lievens, E.; De Geest, K.; Rijckaert, H.; Fernandez, E.G.; Bikondoa, O.; Solano, E.; Paturi, P.; Singh, N.; Pannier, T.; et al. Integration of Solution-Processed BaTiO3 Thin Films with High Pockels Coefficient on Photonic Platforms. Adv. Funct. Mater. 2024, 2403024. [Google Scholar] [CrossRef]
  99. Messner, A.; Eltes, F.; Ma, P.; Abel, S.; Baeuerle, B.; Josten, A.; Heni, W.; Caimi, D.; Fompeyrine, J.; Leuthold, J. Plasmonic Ferroelectric Modulators. J. Light. Technol. 2019, 37, 281–290. [Google Scholar] [CrossRef]
  100. Sun, N.; Sun, D.; Wu, D.; Guo, Y.; Fan, Y.; Zou, F.; Pu, M.; Luo, X. Lowering The Bit-Energy of Electro-Optic Modulators Via Polarization-Phase Modulation in Thin-Film BaTiO3 Ferroelectric Crystal Waveguide. Laser Photon. Rev. 2024, 18, 2300937. [Google Scholar] [CrossRef]
  101. Kohli, M.; Chelladurai, D.; Messner, A.; Horst, Y.; Moor, D.; Winiger, J.; Blatter, T.; Buriakova, T.; Convertino, C.; Eltes, F.; et al. Plasmonic Ferroelectric Modulator Monolithically Integrated on SiN for 216 GBd Data Transmission. J. Light. Technol. 2023, 41, 3825–3831. [Google Scholar] [CrossRef]
  102. Eltes, F.; Mai, C.; Caimi, D.; Kroh, M.; Popoff, Y.; Winzer, G.; Petousi, D.; Lischke, S.; Ortmann, J.E.; Czornomaz, L.; et al. A BaTiO3-Based Electro-Optic Pockels Modulator Monolithically Integrated on an Advanced Silicon Photonics Platform. J. Light. Technol. 2019, 37, 1456–1462. [Google Scholar] [CrossRef]
  103. Riedhauser, A.; Snigirev, V.; Lihachev, G.; Wang, R.N.; Churaev, M.; Caimi, D.; Möhl, C.; Convertino, C.; Eltes, F.; Kippenberg, T.J.; et al. Low-loss heterogeneously integrated barium-titanate-on-silicon-nitride photonics. In Proceedings of the 2023 Conference on Lasers and Electro-Optics (CLEO), San Jose, CA, USA, 7–12 May 2023. [Google Scholar]
  104. Kohli, M.; Chelladurai, D.; Kulmer, L.; Keller, K.; Horst, Y.; Blatter, T.; Winiger, J.; Moor, D.; Buriakova, T.; Zervas, M.; et al. 256 GBd Barium-Titanate-on-SiN Mach-Zehnder Modulator. In Proceedings of the Optical Fiber Communication Conference (OFC), San Diego, CA, USA, 24–28 March 2024. [Google Scholar]
  105. Han, H.; Wang, J.; Wang, Z.; Liu, C.; Xiang, B. Integrated barium titanate electro-optic modulators operating at CMOS-compatible voltage. Appl. Opt. 2023, 62, 6053–6059. [Google Scholar] [CrossRef]
  106. Chatzitheocharis, D.; Ketzaki, D.; Patsamanis, G.; Vyrsokinos, K. BTO-based O-band Sub-Volt CMOS Compatible Plasmonic Racetrack Modulator on Si3N4. In Proceedings of the Frontiers in Optics + Laser Science, Rochester, NY, USA, 16–20 October 2022. [Google Scholar]
  107. Eltes, F.; Villarreal-Garcia, G.E.; Caimi, D.; Siegwart, H.; Gentile, A.A.; Hart, A.; Stark, P.; Marshall, G.D.; Thompson, M.G.; Barreto, J.; et al. An integrated optical modulator operating at cryogenic temperatures. Nat. Mater. 2020, 19, 1164–1168. [Google Scholar] [CrossRef] [PubMed]
  108. Ban, D.; Liu, G.; Yu, H.; Sun, X.; Deng, N.; Qiu, F. High electro-optic coefficient lead zirconate titanate films toward low-power and compact modulators. Opt. Mater. Express 2021, 11, 1733–1741. [Google Scholar] [CrossRef]
  109. Ban, D.; Liu, G.; Yu, H.; Wu, Y.; Qiu, F. Low Driving Voltage and Low Optical Loss Electro-Optic Modulators Based on Lead Zirconate Titanate Thin Film on Silicon Substrate. J. Light. Technol. 2022, 40, 2939–2943. [Google Scholar] [CrossRef]
  110. Yokoyama, S.; Mao, J.; Uemura, F.; Sato, H.; Lu, G.W. 200 Gbit/s Transmitter Based on a Spin-on Ferroelectric Waveguide Modulator. In Proceedings of the Optical Fiber Communications Conference (OFC), San Diego, CA, USA, 5–9 March 2023; pp. 1–3. [Google Scholar]
  111. Marinis, L.D.; Andriolli, N.; Contestabile, G. Analysis of Integration Technologies for High-Speed Analog Neuromorphic Photonics. IEEE J. Sel. Top. Quantum Electron. 2023, 29, 6101409. [Google Scholar] [CrossRef]
  112. Sund, P.I.; Lomonte, E.; Paesani, S.; Wang, Y.; Carolan, J.; Bart, N.; Wieck, A.D.; Ludwig, A.; Midolo, L.; Pernice, W.H.P.; et al. High-speed thin-film lithium niobate quantum processor driven by a solid-state quantum emitter. Sci. Adv. 2023, 9, eadg7268. [Google Scholar] [CrossRef] [PubMed]
  113. Saravi, S.; Pertsch, T.; Setzpfandt, F. Lithium Niobate on Insulator: An Emerging Platform for Integrated Quantum Photonics. Adv. Opt. Mater. 2021, 9, 2100789. [Google Scholar] [CrossRef]
  114. Xu, S.; Ren, Z.; Dong, B.; Zhou, J.; Liu, W.; Lee, C. Mid-Infrared Silicon-on-Lithium-Niobate Electro-Optic Modulators Toward Integrated Spectroscopic Sensing Systems. Adv. Opt. Mater. 2023, 11, 2202228. [Google Scholar] [CrossRef]
  115. Armenise, M.N.; di Toma, A.; Brunetti, G.; Saha, N.; Ciminelli, C. Flexible Photonic Integrated Circuits: A New Paradigm to Process Data on-Board Satellites. In Proceedings of the 2023 23rd International Conference on Transparent Optical Networks (ICTON), Bucharest, Romania, 2–6 July 2023. [Google Scholar]
  116. Tzintzarov, G.N.; Rao, S.G.; Cressler, J.D. Integrated silicon photonics for enabling next-generation space systems. Photonics 2021, 8, 131. [Google Scholar] [CrossRef]
Figure 1. Index ellipsoid of the EO crystal without applied electric field. k is the light beam wave vector.
Figure 1. Index ellipsoid of the EO crystal without applied electric field. k is the light beam wave vector.
Micromachines 15 00865 g001
Figure 2. (a) Schematic of effective EO coefficient module in an a-axis-oriented BaTiO3 platform. (b) Effective EO coefficient as a function of tilted angle ϕ.
Figure 2. (a) Schematic of effective EO coefficient module in an a-axis-oriented BaTiO3 platform. (b) Effective EO coefficient as a function of tilted angle ϕ.
Micromachines 15 00865 g002
Figure 3. Integration schemes of LiNbO3-based modulators. (a) LNOI with etched LiNbO3 waveguides [49]; (b) LiNbO3 thin film bonded to SOI wafers [50]; (c) Si-loaded LiNbO3 thin films [51]; (d) LiNbO3 thin film bonded to SOI wafers using BCB [52].
Figure 3. Integration schemes of LiNbO3-based modulators. (a) LNOI with etched LiNbO3 waveguides [49]; (b) LiNbO3 thin film bonded to SOI wafers [50]; (c) Si-loaded LiNbO3 thin films [51]; (d) LiNbO3 thin film bonded to SOI wafers using BCB [52].
Micromachines 15 00865 g003
Figure 4. Recent developments in LNOI-based modulators. (a) Schematic of an asymmetric waveguides-to-RF pads design [64]; (b) top-view schematic of the periodic capacitively loaded traveling-wave electrode design [58]; (c) schematic of the waveguide cross-section with isolated trenches and a high-refractive-index cladding [65] (adapted from [65] with permission: Copyright © 2024 Wiley-VCH GmbH).
Figure 4. Recent developments in LNOI-based modulators. (a) Schematic of an asymmetric waveguides-to-RF pads design [64]; (b) top-view schematic of the periodic capacitively loaded traveling-wave electrode design [58]; (c) schematic of the waveguide cross-section with isolated trenches and a high-refractive-index cladding [65] (adapted from [65] with permission: Copyright © 2024 Wiley-VCH GmbH).
Micromachines 15 00865 g004
Figure 5. (a) A 110-GHz bandwidth dual-polarization IQ modulator supporting 1.96 Tb/s data transmission [23]. (b) A dual-polarization IQ modulator supporting 1.6 Tb/s data transmission under 256 QAM [66].
Figure 5. (a) A 110-GHz bandwidth dual-polarization IQ modulator supporting 1.96 Tb/s data transmission [23]. (b) A dual-polarization IQ modulator supporting 1.6 Tb/s data transmission under 256 QAM [66].
Micromachines 15 00865 g005
Figure 7. Typical integration schemes of BaTiO3-based modulators. (a) Cross-sectional schematic of a BaTiO3 thin film grown on a MgO substrate, with a deposited SiN layer to form hybrid waveguides [92]. (b) A BaTiO3 thin film grown on an SOI wafer through a SiTrO3 buffer layer, with a deposited amorphous-Si layer to form hybrid waveguides [97] (adapted from [97] with permission: Copyright © 2024, American Chemical Society). (c) A BaTiO3 thin film grown on an SOI wafer, followed by Ar ion milling to form waveguides [91] (adapted from [91] with permission: Copyright © 2024, American Chemical Society). (d) A BaTiO3 thin film bonded to a PIC chip [99].
Figure 7. Typical integration schemes of BaTiO3-based modulators. (a) Cross-sectional schematic of a BaTiO3 thin film grown on a MgO substrate, with a deposited SiN layer to form hybrid waveguides [92]. (b) A BaTiO3 thin film grown on an SOI wafer through a SiTrO3 buffer layer, with a deposited amorphous-Si layer to form hybrid waveguides [97] (adapted from [97] with permission: Copyright © 2024, American Chemical Society). (c) A BaTiO3 thin film grown on an SOI wafer, followed by Ar ion milling to form waveguides [91] (adapted from [91] with permission: Copyright © 2024, American Chemical Society). (d) A BaTiO3 thin film bonded to a PIC chip [99].
Micromachines 15 00865 g007
Figure 8. (a) A c-axis-oriented microring modulator with a bandwidth of over 24 GHz [98] (adapted from [98] with permission: Copyright © 2024 Wiley-VCH GmbH). (b) Schematic of an MZM based on etched BaTiO3 waveguides with effective EO coefficient of 89 pm/V [91] (adapted from [91] with permission: Copyright © 2024, American Chemical Society). (c) A c-axis-oriented etched BaTiO3-based phase-polarization modulation scheme [100] (adapted from [100] with permission: Copyright © 2024 Wiley-VCH GmbH). (d) A plasmonic modulator based on the BaTiO3-SiN bonding technique [101].
Figure 8. (a) A c-axis-oriented microring modulator with a bandwidth of over 24 GHz [98] (adapted from [98] with permission: Copyright © 2024 Wiley-VCH GmbH). (b) Schematic of an MZM based on etched BaTiO3 waveguides with effective EO coefficient of 89 pm/V [91] (adapted from [91] with permission: Copyright © 2024, American Chemical Society). (c) A c-axis-oriented etched BaTiO3-based phase-polarization modulation scheme [100] (adapted from [100] with permission: Copyright © 2024 Wiley-VCH GmbH). (d) A plasmonic modulator based on the BaTiO3-SiN bonding technique [101].
Micromachines 15 00865 g008
Figure 9. (a) Microscope image of a PZT-based microring modulator [30]. (b) Photograph of a fabricated 4-inch LTOI wafer [47]. (c) Microscope image of a fabricated MZM on the LTOI platform [47].
Figure 9. (a) Microscope image of a PZT-based microring modulator [30]. (b) Photograph of a fabricated 4-inch LTOI wafer [47]. (c) Microscope image of a fabricated MZM on the LTOI platform [47].
Micromachines 15 00865 g009
Table 1. Physical properties of ferroelectric materials.
Table 1. Physical properties of ferroelectric materials.
MaterialPoint GroupEO Coefficient (pm/V)Refractive
Index
Curie Temp.
(°C)
Reference
LiNbO33 m r 13 = 9.6
r 22 = 6.8
r 33 = 30.9
r 42 = 32.6
n o = 2.286
n e = 2.2
1140[41,42]
BaTiO34 mm r 13 = 8
r 33 = 28
r 51 = 800
n o = 2.444
n e = 2.383
120[43]
PZT4 mm r c ( 001 ) = 270.2
r c ( 011 ) = 198.2
r c ( 111 ) = 125.3
n o = 2.453
n e = 2.458
340[44,45,46]
LiTaO33 m r 33 = 30.5 n o = 2.119
n e = 2.123
610–700[47]
Table 3. State of the art of BaTiO3-integrated modulators.
Table 3. State of the art of BaTiO3-integrated modulators.
YearStructureVπL
(V·cm)
Bandwidth
(GHz)
Insertion Loss
(dB)
Ref.
2024MRM1.88124NA *[98]
2023MZM2.32NANA[91]
2023MZM0.48NA2[27]
2023Plasmonic0.0144>7020.5[101]
2023MZM0.48262NA[105]
2022PlasmonicNA66.433[106]
2021MZM0.421NANA[90]
2020MRMNA30NA[107]
2019MZM0.232NA[102]
2017Photonic crystal0.664012[26]
* NA: Not available.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, Y.; Sun, M.; Miao, T.; Chen, J. Towards High-Performance Pockels Effect-Based Modulators: Review and Projections. Micromachines 2024, 15, 865. https://doi.org/10.3390/mi15070865

AMA Style

Li Y, Sun M, Miao T, Chen J. Towards High-Performance Pockels Effect-Based Modulators: Review and Projections. Micromachines. 2024; 15(7):865. https://doi.org/10.3390/mi15070865

Chicago/Turabian Style

Li, Yu, Muhan Sun, Ting Miao, and Jianping Chen. 2024. "Towards High-Performance Pockels Effect-Based Modulators: Review and Projections" Micromachines 15, no. 7: 865. https://doi.org/10.3390/mi15070865

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop