Next Article in Journal
A Self-Powered DSSH Circuit with MOSFET Threshold Voltage Management for Piezoelectric Energy Harvesting
Next Article in Special Issue
Beam Steering Technology of Optical Phased Array Based on Silicon Photonic Integrated Chip
Previous Article in Journal
Junction Temperature Optical Sensing Techniques for Power Switching Semiconductors: A Review
Previous Article in Special Issue
A 3C-SiC-on-Insulator-Based Integrated Photonic Platform Using an Anodic Bonding Process with Glass Substrates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Breakthrough in Silicon Photonics Technology in Telecommunications, Biosensing, and Gas Sensing

by
Muhammad Shahbaz
,
Muhammad A. Butt
* and
Ryszard Piramidowicz
Institute of Microelectronics and Optoelectronics, Warsaw University of Technology, Koszykowa 75, 00-662 Warszawa, Poland
*
Author to whom correspondence should be addressed.
Micromachines 2023, 14(8), 1637; https://doi.org/10.3390/mi14081637
Submission received: 2 August 2023 / Revised: 17 August 2023 / Accepted: 18 August 2023 / Published: 19 August 2023
(This article belongs to the Special Issue Silicon Photonic Devices and Integration)

Abstract

:
Silicon photonics has been an area of active research and development. Researchers have been working on enhancing the integration density and intricacy of silicon photonic circuits. This involves the development of advanced fabrication techniques and novel designs to enable more functionalities on a single chip, leading to higher performance and more efficient systems. In this review, we aim to provide a brief overview of the recent advancements in silicon photonic devices employed for telecommunication and sensing (biosensing and gas sensing) applications.

1. Introduction

Silicon (Si) photonics is a groundbreaking technology that merges the fields of Si microelectronics and photonics to enable the manipulation and transmission of light on a Si chip. It leverages the exceptional properties of Si, such as its high refractive index and compatibility with existing electronic manufacturing processes, to create compact and highly efficient optical devices. Si photonics has the potential to revolutionize various domains, including telecommunications, data centers, sensing, and biomedical applications [1,2,3,4,5], by offering high-speed data transmission, low power use, and integration with electronic systems. Its ability to seamlessly integrate photonics with Si electronics opens new avenues for the development of advanced, scalable, and cost-effective solutions for a wide range of applications.
Si, the second most abundant element on Earth after oxygen, possesses exceptional qualities that make it highly suitable for various applications. Its simple cubic crystal structure allows production of defect-free wafers with remarkable purity. Additionally, Si’s thermal conductivity, hardness, and low density are advantageous in semiconductor devices. At a specific wavelength, Si demonstrates a high refractive index and remains transparent to infrared light. The high refractive index of Si allows the miniaturization of devices to incredibly small scales. Moreover, the well-established techniques used in semiconductor processing can be easily applied to Si photonics, enabling cost-effective mass production. Another advantage of Si is its high-quality native oxide, which offers superior characteristics compared with other semiconductors. The oxide serves as an excellent material for Si waveguide (WG) cladding and can host rare-earth dopants for integrated circuits. By controlling the oxide cladding layer, it becomes possible to manipulate light propagation within Si WGs. These properties collectively position Si photonics as a promising solution for integrating photonic circuits.
Si photonics integrated circuits rely on key components to manipulate and control light signals. WGs guide light through total internal reflection, whereas modulators alter light intensity or phase. Photodetectors convert optical signals to electrical ones, and filters selectively transmit specific wavelengths [6,7]. Splitters and couplers distribute and combine signals, and optical amplifiers enhance weak signals. These building blocks enable applications in optical communication, data centers, sensing, and biomedical imaging, propelling advancements in photonics technology [8].
Researchers have made substantial progress in increasing the sensitivity and limit of detection (LoD) of Si photonic sensors. By enhancing the design of WGs and resonators and incorporating advanced materials, these sensors can now detect even smaller quantities of substances, making them suitable for a wider range of applications, including environmental monitoring, healthcare, and security. The integration of sensors on a single Si chip offers numerous benefits, including reduced size, increased portability, and cost-effectiveness. The ability to combine multiple sensors on a single chip allows for compact and robust sensing platforms that can be easily mass-produced. Moreover, advancements in fabrication techniques, such as nanofabrication and complementary metal oxide semiconductor (CMOS)-compatible processes, have enabled the cost-effective mass production of Si photonic sensors [9]. This scalability has opened new opportunities for commercial applications and broader adoption of the technology.
Although Si photonics offers numerous advantages, it is worth noting that other optical platforms, such as indium phosphide (InP) [10], gallium arsenide (GaAs) [11], lithium niobate (LiNbO3) [12,13], and rubidium titanyl phosphate (RTP) [14,15], also have their own unique strengths and are preferred for certain specialized applications. The choice of platform depends on factors such as performance requirements, cost considerations, and the specific needs of the application at hand. For more detailed information on optical platforms and fabrication methods, we recommend readers to see [16].
Si photonics remains an evolving field, continuously pushing the boundaries of optical technology. Several review papers have delved into the diverse facets of Si photonics, shedding light on its current state and prospects [17,18,19]. In this review, we delve into the early research efforts and provide an overview of the significant advancements in the field of Si photonics. We emphasize the critical breakthroughs that have shaped the development of this technology. Furthermore, we explore the diverse range of applications (as shown in Figure 1) that leverage Si photonics, showcasing its versatility and potential impact in various fields.

2. History

Since the 1970s, researchers have envisioned an optical super chip with integrated optical components [20]. Early studies focused on ferroelectric materials such as lithium niobate (LiNbO3) and III-V semiconductors such as gallium arsenide (GaAs) and indium phosphide (InP). LiNbO3 stood out for its significant electrooptic coefficient, enabling optical modulation through the Pockels effect. Meanwhile, the III-V compounds offered advantages in terms of laser fabrication, optical amplification, and electronic integration. Si’s widespread use as the preferred semiconductor in electronics prompted researchers to explore the possibilities of Si photonic circuits. This was mainly driven by the potential benefits of integrating photonics and electronics cost effectively. Si has a high refractive index compared with other common materials used in photonics. This property allows for strong light confinement and efficient light guiding within Si WGs and resonators. Consequently, Si photonic components can be designed on a small scale due to the strong confinement of light within Si WGs. This compactness is especially advantageous for on-chip integration and the creation of complex optical circuits. The investigation of Si photonic circuits started in the mid-1980s and has been ongoing since then. A significant volume of research has been conducted on various materials such as SiON [21,22,23,24,25,26,27], Si3N4 [28,29,30], SiGe [31,32,33,34], SiO2 [35,36,37], and SiC [38,39,40]. These materials have been explored concerning their compatibility with standard CMOS technology. Si dominates as the primary semiconductor material for electronics due to its affordability, well-understood properties, and optical confinement capabilities. However, its indirect bandgap limits its effectiveness as a light-emitting material. Researchers have thus sought ways to modify Si’s structure for light emission [41].
Additional efforts were undertaken to define a completely integrated monolithic optoelectronic super chip for Si hybrid integration. This idea was later modified to introduce a hybrid device that relied on a Si platform, incorporating Si optical WGs [42] as depicted in Figure 2.

3. CMOS Fabrication

CMOS is a semiconductor technology commonly used in electronic devices and integrated circuits and is crucial for Si photonics, enabling the integration of photonic and electronic elements on a single Si chip. CMOS technology is widely adopted for its energy efficiency, scalability, and versatility, making it a fundamental component in the design and manufacture of modern electronics [16]. Photolithography, using optical or extreme ultraviolet (EUV) light, transfers patterns onto a Si wafer coated with photosensitive material. Ultraviolet lithography (UVL) represents a distinct form of photolithography wherein UV light serves as the exposure source [43]. This technique finds extensive application in semiconductor manufacturing and various industries to produce patterns on a substrate. The utilization of UV light with its shorter wavelength enhances resolution and facilitates the creation of smaller, complex devices [44]. Nonetheless, UVL demands meticulous attention and specialized equipment due to its distinctive properties. Electron beam lithography (EBL) is a potent nanofabrication technique that achieves extremely high-resolution patterns in nanometers, enabling precise nanostructures and devices [45]. Though versatile, it is relatively slow and expensive compared with other lithography methods due to the time-consuming electron beam scanning process [46]. Nanoimprint lithography (NIL) offers a cost-effective alternative [47] to high-resolution lithography techniques such as EBL. It enables high-throughput production by processing large substrate areas simultaneously [48]. Additionally, unlike EBL, NIL does not necessitate complex and expensive equipment.
Reactive ion etching (RIE) [49,50] and chemical wet etching [51,52] are common techniques in microfabrication and semiconductor manufacturing. They offer selective material removal from a substrate for pattern creation. Chemical wet etching suits simple and large-area patterning, whereas RIE is better for high-resolution and anisotropic etching [53].
The combination of CMOS technology with various lithography and etching techniques enables the creation of complex and high-performance photonics devices that drive technological advancements in a wide range of industries. As technology continues to evolve, further innovations in CMOS and lithography techniques will likely contribute to even more exciting developments in the field of photonics and beyond.

4. Recent Advances in Si Photonics for Telecommunication

Recent progresses in Si photonics have greatly impacted the field of telecommunications. It combines the advantages of Si-based electronic circuits with the speed and bandwidth of optical communications, enabling the development of highly efficient and scalable optical devices [54,55]. By integrating optical components such as modulators, detectors, and WGs directly onto a Si chip, Si photonics allows for the seamless integration of optical and electronic functionalities. This technology enables high-speed data transmission over long distances, facilitates the growth of data-intensive applications, and provides a pathway for the development of compact, low-cost, and energy-efficient telecommunication systems. Its compatibility with existing Si manufacturing processes makes it an attractive solution for driving the next generation of communication networks, data centers, and high-performance computing infrastructures [56]. In the following sections, we discuss various technologies and applications associated with Si photonics.

4.1. Si-Based Modulators

In recent times, there has been a remarkable advancement in the efficiency of Si optical modulators, which corresponds to substantial research endeavors conducted by academic institutions and businesses on a global scale [57]. The objective behind enhancing the performance of optical modulators in Si is evident. These modulators play a pivotal role in most photonics systems used in data communication applications. By utilizing Si photonics as a cost-effective foundation for constructing these systems, optical-based communication becomes a viable option for numerous short-distance connections. Over the years, numerous methods have been explored to achieve modulation in Si by integrating it with emerging optical materials such as graphene [58,59,60,61]. Chip-scale modulation is an area of great interest, particularly concerning silicon-on-insulator (SOI) modulators [19].
Si modulators employ the phenomenon known as free-carrier plasma dispersion. In the context of Si photonics devices, implanted dopants are used to achieve optical modulation in terms of phase and amplitude. This modulation is accomplished by inducing changes in the complex refractive index of a material containing an excess of either electrons or holes, which is wavelength dependent. Consequently, by manipulating the concentration of free carriers within a WG [62], it becomes possible to modulate the phase and amplitude of the light propagating through it.
Si modulators can be integrated with CMOS technology, which is widely used in the semiconductor industry. This integration enables the creation of photonic integrated circuits that combine both electronic and photonic components on the same chip. This compatibility allows for cost-effective mass production and easy integration with existing electronic systems. These modulators can be fabricated on a small scale, enabling the creation of compact devices and circuits. This compactness is especially valuable for applications where space is limited, such as in data centers, optical interconnects, and portable devices. Moreover, Si modulators are capable of achieving high modulation speeds, making them suitable for high-speed optical communication systems. They can operate at data rates exceeding tens of gigabits per second and are continuously being optimized for even higher speeds [63].
However, Si modulators can exhibit nonlinear behavior, which can lead to signal distortion and the generation of unwanted harmonics. This can limit their performance in high-speed and high-power applications. The operation of Si modulators often involves changing the refractive index of the Si material using electrical signals. This can lead to heat generation, which can induce thermal effects and affect the modulator’s performance and stability [64]. Additionally, the modulation bandwidth of Si modulators can be limited by various factors, including the carrier lifetime and the device’s physical dimensions. This can restrict their performance in applications requiring high data rates.
Multiple studies have documented the use of modulators based on Mach–Zehnder interferometric (MZI) designs [65,66,67,68,69,70,71,72]. A comprehensive study of an on-chip optical modulator utilizing a non-conventional Si-based platform is proposed [73]. The platform is based on the optimum design of an ultra-thin Si-on-insulator (SOI) WG, which exhibits low field confinement within the core WG and high sensitivity to the cladding index. Impressive results were obtained, with an extinction ratio (ER) exceeding 20 dB and an energy per bit of 13.21 fJ/bit for an applied voltage of 0.5 V [73]. The platform’s performance for modulation applications was found to be promising and adequate, with the added advantages of cost effectiveness and ease of fabrication. Researchers, in 2007, introduced ultra-compact Si p+-i-n+ diode Mach–Zehnder electro-optic modulators measuring 100 to 200 µm in length [68]. The active WG structure’s cross-sectional scanning electron microscope (SEM) image is presented in Figure 3a. The performance of these modulators was noteworthy, showcasing remarkable modulation efficiency. Specifically, the Vπ·L figure of merit was determined to be 0.36 V-mm. The experiment successfully demonstrated optical modulation at data rates reaching 10 Gb/s while consuming minimal RF power, with an energy consumption of merely 5 pJ/bit [68].
Another study proposed and investigated a nonlinear porous Si-based all-optical modulator using an asymmetrical MZI design [74]. The MZI comprised two couplers with an identical splitting ratio and two unbalanced arms. In this configuration, a porous Si (PS) WG is placed in only one arm of the MZI, whereas the other arm functions as a short fiber delay line, as shown in Figure 3b. The authors conducted experiments where a pulsed pump and a continuous probe wave were at the same time launched into the input port. The results indicated that the device achieved an outcome signal with approximately 14.10 dB modulation depth at the probe wavelength. The experimental conditions include an initial pulsed pump with a peak of 47.07 dB m, an extinction ratio of 10.88, and a duration of 100 ps. The continuous probe wave has a power of 0 dB m, and the PS WG is 3 mm long [74].
Another popular type of modulator relies on ring resonators (RRs), which are actively being researched by scientists [75,76,77,78,79]. These modulators effectively manipulate light signals by leveraging the unique properties of RRs. They offer versatility and are widely utilized in photonics for various communication and scientific purposes as researchers continue to explore their potential applications [80,81,82,83,84,85,86,87]. A high-speed Si modulator based on cascaded double µ-RRs is explained in a study conducted by the authors of [88]. In Figure 4a, a top view of a microscopic picture is presented, showcasing the fabricated cascaded micro-ring modulator and a schematic 3D view of the interleaved PN junctions. The modulator achieved a modulation rate of 40 Gbit/s, with an ER of 3.9 dB [88]. To enhance the modulator’s performance, the researchers utilized a cascaded double-ring structure that enabled an ultra-high optical bandwidth of 0.41 nm, equivalent to 51 GHz [88]. In another study conducted by researchers [89], the generation of an optical comb consisting of frequency lines with extremely constant power was reported. The change in power among the frequency lines was less than 0.7 dB [89]. To achieve this, a Si RR modulator was employed. The fabrication of this structure took place on the SOI platform, utilizing processes that are compatible with CMOS technology. Figure 4b presents a top-down view micrograph of the ring, demonstrating its structure. Additionally, a diagram outlines the doped sections of the WG in a cross-sectional representation [89]. To establish connectivity with the device, grating couplers were employed, as illustrated in the micrograph. These couplers were intentionally designed to exclusively guide the TE mode within the C-band, where the mode’s primary axis aligns with the x-direction in Figure 4c. To facilitate the transmission and capture of light, two PM fibers were positioned at an 11° angle relative to the chip’s surface normal and precisely aligned with the grating couplers, as illustrated in Figure 4d. In this work, researchers characterized the ring’s complex transfer function and generated five frequency tones with a 10 GHz spacing using a dual-frequency electrical input at 10 and 20 GHz [89]. The optimal operation was observed at a small forward-bias voltage, as indicated by a comparison of comb shapes. Time domain measurements confirmed highly coherent comb signals, producing 20.3 ps-long pulses [89]. In Table 1, a comprehensive analysis is provided, showcasing the performance of different Si-based modulators that feature diverse structures.

4.2. Wavelength Division Multiplexing (WDM) Systems

The rapid growth of capacity demand in various interconnected systems, such as massively parallelized systems, intra-data-center networks, and wider communication networks, necessitates an urgent need to enhance the density of transceiver capacity [90]. Consequently, increasing the number of wavelength-division multiplexing (WDM) channels emerges as one of the most effective approaches to address this issue. Research and development in the field of WDM on Si photonics-integrated circuits (PICs) began nearly twenty years ago, marking the beginning of significant advancements in this area. Since then, the progress and activity in this field have remained consistently high and continue to be active even today [91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107]. The rise of photonic integration technology has led to a notable interest in two prominent components: arrayed WG gratings (AWGs) and micro-ring-based filters. The footprint of a Si AWG is significantly reduced thanks to the substantial index contrast between its Si core and oxide cladding. This contrast allows for efficient light confinement and manipulation within the device. Moreover, the fabrication process of Si AWGs is compatible with well-established CMOS technology, which not only facilitates the integration of AWGs with other functional devices but also leads to cost reductions. This compatibility enables large-scale production of integrated systems, providing a cost-effective solution for various applications. In a study conducted by researchers in [108], a module comprising a modulator utilizing five channels for wavelength-division multiplexing was introduced. The design of this module can be observed in Figure 5a. This module featured the integration of a Si AWG multiplexer with a channel spacing of 200 GHz and an array of electro-absorption modulators capable of achieving a data rate of 20 Gbps [108]. The results demonstrated the module’s promising capability to support transmission capacities of up to 100 Gbps, all within a compact footprint measuring 1.5 × 0.5 mm2 [108].
The SOI platform emerged as a highly promising avenue for realizing optical transceivers. Nevertheless, a notable drawback of the SOI platform pertained to the relatively elevated thermo-optic (TO) coefficient of Si, leading to a significant spectral response drift of approximately 90 (pm/°K) within the C-band range. This level of drift posed a considerable challenge for various applications, prompting the exploration of numerous strategies to mitigate this issue.
In their groundbreaking study, researchers introduced an innovative advancement in wavelength filters that exhibited significantly reduced sensitivity to thermal fluctuations [109]. This achievement marked the first instance of such wavelength filters, which were engineered by combining crystalline Si and hydrogenated amorphous Si (a-Si:H) WGs. The integration of these WGs took place on a common SOI substrate, facilitated by a process flow compatible with CMOS technology. To validate the efficacy of their novel concept, the team meticulously designed and manufactured MZIs and AWGs utilizing this pioneering approach. Subsequently, precise measurements were conducted to evaluate the impact of temperature variations. The results demonstrated an impressively minimal thermal drift, specifically measuring less than 1 (pm/°K) for MZIs and under 10 (pm/°K) for AWGs within the C band range [109]. This achievement not only showcases the practical viability of the proposed wavelength filters but also underscores the potential for enhanced stability in photonic applications.
Researchers have successfully demonstrated a thermally tunable add/drop filter using a long-period grating configuration that combines Si and TiO2 WGs [110]. The implementation of an apodized grating between the WGs achieved a high SLSR of 25.7 dB. Despite a short LPWG length of 800 µm, a narrow bandwidth of 1.4 nm was achieved due to significant group index differences between the silicon and TiO2 WGs. The distinct thermal coefficients of Si and TiO2 contributed to a 25-fold increase in the thermal dependence of the modes. A thermal tuning efficiency of 0.07 nm/mW was achieved using a thin TiN metallic heater. Additionally, cascading two LPWGs demonstrated a channel spacing of 185 GHz with low power consumption [110]. This development holds promise for DWDM applications.
The demand for optical cross-connects and reconfigurable optical add-drop multiplexers requires switches characterized by attributes such as affordability, energy efficiency, and swift sub-millisecond switching speeds. Addressing these requirements, Si emerged as a promising technology due to its capacity to integrate optical components with electronic control circuitry, enabling the development of diverse photonic devices, including switches [111,112,113]. In their study [114], researchers presented the design of a 2 × 2 switching cell, employing a thermo-optic interferometric setup with a critical component, a sub-wavelength grating. The switching cell demonstrated notable characteristics, including an extinction ratio of approximately 13 dB, insertion loss below 2 dB, and crosstalk of 12 dB across a 150 nm bandwidth. Impressively, all of this was accomplished within a remarkably compact footprint of 240 µm × 9 µm [114]. The potential application of this design in flexible telecommunication satellite payloads was showcased by employing the switching cell as a fundamental unit in an 8 × 8 dilated Banyan matrix. This configuration demonstrated impressive attributes such as a broad bandwidth (150 nm), minimal crosstalk (−38 dB), a compact footprint of around 1620 µm × 576 µm, and a relatively low power consumption at 276 mW [114]. In other work, researchers successfully demonstrated a 4 × 4 fully non-blocking crossbar switch fabric by utilizing interferometric thermal phase shifters [115]. These phase shifters employed resistive elements placed around Si WGs to achieve controlled heating. A resulting switching time of 5 μs was recorded while demonstrating energy efficiency, with a power consumption of 41 mW per individual switching element [115].
In the realm of on-chip WDM optical interconnects, Si µ-RRs exhibit remarkable capabilities as wavelength filters, offering significant benefits in terms of their ultra-compact footprint and extraordinary energy efficiency. However, it is important to note that the resonant wavelength of Si-µ-RRs is highly sensitive to both temperature variations and variations in the fabrication process [116,117]. These factors can introduce challenges when it comes to maintaining precise control over the resonant wavelength of the Si-µ-RRs. In a notable advancement, a recent experimental study [118], demonstrated gate-tuning on-chip WDM filters for the first time. The study achieved a substantial wavelength exposure across the entire channel spacing using a Si-µ-RR array (as shown in Figure 5b) that was controlled by high-mobility titanium-doped indium oxide (ITiO) gates [118]. The integrated Si-µ-RRs exhibited an unprecedented level of wavelength tunability, reaching up to 589 pm/V or VπL of 0.050 V cm while maintaining a high-quality factor of 5200 [118].
Figure 5. (a) Schematic diagram of the chip layout [108]; (b) 3D schematic of the on-chip WDM filters. Inset: zoom-in view of µ-RR WG with ITiO/HfO2/Si MOS capacitor [118].
Figure 5. (a) Schematic diagram of the chip layout [108]; (b) 3D schematic of the on-chip WDM filters. Inset: zoom-in view of µ-RR WG with ITiO/HfO2/Si MOS capacitor [118].
Micromachines 14 01637 g005

4.3. Photodetectors

In the realm of optical interconnect technology, the photodetector (PD) stands as a vital component responsible for converting optical signals into electrical ones. As the demand for high-capacity optical interconnect systems continues to surge, the need for high-speed PDs becomes increasingly prominent. To meet this demand, researchers have made significant strides in developing high-performance PDs using diverse absorption materials and innovative structures on the Si photonics platform. Notable examples include germanium (Ge) PDs [119,120,121], germanium tin (GeSn) PDs [122,123,124,125], heterogeneous integrated III–V PDs [3,126,127], all-Si (Si) PDs [128,129], etc. The significance of these advancements lies in their ability to facilitate efficient and reliable communication in systems that demand large data transmission capacities. High-speed PDs play a crucial role in capturing optical signals and converting them into electrical form, enabling seamless data transfer between optical and electronic domains. As the demand for faster and more efficient communication continues to grow, the development of high-speed PDs will remain a focal point of research and innovation in the field of optical interconnect.
The fabrication process of the Ge PD involves the selective growth of a Ge film onto a Si substrate. The point of this fabrication technique is to exploit the desirable characteristics of both materials. Ge is renowned for its excellent optical properties, particularly in the infrared wavelength range, which makes it highly suitable for photodetection applications. On the other hand, Si is widely recognized as a semiconductor material with superior electronic properties. In 2007, researchers presented a Ge p-i-n PD that was seamlessly incorporated into SiON and Si3N4 WGs [130]. They ensured that all the procedures and substances employed were compatible with CMOS technology and could be readily incorporated into the existing integrated circuit process technology. In later work [120], high-speed Ge p-i-n PDs for vertical incidence were introduced, featuring remarkable responsivity, and grown using reduced pressure chemical vapor deposition (CVD) [120]. The authors of [131] presented a demonstration of 100 Gbps Si-contacted Ge WG p-i-n PDs that were successfully integrated on IMEC’s Si photonics platform. In another study [132], Ge WG PDs were created on SOI substrates using selective epitaxial growth. These PDs had a width of 4 μm and were fabricated with different lengths.
In a previous study [133], researchers integrated a compact pin Ge photodetector into a submicron SOI rib WG. They achieved a remarkable reduction in the detector length, which was brought down to 15 μm through a butt-coupling configuration. The schematic representations of the integration of the WG Ge PD and the cross section of the pin diode are shown in Figure 6a,b, respectively. This size reduction allowed the detector to efficiently absorb light at the specific wavelength of 1.55 μm [133]. When subjected to a 4 V reverse bias, the photodetector exhibited a −3 dB bandwidth of 42 GHz. Moreover, it demonstrated an impressive responsivity of 1 A/W at a wavelength of 1.55 μm, coupled with a low dark current density of 60 mA/cm2. Under a −0.5 V bias and at a slightly shorter wavelength of 1.52 μm, the photodetector still maintained a responsivity of 1 A/W [133]. In photonics-integrated circuits, the input of a regular Ge PD consisted of a single-mode Si WG, and the optical connection between the input WG and PD is linked using a tapered Si WG. Notably, a novel design featuring a multi-mode WG input PD was introduced [134]. This innovative approach allowed for the retention of high-order modes from the output end of the MM-WDMs, thereby ensuring the complete transmission of light to the detector. Figure 6c illustrates the cross-sectional architecture of the vertical N-I-P photodetector, whereas Figure 6d displays an SEM image of the manufactured device [134]. The fabricated devices exhibited low dark current (10 nA at −1 V) and high responsivity (>0.75 A/W from 1270 to 1350 nm). They showed a 23 GHz optoelectrical bandwidth at −3 V bias and achieved a clear 50 Gb/s eye diagram with NRZ modulation [134]. These results highlight the devices’ excellent performance for photodetection and high-speed optical communication.
Despite its recent successes, the Ge PD is not suitable for extended wavelength detection, such as the L band or >2000 nm, due to the rapid drop in Ge’s absorption curve beyond 1610 nm. To address this limitation, adding Sn to Ge to create a GeSn alloy is proposed as an alternative solution [122,123,124,125]. Increasing the Sn concentration decreases the direct bandgap of GeSn, resulting in an extended absorption cut-off wavelength. According to [135], researchers successfully showcased the effectiveness of a Si surface passivation technique that was compatible with CMOS technology. The initial experimental validation of a germanium–tin (GeSn) lateral p-i-n PD on a unique GeSn-on-insulator (GeSnOI) substrate was reported [136]. They achieved a cutoff detection beyond 2000 nm, with a responsivity of 0.016 A/W at the wavelength of 2004 nm [136].
In a successful demonstration [137], researchers implemented a GeSn photodetector on an SOI substrate for 2 μm wavelength applications. The 3D schematic of the normally illuminated p-i-n PD is depicted in Figure 7a, whereas Figure 7b presents the top-view SEM image of the device, featuring a mesa with a diameter of 10 μm [137]. The device exhibited a low dark current of approximately 125 mA/cm2 at room temperature under a reverse bias of −1 V. It also achieved an impressive optical responsivity of 14 mA/W for a 2000 nm illumination wavelength under the same bias conditions. Furthermore, the PD demonstrated a remarkable 3 dB bandwidth of around 30 GHz at −3 V [137]. A recent study demonstrated CMOS-compatible GeSn WG photodetectors (WGPDs) with a vertical p-i-n heterojunction design for operation in the 2000 nm wavelength band [138]. The schematic representation of the fabricated device is shown in Figure 7c, whereas an SEM image of the fabricated device is presented in Figure 7d. By incorporating 5.28% Sn into the GeSn active layer, the photodetection range was redshifted to 2090 nm, whereas the proposed GeSn WGPD exhibited improved light–matter interaction and optical confinement. This resulted in a responsivity of up to 0.52 A W−1 at room temperature within the 2000 nm wavelength band [138].
Over an extended duration, there has been extensive research on III-V PDs that exhibit exceptional performance. When compared with group IV PDs, III-V PDs demonstrate lower dark current density, allowing for improved efficiency. Furthermore, they offer wavelength tunability by employing bandgap engineering techniques, enabling the detection of a wide range of light wavelengths. Additionally, III-V PDs can achieve ultrahigh-speed operation, making them highly desirable for applications requiring rapid data processing and communication. In recent years, numerous instances have showcased the successful combination of III–V PDs with Si, employing diverse integration techniques. These methods encompass direct growth [126,139,140], wafer bonding [3,141], and transfer printing [127,142]. Each approach offers a distinct approach to the heterogenous integration of III–V PDs onto Si, resulting in significant advancements in this field. Top-illuminated PIN and modified uni-travelling carrier (MUTC) photodiodes were grown epitaxially on Si templates using InGaAs/InAlAs/InP as the material composition [143]. In this study, the achieved dark currents of the photodiodes were as low as 10 nA at 3 V, which corresponds to a low dark current density of only 0.8 mA/cm2. The responsivity of the photodiodes was reported to be 0.79 A/W, whereas the 3 dB bandwidth reached 9 GHz [143]. As we have already discussed, GeSn PDs extend the wavelength range of Ge PDs but face low responsivity due to lower absorption coefficients at longer wavelengths. In contrast, III–V PDs have no such limitation and easily achieve high responsivity in the short-wavelength or middle-wavelength infrared regimes. The researchers in [144], focused on the combination of InP-based type-II quantum well photodiodes on Si PICs for the range of 2 µm wavelength. In this research, the authors successfully demonstrated a responsivity of 1.2 A/W at a wavelength of 2.32 µm. The photodiodes exhibited a dark current of 12 nA at a bias voltage of −0.5 V at room temperature [144]. Comparing the transfer printed method to the direct growth method for PD production, the transfer printing approach offers the advantages of lower dark current and reduced process complexity. These benefits improve sensitivity and streamline manufacturing, making it an appealing choice for photodetector production. In a study [127], the authors proposed and demonstrated the integration of metal–semiconductor–metal (MSM) PDs based on transfer printing for use in photonic interposers. The PDs were GaAs based and operated at a wavelength of 850 nm. The larger PD exhibited a dark current of 22 nA at a bias of 2 V, whereas the smaller PD had a dark current of 7.2 nA under the same bias conditions [127]. At 2 V bias, the external responsivities for the 850 nm wavelength were measured as 0.117 A/W for the larger PD and 0.1 A/W for the smaller PD [127]. Additionally, the PDs reached a bandwidth of 20 GHz, and open-eye diagrams at a data rate of 40 Gb/s were successfully obtained [127].
All-Si PDs refer to all-Si photodetectors, which are a specific category of high-speed PDs developed within the field of Si photonics. These PDs are designed and fabricated using Si-based materials and have proven to be capable of operating at high speeds. Over time, researchers have demonstrated various types of all-Si PDs [128,129,145,146,147], showcasing the versatility and potential of this technology. The authors of study [148] reported a sub-bandgap linear-absorption-based photodetector operating in avalanche mode at a wavelength of 1550 nm. This photodetector was realized by integrating a PN diode into a Si µ-RR [148].
In another study [149], the authors suggested and operated active resonance wavelength stabilization for Si µ-RRs. This was achieved by utilizing an in-resonator defect-state-absorption (DSA)-based PD specifically designed for optical interconnects [149]. The researchers conducted a demonstration of a CMOS-compatible PD based on all-Si WGs, achieving a measured speed exceeding 35 GHz [150]. The device configurations are illustrated in Figure 8a–c. To create the device, Si ions were inserted into an intrinsic Si ridge WG, which was then subjected to annealing. This PD exhibited operational capabilities within the wavelength range of 1100 to 1750 nm, with internal responsivity varying from 0.5 to 10 A/W under different bias voltages [150]. With the assistance of a cavity enhancement effect, numerous photodiodes in the past demonstrated significantly high responsivity at telecommunication wavelengths, specifically at 1310 nm. However, the underlying mechanisms that contribute to such heightened responsivity have yet to be fully understood. In a recent study [151], the researchers systematically investigated an all-Si micro-ring as a photodiode to unravel the diverse absorption processes involved [151]. The schematic representation of the all-Si MRR APD is shown in Figure 8d [151]. At a bias voltage of −6.4 V, the micro-ring exhibited a responsivity of approximately 0.53 A/W, accompanied by avalanche gain. The device further showcased a 3 dB bandwidth of around 25.5 GHz and produced open-eye diagrams with a transmission rate of up to 100 Gb/s, as shown in Figure 8e,f [151]. Table 2 provides a detailed and comprehensive analysis, offering a thorough comparison and summary of the performance of different optical photodetectors.

5. Si Photonics Sensors

Si photonic sensors have gained significant attention and interest in both gas sensing and biosensing applications due to their exclusive advantages, such as high sensitivity, miniaturization capabilities, and potential for integration with existing Si-based electronic systems [152]. In this section, we reviewed recent advancements in Si photonic sensors employed in biosensing and gas-sensing applications.

5.1. Si Photonics in Biosensing

Technological progress has brought about the convergence of various fields of science and engineering, such as electronics, medicine, and biology. This convergence has led to the development of biosensors, which are now commonly associated with point-of-care medical diagnosis and precise healthcare [153]. Biosensors have found applications in multiple areas, involving food safety, pharmaceutical research, healthcare, and cancer research [4,154]. These devices can identify and transform biological signals into electrical or optical signals [155].
Si-based materials offer significant advantages over alternative substances such as lithium and GaAs, which are commonly utilized in optical biosensors. One key advantage is their comparatively lower cost, making them more economically viable. Additionally, they can be produced on a large scale utilizing CMOS technology [156]. Moreover, Si-based materials exhibit a substantial difference in refractive index due to the distinct refractive index values of Si and Si dioxide, which are employed as the core and cladding materials, respectively. This discrepancy in the refractive index helps minimize potential errors that may arise from fluctuations in the refractive index of the sensor interface. The use of SOI wafers enables the production of both electronic and photonic components on a single chip, resulting in a compact platform consolidated on a solitary Si-based substrate [157]. In optical biosensor applications, two primary factors are consistently considered: sensitivity (S) and quality factor (Q) [158,159]. These parameters play a crucial role in distinguishing the performance of various sensors.
S is the measurement of the ratio between the change in sensor output and the corresponding change in the quantity being measured. It serves as an indicator of how effectively the sensor can detect and respond to variations in the measured parameter. The S of a biosensor is typically calculated as follows:
S = Δ λ Δ n
Q is another essential parameter in optical biosensing. It characterizes the effectiveness and selectivity of a biosensor. A higher Q indicates a more precise and accurate measurement, with reduced noise and interference. The Q is determined by factors such as the sensor’s resonance properties, the stability of its components, and the overall system design. The Q is calculated by the following expression:
Q = Δ λ ( r e s . ) Δ λ ( F W H M ) .
Both S and Q are crucial considerations when evaluating the performance and reliability of optical biosensors. Achieving high sensitivity and a high-quality factor is essential for accurate and precise measurements in biosensing applications.
To enhance the sensitivity of the Si-based photonic sensors, researchers have proposed highly sensitive WG structures such as suspended silicon WG [160,161,162] and subwavelength grating (SWG) WGs [163]. Suspended WGs are typically fabricated on SOI substrates, where a thin silicon layer is separated from a SiO2 layer by a buried oxide layer. The term “suspended” refers to the fact that the Si WG core is physically suspended above the substrate by removing the surrounding SiO2, thus allowing for tighter confinement of the optical mode and interaction with the surrounding environment. The suspended design allows for strong interaction between the optical mode and the analyte being sensed. This leads to increased sensitivity, making them suitable for detecting even small changes in the refractive index or other properties of the surrounding medium. Moreover, the strong confinement of light in the WG core enables compact device designs. This is particularly useful for on-chip integration with other optical components and for creating dense arrays of sensors.
SWG WGs use a periodic pattern of subwavelength-sized features to guide and manipulate light. They have gained significant attention in the field of photonics and sensing due to their unique properties and capabilities [164]. SWG WGs can be used for various applications, including sensing, due to their ability to confine and interact with light in a compact and efficient manner [165]. The periodic subwavelength structure of the WG can be tailored to achieve strong light–matter interactions with specific analytes. This enhanced interaction can enable higher sensitivities in detecting changes in refractive index, temperature, or other environmental parameters [166]. SWG WGs support the propagation of evanescent fields beyond the core region of the WG. These evanescent fields can be used to interact with substances on the WG’s surface. By monitoring changes in the evanescent field due to the presence of target molecules or changes in the surrounding medium, sensitive and label-free sensing can be achieved [167].
Extensive research has been conducted in the field of biosensors based on Si photonics, resulting in the development of several types with diverse applications [168,169,170,171,172,173,174]. These biosensors utilize the advantageous properties of Si photonics to detect and analyze biological substances effectively. The initial implementation of an interferometer WG-driven biosensing application was documented during the early 1990s [175]. This interferometer employed the Mach–Zehnder interferometer (MZI) principle, where light is divided into two separate arms and then both signals are brought together once again using a Y-shaped junction [175]. In the year 2000, the initial use of a Young interferometer (YI) as a biosensor was documented, achieving a bulk sensitivity of 9 × 10−8 RIU [176]. The YI employs a Y-junction to split a single optical path into two separate arms. Unlike the MZI, the interference pattern is generated outside the chip since there is no recombination of the two arms. The authors of [177] introduced a biosensor based on Si3N4 slot WGs, utilizing a Mach–Zehnder Interferometer (MZI) design, as shown in Figure 9 [177]. The biosensor was specifically engineered to exhibit minimal temperature dependency while maintaining high sensitivity and a low LoD. The measured surface sensitivity and LoD were reported to be 7.16 nm/(ngmm−2) and 1.30 (pgmm−2), respectively [177]. Additionally, the temperature dependence was found to be as low as 5.0 pm/°C. On the other hand, when water was used as the cladding material, the bulk sensitivity and LOD were reported to reach 1730(2π)/RIU and 1.29 × 10−5 RIU, respectively. By incorporating the Vernier effect via cascaded MZI structures, the authors achieve a sensitivity improvement factor of 8.38, resulting in a surface LoD of 0.155 (pgmm−2) [177].
The efficiency of the photonic WG sensor can be validated by considering the parameter of bulk sensitivity. Current improvements in point-of-care Si photonic biosensing have led researchers to focus on identifying effective methods for improving sensitivity. One approach involves integrating a microfluidic channel made of polydimethylsiloxane (PDMS) into the sensor architecture. However, this integration may result in decreased sensitivity due to molecule leakage at the edges of the channel [178]. To address this issue, a Si3N4 MZI is utilized, which takes advantage of the refractive index of different cancer cells [179]. The authors examined the effectiveness of both gradient rib and gradient rib-slot WGs in this analysis, as depicted in Figure 10 [179]. This structure is specifically proposed to securely secure the liquid sample without the use of PDMS material. Compared with the gradient rib WG, this novel WG demonstrates significantly higher WG bulk sensitivity and device bulk sensitivity [179]. The researchers were able to achieve a WG bulk sensitivity of 2.0699 RIU/RIU and a device sensitivity of 568 nm/RIU [179].
Researchers have designed a RR-based biosensor aimed at detecting cancer cells and hemoglobin concentrations while optimizing its structure to enhance S and Q [180]. The biosensor demonstrated the capability to detect a range of cancer cells and hemoglobin concentrations with a remarkable sensitivity of 200 nm/RIU and a high Q factor of approximately 2000 [180]. In a recent study conducted by researchers, an optical-based label-free biosensor was developed and optimized for sensing applications. The biosensor consisted of two indirectly coupled double-slot-WG-based µ-RRs, as shown in Figure 11 [181]. The optimized system was then employed to detect hemoglobin concentration, specifically targeting the identification of anemia disease [181]. To evaluate the biosensor’s performance, nine different concentrations of hemoglobin were introduced to the sensor for both men and women. The study aimed to determine the condition of anemia based on gender and various levels of the disease, involving normal, mild, moderate, severe, and deadly conditions [181]. Remarkably, the biosensor exhibited a high sensitivity of 1024 nm/RIU, accompanied by a minimum deflection limit of 4.88 × 10−6 RIU [181]. These results demonstrate the biosensor’s ability to provide precise measurements at a microscale, presenting a promising lab-on-a-chip microdevice for monitoring the health of individuals with anemia.
The application of porous Si (PSi) as an optical sensor to detect chemicals and molecular interactions has been explored. The researchers employed electrochemical etching of crystalline Si in solutions containing HF to create various PSi layers. Additionally, they employed a combination of physical, physicochemical, chemical, and electrochemical post-procedures to further enhance the properties of the PSi layers [182,183]. PhC-based biosensors have captured significant attention as a promising and groundbreaking technology. Researchers have conducted investigations into Si-based PhC devices on the SOI platform, and these devices have experienced rapid development. This technological advancement has led to the creation of various architectures, such as 1D PhC [172] and 2D PhC [184], specifically designed for the detection of biomolecules.
In a study conducted by researchers [185], a biosensor platform was introduced that utilized microcavities coupled to an optical WG. The researchers emphasized the high sensitivity of this platform, measuring approximately 97 nm/RIU. Furthermore, the biosensor exhibited an LoD as low as 0.0055 fg [185]. In their research, the detection of the hepatitis B virus was monitored using a 1D PhC [186]. The study focused on utilizing an antibody to specifically detect the surface antigen (HBsAg-ayw). The biosensor based on the 1D-PhC design achieved an impressive LoD of 460 pg [186]. In 2021, a study was conducted to suggest a refractive index sensor capable of the detection of cancer and diabetes using PhC at the same time [187]. The fundamental structure of the proposed PhC consisted of Si rods arranged in a hexagonal lattice within an air bed. To facilitate the measurements, two tubes were utilized to accommodate the samples of cancerous or diabetic substances, as illustrated in Figure 12a [187]. The researchers recognized the importance of sensitivity and accuracy in sensor design. Consequently, they evaluated the performance of the sensor by considering the FWHM value, which was measured at 1.8 nm. Moreover, they assessed the sensitivity and FOM of the sensor following modifications made to the shape of the dielectric rods and nanocavities. The results indicated that the sensor achieved a best sensitivity value of S = 20,393 nm/RIU and an FOM of 9104.017 ± 606.93 RIU−1 [187]. A recent study proposed the utilization of a RR in 2D PhC biosensors [188]. The biosensor configuration, as depicted in Figure 12b, consists of an RR accompanied by two bus WGs. To form the resonator and WG, a row of rods was selectively removed from the structure. Specifically, the ring RR was composed of a large green dielectric rod encircled by four smaller rods, measuring 340 nm and 162 nm in radius, respectively [188]. The objective of this design was to enable the detection and differentiation of both normal and cancerous cells. The researchers assessed the performance of the proposed structure based on various parameters. The obtained average values for S, Q, FOM, and transmission were 308.5 (nm/RIU), 3803.55, 848.06 RIU−1, and 98.78%, respectively [188]. In another recent study [189], the authors introduced a biomedical sensor based on PhC, as illustrated in Figure 12c. The sensor was designed to accurately identify and differentiate between normal and abnormal brain tissues, including lesions, tumors, and cancerous tissues. Impressively, the sensor exhibited remarkable performance metrics. Specifically, it achieved a sensitivity of 1332 nm/RIU. Additionally, the sensor demonstrated an exceptionally low LoD of 9.08 × 10−6 and exhibited an ultra-high quality factor of 16,254 [189]. A list of some of the noteworthy proposed biosensors is presented in Table 3.

5.2. Si Photonics in Gas Sensing

The significance of gas sensing has been increasingly recognized in diverse domains and for a broad spectrum of uses, including the detection of dangerous and harmful gases, inspection in industrial settings, and monitoring of the environment [190,191,192]. Numerous optical gas sensors have been created throughout time, benefiting from their remarkable attributes such as heightened sensitivity, dynamic range, stability, rapid response time, and the ability to perform multiple measurements simultaneously [193,194,195,196,197,198,199]. The ability of RI sensing to directly detect biomolecules without the need for labelling and its remarkable sensitivity to even slight variations in the surrounding medium has garnered growing attention. In 2003, an optical nanosensor utilizing Si microelectronics technology was developed; it featured an integrated MZI structure formed by total internal reflection (TIR) WGs [200]. A novel fiber MZI sensor, utilizing a single “S”-shaped fiber taper, was successfully constructed by employing nonaxial pull during the fusion splicing process [201]. The fabricated structure had a characteristic size of 660 μm in length and an axial offset of 96 μm. This particular MZI, based on the S-shaped fiber taper, demonstrated a high RI sensitivity of 1590 nm/RIU within the refractive index range of 1.409–1.425. Additionally, it exhibited a strain sensitivity of approximately −60 pm/microstrain [201]. In another study, a novel mid-infrared (MIR) gas sensor utilizing a suspended Si WG was proposed [202]. Two optimized designs for wavelength and intensity interrogation were developed, achieving high sensitivity and figure of merit [202]. In a recent study [203], researchers introduced a compact refractive index gas sensor by employing an SOI loop-terminated MZI (LT-MZI) and incorporating a slot WG in the sensing arm, as depicted in Figure 13a [203]. Despite its short sensing arm length of 150 μm, the sensor achieved a device sensitivity of 1070 nm/RIU and an impressive figure of merit (FOM) of 280.8 RIU−1 at a wavelength of 1.55 μm [203]. Another recent study, [204], examined the gas sensing capabilities of a compact Si photonics MZI with a coiled sensing arm, as shown in Figure 13b. The sensor, fabricated using deep UV lithography, demonstrated a sensitivity of approximately 1458 nm/RIU and an LoD of approximately 8.5 × 10−5 RIU when tested with helium (He) and nitrogen (N2) gases [204]. The sensor’s temperature sensitivity was found to be 166 pm/°C. Incorporating a cladded ring-resonator post-MZI (Figure 13b) allowed for accurate temperature compensation and resolved temperature drift due to gas flow [204].
µ-RRs in SOI structures have gained attention in sensing due to their high sensitivity to refractive index changes, wavelength multiplexing capability, and compact size [205,206,207]. µ-RRs offer precise detection of analytes, simultaneous measurement of multiple parameters, and integration into portable sensing systems, making them promising for diverse applications. In 2011, the on-chip interrogation of a gas sensor based on SOI µ-RRs was demonstrated using a compact AWG spectrometer [208]. The researchers utilized a 200 GHz SOI AWG with closely spaced output channels to analyze the response of an µ-RR sensor to ethanol vapor concentrations ranging from 100 to 1000 ppm [208]. Researchers, in [209], demonstrated a chip-scale photonic system using a slotted µ-RR cavity, as shown in Figure 14a, to perceive minute refractive index changes in acetylene gas [209]. The nanoscale slot geometry resulted in a significant interaction factor of 0.64, leading to a device sensitivity of 490 nm/RIU. By utilizing a resonator with a Q factor of 5000, they successfully detected refractive index changes at the order of 10−4 [209]. Recently, researchers investigated and introduced a compact grating double-slot µ-RR (GDS-µ-RR) on an SOI platform as a potential solution for achieving a refractive index sensor with a wide measurement range and high sensitivity [210]. The refractive index sensing experiments yielded a sensitivity of 433.33 nm/RIU, a Q-factor of 4325, and a DL value of 8.26 × 10−4 RIU [210]. The researchers concluded that the proposed compact GDS-µ-RR exhibited outstanding performance and held significant potential for various sensing applications. Another study on µ-RR, suggested a high-performance refractive index sensor utilizing a single trapezoidal subwavelength grating slot-µ-RR (T-SWGS-µ-RR)) [211]. The T-SWGS-µ-RR design shown in Figure 14b, achieved a high sensitivity of 823 nm/RIU, a large Q-factor of 2.50 × 104, and a low LoD of 7.53 × 10−5 RIU [211]. The researchers also demonstrated an improved sensitivity of 12,151 nm/RIU and a lower LOD of 2.47 × 10−5 RIU by employing two cascaded T-SWGS-µ-RRs, as depicted in Figure 14c [211].
The MIR region has become the subject of significant interest and research due to its diverse range of applications. One of the reasons the mid-IR region is particularly significant is because it holds the absorption lines of numerous important gases. Gases such as carbon dioxide (CO2), carbon monoxide (CO), and methane (CH4) [4,193,212,213] exhibit characteristic absorption patterns in the MIR range. These absorption lines correspond to specific vibrational and rotational modes of the gas molecules, allowing for their identification and quantification [214,215]. The ability to detect and measure these gases accurately is crucial for various applications, including environmental monitoring, industrial safety, atmospheric studies, and climate research. In [213], Khonina et al. achieved a substantial development in the evanescent field ratio by transforming a ridge WG into a dual hybrid plasmonic WG, as shown in Figure 15a. The WG geometry was adjusted at 3.392 μm, aligning with the absorption line of CH4 gas. The suggested method allowed for a larger WG cross section, facilitating flexible butt coupling of light with an augmented evanescent field. The study validated the results through finite element method (FEM) analysis and reported an elevated evanescent field ratio of 0.74 and a propagation loss of 0.7 dB/μm. A sensitivity of 0.0715 (mW/gas conc.) was also calculated by measuring the decay in transmission power due to gas absorption in the medium [213]. The schematic of the proposed gas sensing procedure is shown in Figure 15b.
In [216], Kazanskiy et al. proposed a polarization-independent strategy of a hybrid plasmonic WG (HPWG) that can be employed as an evanescent field absorption gas sensor, as illustrated in Figure 16, optimized at a wavelength of 3.392 µm for CH4 gas absorption [216]. The HPWG exhibited high sensitivity (Smode) and evanescent field ratio (EFR) for both TE and TM hybrid modes [216]. Modal analysis using the FEM confirmed the TE mode with Smode = 0.94 and EFR = 0.704 and the TM mode with Smode = 0.86 and EFR = 0.67. For a 20 µm-long HPWG at 60% gas concentration, both modes achieved a power dissipation of approximately 3 dB [216].
In the past, researchers developed numerous microstructured gas detectors to integrate them into micro-GC and/or microfluidic systems for rapid on-site gas detection [217,218]. Among the various detection methods, an encouraging approach relied on the utilization of the Fabry–Perot (FP) cavity. This detection scheme had gained popularity, not only in gas detection but also in other fields such as biosensing such as for temperature, strain, and humidity sensing. The FP cavity was favored due to its uncomplicated fabrication process and straightforward measurement setup. Researchers, in [219], investigated a stable optofluidic FP resonator that comprised Si cylindrical Bragg mirrors along with a central capillary tube [219]. In other work, a non-destructive and versatile microfluidic-based Fabry–Pérot (FP) gas detector (as shown in Figure 17a) was created and investigated by researchers [220]. The detector underwent testing using different analytes possessing diverse chemical and physical properties, as well as various concentrations. The experimental outcomes revealed a sensitivity of 812.5 nm/RIU and an LoD of 1.2 × 10−6 RIU for the detector [220].
PhC gas sensors have gained significant attention in recent years due to their high sensitivity and selectivity in detecting various gases [221,222,223,224,225,226]. These sensors are based on the principle of manipulating light propagation through a periodic arrangement of materials with different refractive indices. An investigation was conducted by researchers [227] on an air-slot photonic crystal cavity to achieve high-precision refractive index sensing. The cavity exhibited a high Q-factor of approximately 2.6 × 104, and a significant overlap was observed between the resonant mode and the hollow core region. The experimental results demonstrated a sensitivity of 510 nm/RIU and an LoD below 1 × 10−5 RIU [227]. A series of Ln slot PhC microcavities were proposed and experimentally demonstrated (in [228]) for their application as refractive index gas sensors, as shown in Figure 17b. The cavities, composed of a Si slab triangular PhC with n holes replaced by slots, showed exponential increases in quality factor with increasing n. An L9 slot PhC microcavity achieved a high-quality factor exceeding 30,000, a sensitivity of 421 nm/RIU, and an LoD below 1 × 10−5 RIU [228].
Figure 17. (a) Three-dimensional diagram of the microfluidic-based FP gas detector [220]. (b) Top view of L9 slot PhC cavity [228].
Figure 17. (a) Three-dimensional diagram of the microfluidic-based FP gas detector [220]. (b) Top view of L9 slot PhC cavity [228].
Micromachines 14 01637 g017
A comprehensive comparison and summary of the performance of various optical gas sensors can be found in Table 4. By examining the data presented in Table 4, readers can gain insights into the strengths and limitations of each sensor type, aiding in the selection and implementation of appropriate optical gas sensing technologies for specific applications.

6. Concluding Remarks

Si photonics is an advancing field of technology that combines optics and electronics using Si as the material platform. It focuses on the development of integrated photonic circuits that manipulate and control light signals, such as how electronic circuits manipulate electrical signals. Si photonics holds tremendous potential to revolutionize telecommunication by enabling faster, more efficient, and cost-effective data transmission and communication systems. As this technology continues to evolve, it is expected to play an increasingly crucial role in shaping the future of telecommunication networks. One of the primary applications of Si photonics in telecommunications is for high-speed, energy-efficient optical interconnects. These interconnects are used to transfer data between different components within data centers and supercomputers. Si photonics allows for the integration of high-performance optical transmitters, receivers, and WGs on a single chip, enabling faster data transfer rates and reducing power consumption compared with traditional copper-based interconnectors. Si photonics can be used to create optical amplifiers and regenerators that boost the optical signal’s strength, compensating for signal losses in long-distance optical communication. These components help maintain the quality and reach of optical signals in telecommunication networks.
Si photonics enables the development of compact and cost-effective optical transceivers used in fiberoptic communication systems. These transceivers integrate lasers, modulators, detectors, and other optical components on a Si chip. They are used for data transmission over long distances in optical networks, such as metropolitan area networks (MANs) and long-haul communications. Moreover, Si photonics provides the means to create compact and efficient wavelength division multiplexing devices, enabling higher data transmission capacity and increasing network bandwidth. Si photonic switches are also used in optical networks to route data packets efficiently. They can be reconfigured rapidly to establish various communication paths, enabling dynamic and flexible network architectures. Si photonics’ fast response times and low power consumption make it ideal for next-generation optical switches.
Additionally, Si photonic sensors leverage the unique properties of Si to interact with light and enable highly sensitive and selective detection. These sensing devices find applications in diverse fields, including environmental monitoring, healthcare, industrial process control, and more. Gas sensing devices based on the Si platform can detect and quantify the presence of specific gases in the environment. As mentioned earlier, the interaction between gas molecules and light in Si WGs can lead to changes in light properties, enabling the detection of various gases with high sensitivity. Moreover, Si photonics biosensors are designed to detect and analyze biological molecules or entities. By functionalizing the Si surface with specific biomolecule receptors (e.g., antibodies or DNA probes), these sensors can detect the binding of target biomolecules, such as proteins or DNA, leading to measurable changes in the transmitted light. Si photonics chemical sensors can also be used to detect and quantify the presence of certain chemicals or analytes. They are widely employed in environmental monitoring and industrial safety applications. The advantages of Si photonics sensors include high sensitivity, label-free detection (in the case of biosensors), compact size, and the potential for integration with existing Si-based electronic and photonic systems.
We believe that Si photonics sensors are still an active area of research and development and that ongoing advancements in materials, fabrication techniques, and device design are continuously improving their performance and expanding their range of applications. As the technology matures, Si photonics sensors are expected to play an increasingly significant role in diverse sensing applications, contributing to advancements in fields such as environmental monitoring, healthcare, and beyond.

Author Contributions

Conceptualization, M.A.B. and M.S.; methodology, M.A.B. and M.S.; software, M.A.B.; validation, M.A.B., M.S. and R.P.; formal analysis, M.S.; investigation, M.S.; resources, R.P.; data curation, M.S.; writing—original draft preparation, M.S.; writing—review and editing, M.A.B.; visualization, M.S.; supervision, M.A.B. and R.P.; project administration, R.P.; funding acquisition, M.A.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge the equal contribution of all the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yamada, K.; Tsuchizawa, T.; Nishi, H.; Kou, R.; Hiraki, T.; Takeda, K.; Fukuda, H.; Ishikawa, Y.; Wada, K.; Yamamoto, T. High-performance silicon photonics technology for telecommunications applications. Sci. Technol. Adv. Mater. 2014, 15, 024603. [Google Scholar] [CrossRef]
  2. Sato, K.I. Design and Performance of Large Port Count Optical Switches for Intra Data Centre Application. In Proceedings of the 2020 22nd International Conference on Transparent Optical Networks (ICTON), Bari, Italy, 19–23 July 2020; pp. 1–4. [Google Scholar]
  3. Roelkens, G.; Abassi, A.; Cardile, P.; Dave, U.; De Groote, A.; De Koninck, Y.; Dhoore, S.; Fu, X.; Gassenq, A.; Hattasan, N.; et al. III-V-on-Silicon Photonic Devices for Optical Communication and Sensing. Photonics 2015, 2, 969–1004. [Google Scholar] [CrossRef]
  4. Kazanskiy, N.L.; Khonina, S.N.; Butt, M.A. Advancement in Silicon Integrated Photonics Technologies for Sensing Applications in Near-Infrared and Mid-Infrared Region: A Review. Photonics 2022, 9, 331. [Google Scholar] [CrossRef]
  5. Sajan, S.C.; Singh, A.; Sharma, P.K.; Kumar, S. Silicon Photonics Biosensors for Cancer Cells Detection—A Review. IEEE Sens. J. 2023, 23, 3366–3377. [Google Scholar] [CrossRef]
  6. Khonina, S.N.; Kazanskiy, N.L.; Butt, M.A. Spectral characteristics of broad band-rejection filter based on Bragg grating, one-dimensional photonic crystal, and subwavelength grating waveguide—IOPscience. Phys. Scr. 2021, 96, 055505. Available online: https://iopscience.iop.org/article/10.1088/1402-4896/abe6be (accessed on 18 July 2023). [CrossRef]
  7. Butt, M.A.; Kazanskiy, N.L.; Khonina, S.N. Advances in Waveguide Bragg Grating Structures, Platforms, and Applications: An Up-to-Date Appraisal. Biosensors 2022, 12, 497. [Google Scholar] [CrossRef]
  8. Butt, M.A.; Khonina, S.N.; Kazanskiy, N.L. A T-shaped 1 × 8 balanced optical power splitter based on 90° bend asymmetric vertical slot waveguides. Laser Phys. 2019, 29, 046207. [Google Scholar] [CrossRef]
  9. Butt, M.A.; Khonina, S.N.; Kazanskiy, N.L. Ultrashort inverted tapered silicon ridge-to-slot waveguide coupler at 1.55 µm and 3.392 µm wavelength. Appl. Opt. 2020, 59, 7821–7828. [Google Scholar] [CrossRef]
  10. Melchior, H. Integrated Photonics Research (1998). In Indium Phosphide Photonic Waveguide Devices and Their Fiber Pigtailing; Optica Publishing Group: British, CO, Canada, 1998; p. IMA2. Available online: https://opg.optica.org/abstract.cfm?uri=IPR-1998-IMA2 (accessed on 18 July 2023).
  11. Wang, J.; Santamato, A.; Jiang, P.; Bonneau, D.; Engin, E.; Silverstone, J.W.; Lermer, M.; Beetz, J.; Kamp, M.; Höfling, S.; et al. Gallium arsenide (GaAs) quantum photonic waveguide circuits. Opt. Commun. 2014, 327, 49–55. [Google Scholar] [CrossRef]
  12. Al-Douri, Y. Optical Waveguide of Lithium Niobate Nanophotonics. In Integrated Nanophotonics; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2023; pp. 277–312. [Google Scholar]
  13. Chen, Q.; Zhu, Y.; Wu, D.; Li, T.; Li, Z.; Lu, C.; Chiang, K.S.; Zhang, X. Electrically generated optical waveguide in a lithium-niobate thin film. Opt. Express 2020, 28, 29895–29903. [Google Scholar] [CrossRef]
  14. Butt, M.A.; Solé, R.; Pujol, M.C.; Ródenas, A.; Lifante, G.; Choudhary, A.; Murugan, G.S.; Shepherd, D.P.; Wilkinson, J.S.; Aguiló, M.; et al. Fabrication of Y-Splitters and Mach–Zehnder Structures on (Yb,Nb):RbTiOPO4/RbTiOPO4 Epitaxial Layers by Reactive Ion Etching. J. Light. Technol. 2015, 33, 1863–1871. [Google Scholar] [CrossRef]
  15. Butt, M.A.; Nguyen, H.D.; Ródenas, A.; Romero, C.; Moreno, P.; de Aldana, J.R.V.; Aguiló, M.; Solé, R.M.; Pujol, M.C.; Díaz, F. Low-repetition rate femtosecond laser writing of optical waveguides in KTP crystals: Analysis of anisotropic refractive index changes. Opt. Express 2015, 23, 15343–15355. [Google Scholar] [CrossRef] [PubMed]
  16. Butt, M.A. Integrated Optics: Platforms and Fabrication Methods. Encyclopedia 2023, 3, 824–838. [Google Scholar] [CrossRef]
  17. Fang, Z.; Zhao, C.Z. Recent progress in silicon photonics: A review. Int. Sch. Res. Not. 2012, 2012, 428690. [Google Scholar] [CrossRef]
  18. Soref, R. The past, present, and future of silicon photonics. IEEE J. Sel. Top. Quantum Electron. 2006, 12, 1678–1687. [Google Scholar] [CrossRef]
  19. Reed, G.T.; Mashanovich, G.; Gardes, F.Y.; Thomson, D.J. Silicon optical modulators. Nat. Photonics 2010, 4, 518–526. [Google Scholar] [CrossRef]
  20. Miller, S.E. Integrated optics: An introduction. Bell Syst. Tech. J. 1969, 48, 2059–2069. [Google Scholar] [CrossRef]
  21. Gehler, J.; Bräuer, A.; Karthe, W.; Jäger, M. Antiresonant reflecting optical waveguides in strip configuration. Appl. Phys. Lett. 1994, 64, 276–278. [Google Scholar] [CrossRef]
  22. Aarnio, J.; Heimala, P.; Del Giudice, M.; Bruno, F. Birefringence control and dispersion characteristics of silicon oxynitride optical waveguides. Electron. Lett. 1991, 25, 2317–2318. [Google Scholar] [CrossRef]
  23. Zurhelle, D.; Schmidt, J.P.; Hoffmann, R.; Sander, D.; Mueller, J. Coupling structures for active and passive integrated optoelectronic components and circuits on silicon. In Proceedings of the Optoelectronic Integrated Circuit Materials, Physics, and Devices, San Jose, CA, USA, 6–9 February 1995; SPIE: Bellingham, WA, USA, 1995; Volume 2397, pp. 666–677. [Google Scholar]
  24. Gleine, W.; Muller, J. Laser trimming of SiON components for integrated optics. J. Light. Technol. 1991, 9, 1626–1629. [Google Scholar] [CrossRef]
  25. Müller, J.; Dahms, U.P.; Mahnke, M.; Wunderlich, S. Refractive index relaxation in PECVD-and LPCVD-SiON-waveguides on silicon substrates. In Proceedings of the Conference on Integrated Optics, Delft, The Netherlands, 3–6 April 1995; pp. 303–308. [Google Scholar]
  26. Wunderlich, S.; Schmidt, J.P.; Müller, J. Integration of SiON waveguides and photodiodes on silicon substrates. Appl. Opt. 1992, 31, 4186–4189. [Google Scholar] [CrossRef] [PubMed]
  27. Tu, Y.K.; Chou, J.C.; Cheng, S.P. Single-mode SiON/SiO2/Si optical waveguides prepared by plasma-enhanced chemical vapor deposition. Fiber Integr. Opt. 1995, 14, 133–139. [Google Scholar] [CrossRef]
  28. Henry, C.H.; Blonder, G.E.; Kazarinov, R.F. Glass waveguides on silicon for hybrid optical packaging. J. Light. Technol. 1989, 7, 1530–1539. [Google Scholar] [CrossRef]
  29. Hickernell, F.S.; Seaton, C.T. Channelized optical waveguides on silicon. Integr. Packag. Optoelectron. Devices 1987, 703, 164–174. [Google Scholar]
  30. Nagata, T.; Tanaka, T.; Miyake, K.; Kurotaki, H.; Yokoyama, S.Y.S.; Koyanagi, M.K.M. Micron-size optical waveguide for optoelectronic integrated circuit. Jpn. J. Appl. Phys. 1994, 33, 822. [Google Scholar] [CrossRef]
  31. Soref, R.A.; Schmidtchen, J.; Petermann, K. Large single-mode rib waveguides in GeSi-Si and Si-on-SiO/sub 2. IEEE J. Quantum Electron. 1991, 27, 1971–1974. [Google Scholar] [CrossRef]
  32. Splett, A.O.; Schmidtchen, J.; Schueppert, B.; Petermann, K. Integrated optical channel waveguides in silicon using SiGe alloys. In Proceedings of the Physical Concepts of Materials for Novel Optoelectronic Device Applications II: Device Physics and Applications, Aachen, Germany, 28 October–2 November 1991; SPIE: Bellingham, WA, USA, 1991; Volume 1362, pp. 827–833. [Google Scholar]
  33. Namavar, F.; Soref, R.A. Optical waveguiding in Si/Si1− x Ge x/Si heterostructures. J. Appl. Phys. 1991, 70, 3370–3372. [Google Scholar] [CrossRef]
  34. Weiss, B.L.; Yang, Z.; Namavar, F. Wavelength dependent propagation loss characteristics of SiGe/Si planar waveguides. Electron. Lett. 1992, 28, 2218–2220. [Google Scholar] [CrossRef]
  35. Grant, M.F. Integrated optical waveguide devices on silicon for optical communications. In Proceedings of the IEE Colloquium on Planar Silicon Hybrid Optoelectronics (Digest No. 1994/198), London, UK, 24 October 1994; IET: Stevenage, UK, 1994; pp. 1/1–110. [Google Scholar]
  36. Najafi, S.I.; Honkanen, S.; Tervonen, A. Recent progress in glass integrated optical circuits. Integr. Opt. Microstruct. II 1994, 2291, 6–12. [Google Scholar]
  37. Habara, K.; Matsunaga, T. Silica glass waveguide planar lightwave circuit applications in photonic switching. In Proceedings of the Photonic Switching, Minsk, Belarus, 1–3 July 1992; SPIE: Bellingham, WA, USA, 1993; Volume 1807, pp. 418–421. [Google Scholar]
  38. Tang, X.; Wongchotigul, K.; Spencer, M.G. Optical waveguide formed by cubic silicon carbide on sapphire substrates. Appl. Phys. Lett. 1991, 58, 917–918. [Google Scholar] [CrossRef]
  39. Jackson, S.M.; Reed, G.T.; Reeson, K.J. Wave-Guiding in epitaxial 3c-Silicon carbide on silicon. Electron. Lett. 1995, 31, 1438–1439. [Google Scholar] [CrossRef]
  40. Liu, Y.M.; Prucnal, P.R. Low-loss silicon carbide optical waveguides for silicon-based optoelectronic devices. IEEE Photonics Technol. Lett. 1993, 5, 704–707. [Google Scholar] [CrossRef]
  41. Iyer, S.S.; Xie, Y.-H. Light Emission from Silicon. Science 1993, 260, 40–46. [Google Scholar] [CrossRef]
  42. Soref, R.A. Silicon-based optoelectronics. Proc. IEEE 1993, 81, 1687–1706. [Google Scholar] [CrossRef]
  43. Huang, Y.; Zhao, Q.; Kamyab, L.; Rostami, A.; Capolino, F.; Boyraz, O. Sub-micron silicon nitride waveguide fabrication using conventional optical lithography. Opt. Express 2015, 23, 6780–6786. [Google Scholar] [CrossRef]
  44. Okazaki, S. High resolution optical lithography or high throughput electron beam lithography: The technical struggle from the micro to the nano-fabrication evolution. Microelectron. Eng. 2015, 133, 23–35. [Google Scholar] [CrossRef]
  45. Zheng, Y.; Gao, P.; Tang, X.; Liu, J.; Duan, J. Effects of electron beam lithography process parameters on structure of silicon optical waveguide based on SOI. J. Cent. South Univ. 2022, 29, 3335–3345. [Google Scholar] [CrossRef]
  46. Broers, A.N.; Hoole, A.C.F.; Ryan, J.M. Electron beam lithography—Resolution limits. Microelectron. Eng. 1996, 32, 131–142. [Google Scholar] [CrossRef]
  47. Bruinink, C.M.; Burresi, M.; de Boer, M.J.; Segerink, F.B.; Jansen, H.V.; Berenschot, E.; Reinhoudt, D.N.; Huskens, J.; Kuipers, L. Nanoimprint Lithography for Nanophotonics in Silicon. Nano Lett. 2008, 8, 2872–2877. [Google Scholar] [CrossRef]
  48. Shneidman, A.V.; Becker, K.P.; Lukas, M.A.; Torgerson, N.; Wang, C.; Reshef, O.; Burek, M.J.; Paul, K.; McLellan, J.; Lončar, M. All-Polymer Integrated Optical Resonators by Roll-to-Roll Nanoimprint Lithography. ACS Photonics 2018, 5, 1839–1845. [Google Scholar] [CrossRef]
  49. Hiscocks, M.P.; Ganesan, K.; Gibson, B.C.; Huntington, S.T.; Ladouceur, F.; Prawer, S. Diamond waveguides fabricated by reactive ion etching. Opt. Express 2008, 16, 19512–19519. [Google Scholar] [CrossRef] [PubMed]
  50. Kim, T.; Lee, J. Optimization of deep reactive ion etching for microscale silicon hole arrays with high aspect ratio. Micro Nano Syst. Lett. 2022, 10, 12. [Google Scholar] [CrossRef]
  51. Bauhuber, M.; Mikrievskij, A.; Lechner, A. Isotropic wet chemical etching of deep channels with optical surface quality in silicon with HNA based etching solutions. Mater. Sci. Semicond. Process. 2013, 16, 1428–1433. [Google Scholar] [CrossRef]
  52. Butt, M.A.; Kozłowski, Ł.; Piramidowicz, R. Numerical scrutiny of a silica-titania-based reverse rib waveguide with vertical and rounded sidewalls. Appl. Opt. 2023, 62, 1296–1302. [Google Scholar] [CrossRef] [PubMed]
  53. Li, W.; Luo, Y.; Xiong, B.; Sun, C.; Wang, L.; Wang, J.; Han, Y.; Yan, J.; Wei, T.; Lu, H. Fabrication of GaN-based ridge waveguides with very smooth and vertical sidewalls by combined plasma dry etching and wet chemical etching. Phys. Status Solidi A 2015, 212, 2341–2344. [Google Scholar] [CrossRef]
  54. Butt, M.A.; Khonina, S.N.; Kazanskiy, N.L. 2D-Photonic crystal heterostructures for the realization of compact photonic devices. Photonics Nanostructures Fundam. Appl. 2021, 44, 100903. [Google Scholar] [CrossRef]
  55. Butt, M.A.; Kazanskiy, N.L. Two-dimensional photonic crystal heterostructure for light steering and TM-polarization maintaining applications—IOPscience. Laser Phys. 2021, 31, 036201. Available online: https://iopscience.iop.org/article/10.1088/1555-6611/abd8ca (accessed on 18 July 2023). [CrossRef]
  56. Quack, N.; Takabayashi, A.Y.; Sattari, H.; Edinger, P.; Jo, G.; Bleiker, S.J.; Errando-Herranz, C.; Gylfason, K.B.; Niklaus, F.; Khan, U.; et al. Integrated silicon photonic MEMS. Microsyst. Nanoeng. 2023, 9, 27. [Google Scholar] [CrossRef]
  57. Yue, H.; Chen, K.; Chu, T. Ultrahigh-linearity dual-drive scheme using a single silicon modulator. Opt. Lett. 2023, 48, 2995–2998. Available online: https://opg.optica.org/ol/abstract.cfm?uri=ol-48-11-2995 (accessed on 18 July 2023). [CrossRef]
  58. 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]
  59. Yue, G.; Xing, Z.; Hu, H.; Cheng, Z.; Lu, G.-W.; Liu, T. Graphene-based dual-mode modulators. Opt. Express 2020, 28, 18456–18471. [Google Scholar] [CrossRef]
  60. Phatak, A.; Cheng, Z.; Qin, C.; Goda, K. Design of electro-optic modulators based on graphene-on-silicon slot waveguides. Opt. Lett. 2016, 41, 2501–2504. [Google Scholar] [CrossRef] [PubMed]
  61. Sorianello, V.; Midrio, M.; Contestabile, G.; Asselberghs, I.; Van Campenhout, J.; Huyghebaert, C.; Goykhman, I.; Ott, A.K.; Ferrari, A.C.; Romagnoli, M. Graphene–silicon phase modulators with gigahertz bandwidth. Nat. Photonics 2018, 12, 40–44. [Google Scholar] [CrossRef]
  62. Soref, R.; Bennett, B. Electrooptical effects in silicon. IEEE J. Quantum Electron. 1987, 23, 123–129. [Google Scholar] [CrossRef]
  63. Witzens, J. High-Speed Silicon Photonics Modulators. Proc. IEEE 2018, 106, 2158–2182. [Google Scholar] [CrossRef]
  64. Pant, B.; Zhang, W.; Ebert, M.; Yan, X.; Du, H.; Banakar, M.; Tran, D.T.; Qi, Y.; Rowe, D.; Jeyaselvan, V.; et al. Study into the spread of heat from thermo-optic silicon photonic elements. Opt. Express 2021, 29, 36461–36468. [Google Scholar] [CrossRef]
  65. Fasih, M.; Yoon, H.; Kwan, N.; Kim, J.; Yoon, J.; Sikder, R.I.; Kurt, H.; Park, H.H. Optical Sideband Modulation in Silicon Photonics Platform Using Mach-Zehnder Interferometers. In Proceedings of the 2022 24th International Conference on Advanced Communication Technology (ICACT), PyeongChang, Republic of Korea, 13–16 February 2022; pp. 288–292. [Google Scholar]
  66. Gui, Y.; Nouri, B.M.; Miscuglio, M.; Amin, R.; Wang, H.; Khurgin, J.B.; Dalir, H.; Sorger, V.J. 100 GHz micrometer-compact broadband monolithic ITO Mach–Zehnder interferometer modulator enabling 3500 times higher packing density. Nanophotonics 2022, 11, 4001–4009. [Google Scholar] [CrossRef]
  67. Xu, Q.; Schmidt, B.; Pradhan, S.; Lipson, M. Micrometre-scale silicon electro-optic modulator. Nature 2005, 435, 325–327. [Google Scholar] [CrossRef]
  68. Green, W.M.J.; Rooks, M.J.; Sekaric, L.; Vlasov, Y.A. Ultra-compact, low RF power, 10 Gb/s silicon Mach-Zehnder modulator. Opt. Express 2007, 15, 17106–17113. [Google Scholar] [CrossRef]
  69. Jiang, W.; Gu, L.; Chen, X.; Chen, R.T. Photonic crystal waveguide modulators for silicon photonics: Device physics and some recent progress. Solid State Electronics 2007, 51, 1278–1286. [Google Scholar] [CrossRef]
  70. Wang, C.; Zhang, M.; Chen, X.; Bertrand, M.; Shams-Ansari, A.; Chandrasekhar, S.; Winzer, P.; Lončar, M. Integrated lithium niobate electro-optic modulators operating at CMOS-compatible voltages. Nature 2018, 562, 101–104. [Google Scholar] [CrossRef] [PubMed]
  71. Sato, H.; Miura, H.; Qiu, F.; Spring, A.M.; Kashino, T.; Kikuchi, T.; Ozawa, M.; Nawata, H.; Odoi, K.; Yokoyama, S. Low driving voltage Mach-Zehnder interference modulator constructed from an electro-optic polymer on ultra-thin silicon with a broadband operation. Opt. Express 2017, 25, 768–775. [Google Scholar] [CrossRef]
  72. Kieninger, C.; Füllner, C.; Zwickel, H.; Kutuvantavida, Y.; Kemal, J.N.; Eschenbaum, C.; Elder, D.L.; Dalton, L.R.; Freude, W.; Randel, S. Silicon-organic hybrid (SOH) Mach-Zehnder modulators for 100 GBd PAM4 signaling with sub-1 dB phase-shifter loss. Opt. Express 2020, 28, 24693–24707. [Google Scholar] [CrossRef] [PubMed]
  73. Ayoub, A.B.; Swillam, M.A. Optical modulator using ultra-thin silicon waveguide in SOI hybrid technology. Opt. Quantum Electron. 2022, 54, 181. [Google Scholar] [CrossRef]
  74. Xiao, L.; Wu, J.-W. Porous silicon based all-optical modulator using asymmetrical Mach–Zehnder interferometer configuration. Opt. Commun. 2015, 338, 246–252. [Google Scholar] [CrossRef]
  75. Zhang, W.; Ebert, M.; Li, K.; Chen, B.; Yan, X.; Du, H.; Banakar, M.; Tran, D.T.; Littlejohns, C.G.; Scofield, A.; et al. Harnessing plasma absorption in silicon MOS ring modulators. Nat. Photonics 2023, 17, 273–279. [Google Scholar] [CrossRef]
  76. Qiu, F.; Han, Y. Electro-optic polymer ring resonator modulators [Invited]. Chin. Opt. Lett. 2021, 19, 041301. [Google Scholar] [CrossRef]
  77. Manipatruni, S.; Xu, Q.; Schmidt, B.; Shakya, J.; Lipson, M. High Speed Carrier Injection 18 Gb/s Silicon Micro-ring Electro-optic Modulator. In Proceedings of the LEOS 2007—IEEE Lasers and Electro-Optics Society Annual Meeting Conference Proceedings, Lake Buena Vista, FL, USA, 21–25 October 2007; pp. 537–538. [Google Scholar]
  78. Jayatilleka, H.; Frish, H.; Kumar, R.; Heck, J.; Ma, C.; Sakib, M.N.; Huang, D.; Rong, H. Post-Fabrication Trimming of Silicon Photonic Ring Resonators at Wafer-Scale. J. Light. Technol. 2021, 39, 5083–5088. [Google Scholar] [CrossRef]
  79. Chen, L.; Preston, K.; Manipatruni, S.; Lipson, M. Integrated GHz silicon photonic interconnect with micrometer-scale modulators and detectors. Opt. Express 2009, 17, 15248–15256. [Google Scholar] [CrossRef]
  80. Zheng, X.; Chang, E.; Amberg, P.; Shubin, I.; Lexau, J.; Liu, F.; Thacker, H.; Djordjevic, S.S.; Lin, S.; Luo, Y.; et al. A high-speed, tunable silicon photonic ring modulator integrated with ultra-efficient active wavelength control. Opt. Express 2014, 22, 12628–12633. [Google Scholar] [CrossRef]
  81. Li, J.; Li, G.; Zheng, X.; Raj, K.; Krishnamoorthy, A.V.; Buckwalter, J.F. A 25-Gb/s Monolithic Optical Transmitter with Micro-Ring Modulator in 130-nm SoI CMOS. IEEE Photonics Technol. Lett. 2013, 25, 1901–1903. [Google Scholar] [CrossRef]
  82. Pantouvaki, M.; Verheyen, P.; De Coster, J.; Lepage, G.; Absil, P.; Van Campenhout, J. 56Gb/s ring modulator on a 300mm silicon photonics platform. In Proceedings of the 2015 European Conference on Optical Communication (ECOC), Valencia, Spain, 27 September–1 October 2015; pp. 1–3. [Google Scholar]
  83. Ziebell, M.; Marris-Morini, D.; Rasigade, G.; Crozat, P.; Fédéli, J.-M.; Grosse, P.; Cassan, E.; Vivien, L. Ten Gbit/s ring resonator silicon modulator based on interdigitated PN junctions. Opt. Express 2011, 19, 14690–14695. [Google Scholar] [CrossRef] [PubMed]
  84. Moazeni, S.; Lin, S.; Wade, M.; Alloatti, L.; Ram, R.J.; Popović, M.; Stojanović, V. A 40-Gb/s PAM-4 Transmitter Based on a Ring-Resonator Optical DAC in 45-nm SOI CMOS. IEEE J. Solid-State Circuits 2017, 52, 3503–3516. [Google Scholar] [CrossRef]
  85. Li, G.; Krishnamoorthy, A.V.; Shubin, I.; Yao, J.; Luo, Y.; Thacker, H.; Zheng, X.; Raj, K.; Cunningham, J.E. Ring Resonator Modulators in Silicon for Interchip Photonic Links. IEEE J. Sel. Top. Quantum Electron. 2013, 19, 95–113. [Google Scholar] [CrossRef]
  86. Hai, M.S.; Fard, M.M.P.; Liboiron-Ladouceur, O. A low-voltage PAM-4 SOI ring-based modulator. In Proceedings of the 2014 IEEE Photonics Conference, San Diego, CA, USA, 12–16 October 2014; pp. 194–195. [Google Scholar]
  87. Dong, P.; Liao, S.; Feng, D.; Liang, H.; Zheng, D.; Shafiiha, R.; Kung, C.-C.; Qian, W.; Li, G.; Zheng, X. Low V pp, ultralow-energy, compact, high-speed silicon electro-optic modulator. Opt. Express 2009, 17, 22484–22490. [Google Scholar] [CrossRef]
  88. Hu, Y.; Xiao, X.; Xu, H.; Li, X.; Xiong, K.; Li, Z.; Chu, T.; Yu, Y.; Yu, J. High-speed silicon modulator based on cascaded microring resonators. Opt. Express 2012, 20, 15079–15085. [Google Scholar] [CrossRef]
  89. Demirtzioglou, I.; Lacava, C.; Bottrill, K.R.H.; Thomson, D.J.; Reed, G.T.; Richardson, D.J.; Petropoulos, P. Frequency comb generation in a silicon ring resonator modulator. Opt. Express 2018, 26, 790–796. [Google Scholar] [CrossRef]
  90. Khonina, S.N.; Kazanskiy, N.L.; Butt, M.A.; Karpeev, S.V. Optical multiplexing techniques and their marriage for on-chip and optical fiber communication: A review. Opto-Electron. Adv. 2022, 5, 210127. [Google Scholar] [CrossRef]
  91. Wang, Q. Investigation progress on key photonic integration for application in optical communication network. Sci. China Ser. F 2003, 46, 60–66. [Google Scholar] [CrossRef]
  92. Irace, A.; Breglio, G. Silicon-based optoelectronic filters based on a Bragg grating and P-i-N diode for DWDM optical networks. In Proceedings of the Laser Diodes, Optoelectronic Devices, and Heterogenous Integration, Bruges, Belgium, 28 October–1 November 2002; Volume 4947, pp. 68–73. Available online: https://www.spiedigitallibrary.org/conference-proceedings-of-spie/4947.toc (accessed on 28 October 2002).
  93. Fukazawa, T.; Ohno, F.; Baba, T. Very Compact Arrayed-Waveguide-Grating Demultiplexer Using Si Photonic Wire Waveguides. Jpn. J. Appl. Phys. 2004, 43, L673. [Google Scholar] [CrossRef]
  94. Brouckaert, J.; Bogaerts, W.; Dumon, P.; Van Thourhout, D.; Baets, R. Planar Concave Grating Demultiplexer Fabricated on a Nanophotonic Silicon-on-Insulator Platform. J. Light. Technol. 2007, 25, 1269–1275. [Google Scholar] [CrossRef]
  95. Kim, D.W.; Barkai, A.; Jones, R.; Elek, N.; Nguyen, H.; Liu, A. Silicon-on-insulator eight-channel optical multiplexer based on a cascade of asymmetric Mach–Zehnder interferometers. Opt. Lett. 2008, 33, 530–532. [Google Scholar] [CrossRef] [PubMed]
  96. Horst, F.; Green, W.M.J.; Offrein, B.J.; Vlasov, Y.A. Silicon-on-Insulator Echelle Grating WDM Demultiplexers With Two Stigmatic Points. IEEE Photonics Technol. Lett. 2009, 21, 1743–1745. [Google Scholar] [CrossRef]
  97. Fiorentino, M.; Peng, Z.; Binkert, N.; Beausoleil, R.G. Devices and architectures for large scale integrated silicon photonics circuits. In Proceedings of the IEEE Winter Topicals 2011, Keystone, CO, USA, 10–12 January 2011; pp. 131–132. [Google Scholar]
  98. Feng, D.; Qian, W.; Liang, H.; Luff, B.J.; Asghari, M. High-Speed Receiver Technology on the SOI Platform. IEEE J. Sel. Top. Quantum Electron. 2013, 19, 3800108. [Google Scholar] [CrossRef]
  99. Horst, F.; Green, W.M.J.; Assefa, S.; Shank, S.M.; Vlasov, Y.A.; Offrein, B.J. Cascaded Mach-Zehnder wavelength filters in silicon photonics for low loss and flat pass-band WDM (de-)multiplexing. Opt. Express 2013, 21, 11652–11658. [Google Scholar] [CrossRef]
  100. Okamoto, K. Wavelength-Division-Multiplexing Devices in Thin SOI: Advances and Prospects. IEEE J. Sel. Top. Quantum Electron. 2014, 20, 248–257. [Google Scholar] [CrossRef]
  101. Wang, J.; Chen, S.; Dai, D. Silicon hybrid demultiplexer with 64 channels for wavelength/mode-division multiplexed on-chip optical interconnects. Opt. Lett. 2014, 39, 6993–6996. [Google Scholar] [CrossRef]
  102. Chen, S.; Shi, Y.; He, S.; Dai, D. Compact monolithically-integrated hybrid (de)multiplexer based on silicon-on-insulator nanowires for PDM-WDM systems. Opt. Express 2015, 23, 12840–12849. [Google Scholar] [CrossRef]
  103. Castellan, C.; Tondini, S.; Mancinelli, M.; Kopp, C.; Pavesi, L. Low crosstalk silicon arrayed waveguide gratings for on-chip optical multiplexing. In Proceedings of the Silicon Photonics: From Fundamental Research to Manufacturing, Strasbourg, France, 23–26 April 2018; Volume 10686, p. 106860V. [Google Scholar]
  104. Jeong, S.H.; Onawa, Y.; Shimura, D.; Okayama, H.; Aoki, T.; Yaegashi, H.; Horikawa, T.; Nakamura, T. Polarization Diversified 16λ Demultiplexer Based on Silicon Wire Delayed Interferometers and Arrayed Waveguide Gratings. J. Light. Technol. 2020, 38, 2680–2687. [Google Scholar] [CrossRef]
  105. Hammood, M.; Mistry, A.; Yun, H.; Ma, M.; Chrostowski, L.; Jaeger, N.A.F. Four-channel, Silicon Photonic, Wavelength Multiplexer-Demultiplexer With High Channel Isolations. In Proceedings of the Optical Fiber Communication Conference, San Diego, CA, USA, 8–12 March 2020; Optica Publishing Group: San Diego, CA, USA, 2020; p. M3F.5. [Google Scholar]
  106. Shen, X.; Chen, B.; Zhu, Y.; Shi, W. Silicon photonic integrated circuits and its application in data center. In Proceedings of the Seventh Symposium on Novel Photoelectronic Detection Technology and Applications, Kunming, China, 5–7 November 2020; SPIE: Bellingham, WA, USA, 2021; Volume 11763, p. 1176380. [Google Scholar]
  107. Wang, Y.; Wang, S.; Novick, A.; James, A.; Parsons, R.; Rizzo, A.; Bergman, K. Dispersion-Engineered and Fabrication-Robust SOI Waveguides for Ultra-Broadband DWDM. In Proceedings of the 2023 Optical Fiber Communications Conference and Exhibition (OFC), San Diego, CA, USA, 5–9 March 2023; pp. 1–3. [Google Scholar]
  108. Fu, X.; Cheng, J.; Huang, Q.; Hu, Y.; Xie, W.; Tassaert, M.; Verbist, J.; Ma, K.; Zhang, J.; Chen, K.; et al. 5 x 20 Gb/s heterogeneously integrated III-V on silicon electro-absorption modulator array with arrayed waveguide grating multiplexer. Opt. Express 2015, 23, 18686–18693. [Google Scholar] [CrossRef]
  109. Trita, A.; Thomas, A.; Rickman, A. CMOS compatible athermal silicon photonic filters based on hydrogenated amorphous silicon. Opt. Express 2022, 30, 19311–19319. [Google Scholar] [CrossRef] [PubMed]
  110. 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] [PubMed]
  111. Jalali, B.; Fathpour, S. Silicon Photonics. J. Light. Technol. 2006, 24, 4600–4615. [Google Scholar] [CrossRef]
  112. Zhang, Y.; Zhang, R.; Zhu, Q.; Yuan, Y.; Su, Y. Architecture and Devices for Silicon Photonic Switching in Wavelength, Polarization and Mode. J. Light. Technol. 2020, 38, 215–225. [Google Scholar] [CrossRef]
  113. Fokine, M.; Nilsson, L.E.; Claesson, Å.; Berlemont, D.; Kjellberg, L.; Krummenacher, L.; Margulis, W. Integrated fiber Mach–Zehnder interferometer for electro-optic switching. Opt. Lett. 2002, 27, 1643–1645. [Google Scholar] [CrossRef]
  114. Brunetti, G.; Marocco, G.; Benedetto, A.D.; Giorgio, A.; Armenise, M.N.; Ciminelli, C. Design of a large bandwidth 2 × 2 interferometric switching cell based on a sub-wavelength grating. J. Opt. 2021, 23, 085801. [Google Scholar] [CrossRef]
  115. Patel, D.; Veerasubramanian, V.; Ghosh, S.; Shi, W.; Samani, A.; Zhong, Q.; Plant, D.V. A 4×4 fully non-blocking switch on SOI based on interferometric thermo-optic phase shifters. In Proceedings of the 2014 Optical Interconnects Conference, San Diego, CA, USA, 4–7 May 2014; pp. 104–105. [Google Scholar]
  116. Padmaraju, K.; Bergman, K. Resolving the thermal challenges for silicon microring resonator devices. Nanophotonics 2014, 3, 269–281. [Google Scholar] [CrossRef]
  117. Nikdast, M.; Nicolescu, G.; Trajkovic, J.; Liboiron-Ladouceur, O. Chip-Scale Silicon Photonic Interconnects: A Formal Study on Fabrication Non-Uniformity. J. Light. Technol. 2016, 34, 3682–3695. [Google Scholar] [CrossRef]
  118. Hsu, W.C.; Nujhat, N.; Kupp, B.; Conley, J.F.; Wang, A.X. On-chip wavelength division multiplexing filters using extremely efficient gate-driven silicon microring resonator array. Sci. Rep. 2023, 13, 5269. [Google Scholar] [CrossRef]
  119. Li, X.; Zhu, Y.; Liu, Z.; Peng, L.; Liu, X.; Niu, C.; Zheng, J.; Zuo, Y.; Cheng, B. 75 GHz germanium waveguide photodetector with 64 Gbps data rates utilizing an inductive-gain-peaking technique. J. Semicond. 2023, 44, 012301. [Google Scholar] [CrossRef]
  120. Suh, D.; Kim, S.; Joo, J.; Kim, G. 36-GHz High-Responsivity Ge Photodetectors Grown by RPCVD. IEEE Photonics Technol. Lett. 2009, 21, 672–674. [Google Scholar] [CrossRef]
  121. Lischke, S.; Peczek, A.; Morgan, J.S.; Sun, K.; Steckler, D.; Yamamoto, Y.; Korndörfer, F.; Mai, C.; Marschmeyer, S.; Fraschke, M.; et al. Ultra-fast germanium photodiode with 3-dB bandwidth of 265 GHz. Nat. Photonics 2021, 15, 925–931. [Google Scholar] [CrossRef]
  122. Wang, H.; Zhang, J.; Zhang, G.; Chen, Y.; Huang, Y.-C.; Gong, X. High-speed and high-responsivity p-i-n waveguide photodetector at a 2 µm wavelength with a Ge0.92Sn0.08/Ge multiple-quantum-well active layer. Opt. Lett. 2021, 46, 2099–2102. [Google Scholar] [CrossRef] [PubMed]
  123. Kim, S.; Bhargava, N.; Gupta, J.; Coppinger, M.; Kolodzey, J. Infrared photoresponse of GeSn/n-Ge heterojunctions grown by molecular beam epitaxy. Opt. Express 2014, 22, 11029–11034. [Google Scholar] [CrossRef]
  124. Gassenq, A.; Gencarelli, F.; Van Campenhout, J.; Shimura, Y.; Loo, R.; Narcy, G.; Vincent, B.; Roelkens, G. GeSn/Ge heterostructure short-wave infrared photodetectors on silicon. Opt. Express 2012, 20, 27297–27303. [Google Scholar] [CrossRef]
  125. Roucka, R.; Mathews, J.; Weng, C.; Beeler, R.; Tolle, J.; Menéndez, J.; Kouvetakis, J. High-Performance Near-IR Photodiodes: A Novel Chemistry-Based Approach to Ge and Ge–Sn Devices Integrated on Silicon. IEEE J. Quantum Electron. 2011, 47, 213–222. [Google Scholar] [CrossRef]
  126. Bartels, A.; Peiner, E.; Tang, G.P.; Klockenbrink, R.; Wehmann, H.H.; Schlachetzki, A. Performance of InGaAs metal-semiconductor-metal photodetectors on Si. IEEE Photonics Technol. Lett. 1996, 8, 670–672. [Google Scholar] [CrossRef]
  127. Chen, G.; Goyvaerts, J.; Kumari, S.; Van Kerrebrouck, J.; Muneeb, M.; Uvin, S.; Yu, Y.; Roelkens, G. Integration of high-speed GaAs metal-semiconductor-metal photodetectors by means of transfer printing for 850 nm wavelength photonic interposers. Opt. Express 2018, 26, 6351–6359. [Google Scholar] [CrossRef]
  128. Knights, A.P.; Bradley, J.D.B.; Gou, S.H.; Jessop, P.E. Silicon-on-insulator waveguide photodetector with self-ion-implantation-engineered-enhanced infrared response. J. Vac. Sci. Technol. A 2006, 24, 783–786. [Google Scholar] [CrossRef]
  129. Ackert, J.J.; Fiorentino, M.; Logan, D.F.; Beausoleil, R.G.; Jessop, P.E.; Knights, A.P. Silicon-on-insulator microring resonator defect-based photodetector with 3.5-GHz bandwidth. J. Nanophotonics 2011, 5, 059507-1–059507-6. [Google Scholar] [CrossRef]
  130. Ahn, D.; Hong, C.; Liu, J.; Giziewicz, W.; Beals, M.; Kimerling, L.C.; Michel, J.; Chen, J.; Kärtner, F.X. High performance, waveguide integrated Ge photodetectors. Opt. Express 2007, 15, 3916–3921. [Google Scholar] [CrossRef] [PubMed]
  131. Chen, H.; Galili, M.; Verheyen, P.; De Heyn, P.; Lepage, G.; De Coster, J.; Balakrishnan, S.; Absil, P.; Oxenlowe, L.; Van Campenhout, J.; et al. 100-Gbps RZ Data Reception in 67-GHz Si-Contacted Germanium Waveguide p-i-n Photodetectors. J. Light. Technol. 2017, 35, 722–726. [Google Scholar] [CrossRef]
  132. Li, X.L.; Liu, Z.; Peng, L.Z.; Liu, X.Q.; Wang, N.; Zhao, Y.; Zheng, J.; Zuo, Y.H.; Xue, C.L.; Cheng, B.W. High-Performance Germanium Waveguide Photodetectors on Silicon. Chin. Phys. Lett. 2020, 37, 038503. [Google Scholar] [CrossRef]
  133. Vivien, L.; Osmond, J.; Fédéli, J.M.; Marris-Morini, D.; Crozat, P.; Damlencourt, J.F.; Cassan, E.; Lecunff, Y.; Laval, S. 42 GHz p.i.n Germanium photodetector integrated in a silicon-on-insulator waveguide. Opt. Express 2009, 17, 6252–6257. [Google Scholar] [CrossRef]
  134. Wu, L.; Lv, D.; Zhao, N.; Wang, R.; Wu, A. Research on Germanium Photodetector with Multi-Mode Waveguide Input. Photonics 2023, 10, 455. [Google Scholar] [CrossRef]
  135. Dong, Y.; Wang, W.; Lei, D.; Gong, X.; Zhou, Q.; Lee, S.Y.; Loke, W.K.; Yoon, S.F.; Tok, E.S.; Liang, G.; et al. Suppression of dark current in germanium-tin on silicon p-i-n photodiode by a silicon surface passivation technique. Opt. Express 2015, 23, 18611–18619. [Google Scholar] [CrossRef]
  136. Xu, S.; Huang, Y.C.; Lee, K.H.; Wang, W.; Dong, Y.; Lei, D.; Masudy-Panah, S.; Tan, C.S.; Gong, X.; Yeo, Y.C. GeSn lateral p-i-n photodetector on insulating substrate. Opt. Express 2018, 26, 17312–17321. [Google Scholar] [CrossRef]
  137. Li, X.; Peng, L.; Liu, Z.; Zhou, Z.; Zheng, J.; Xue, C.; Zuo, Y.; Chen, B.; Cheng, B. 30 GHz GeSn photodetector on SOI substrate for 2 µm wavelength application. Photonics Res. 2021, 9, 494–500. [Google Scholar] [CrossRef]
  138. Liu, C.H.; Bansal, R.; Wu, C.W.; Jheng, Y.T.; Chang, G.E. GeSn Waveguide Photodetectors with Vertical p–i–n Heterostructure for Integrated Photonics in the 2 μm Wavelength Band. Adv. Photonics Res. 2022, 3, 2100330. [Google Scholar] [CrossRef]
  139. Xue, Y.; Han, Y.; Tong, Y.; Yan, Z.; Wang, Y.; Zhang, Z.; Tsang, H.K.; Lau, K.M. High-performance III-V photodetectors on a monolithic InP/SOI platform. Optica 2021, 8, 1204–1209. [Google Scholar] [CrossRef]
  140. Wan, Y.; Zhang, Z.; Chao, R.; Norman, J.; Jung, D.; Shang, C.; Li, Q.; Kennedy, M.J.; Liang, D.; Zhang, C. Monolithically integrated InAs/InGaAs quantum dot photodetectors on silicon substrates. Opt. Express 2017, 25, 27715–27723. [Google Scholar] [CrossRef] [PubMed]
  141. Shen, L.; Jiao, Y.; Yao, W.; Cao, Z.; van Engelen, J.P.; Roelkens, G.; Smit, M.K.; van der Tol, J.J.G.M. High-bandwidth uni-traveling carrier waveguide photodetector on an InP-membrane-on-silicon platform. Opt. Express 2016, 24, 8290–8301. [Google Scholar] [CrossRef] [PubMed]
  142. Zhang, J.; De Groote, A.; Abbasi, A.; Loi, R.; O’Callaghan, J.; Corbett, B.; Trindade, A.J.; Bower, C.A.; Roelkens, G. Silicon photonics fiber-to-the-home transceiver array based on transfer-printing-based integration of III-V photodetectors. Opt. Express 2017, 25, 14290–14299. [Google Scholar] [CrossRef] [PubMed]
  143. Sun, K.; Jung, D.; Shang, C.; Liu, A.; Morgan, J.; Zang, J.; Li, Q.; Klamkin, J.; Bowers, J.E.; Beling, A. Low dark current III–V on silicon photodiodes by heteroepitaxy. Opt. Express 2018, 26, 13605–13613. [Google Scholar] [CrossRef]
  144. Wang, R.; Sprengel, S.; Muneeb, M.; Boehm, G.; Baets, R.; Amann, M.-C.; Roelkens, G. 2 μm wavelength range InP-based type-II quantum well photodiodes heterogeneously integrated on silicon photonic integrated circuits. Opt. Express 2015, 23, 26834–26841. [Google Scholar] [CrossRef]
  145. Baehr-Jones, T.; Hochberg, M.; Scherer, A. Photodetection in silicon beyond the band edge with surface states. Opt. Express 2008, 16, 1659–1668. [Google Scholar] [CrossRef]
  146. Goykhman, I.; Desiatov, B.; Khurgin, J.; Shappir, J.; Levy, U. Waveguide based compact silicon Schottky photodetector with enhanced responsivity in the telecom spectral band. Opt. Express 2012, 20, 28594–28602. [Google Scholar] [CrossRef]
  147. Tanabe, T.; Sumikura, H.; Taniyama, H.; Shinya, A.; Notomi, M. All-silicon sub-Gb/s telecom detector with low dark current and high quantum efficiency on chip. Appl. Phys. Lett. 2010, 96, 101103. [Google Scholar] [CrossRef]
  148. Li, Y.; Feng, S.; Zhang, Y.; Poon, A.W. Sub-bandgap linear-absorption-based photodetectors in avalanche mode in PN-diode-integrated silicon microring resonators. Opt. Lett. 2013, 38, 5200–5203. [Google Scholar] [CrossRef]
  149. Li, Y.; Poon, A.W. Active resonance wavelength stabilization for silicon microring resonators with an in-resonator defect-state-absorption-based photodetector. Opt. Express 2015, 23, 360–372. [Google Scholar] [CrossRef]
  150. Geis, M.W.; Spector, S.J.; Grein, M.E.; Yoon, J.U.; Lennon, D.M.; Lyszczarz, T.M. Silicon waveguide infrared photodiodes with >35 GHz bandwidth and phototransistors with 50 AW-1 response. Opt. Express 2009, 17, 5193–5204. [Google Scholar] [CrossRef] [PubMed]
  151. Yuan, Y.; Sorin, W.V.; Liang, D.; Cheung, S.; Peng, Y.; Jain, M.; Huang, Z.; Fiorentino, M.; Beausoleil, R.G. Mechanisms of enhanced sub-bandgap absorption in high-speed all-silicon avalanche photodiodes. Photonics Res. 2023, 11, 337–346. [Google Scholar] [CrossRef]
  152. Conteduca, D.; Arruda, G.S.; Barth, I.; Wang, Y.; Krauss, T.F.; Martins, E.R. Beyond Q: The Importance of the Resonance Amplitude for Photonic Sensors. ACS Photonics 2022, 9, 1757–1763. [Google Scholar] [CrossRef] [PubMed]
  153. Torrijos-Morán, L.; Lisboa, B.D.; Soler, M.; Lechuga, L.M.; García-Rupérez, J. Integrated optical bimodal waveguide biosensors: Principles and applications. Results Opt. 2022, 9, 100285. [Google Scholar] [CrossRef]
  154. Butt, M.A.; Kazanskiy, N.L.; Khonina, S.N.; Voronkov, G.S.; Grakhova, E.P.; Kutluyarov, R.V. A Review on Photonic Sensing Technologies: Status and Outlook. Biosensors 2023, 13, 568. [Google Scholar] [CrossRef]
  155. Shahbaz, M.; Butt, M.A.; Piramidowicz, R. A Concise Review of the Progress in Photonic Sensing Devices. Photonics 2023, 10, 698. [Google Scholar] [CrossRef]
  156. Chen, X.; Milosevic, M.M.; Stanković, S.; Reynolds, S.; Bucio, T.D.; Li, K.; Thomson, D.J.; Gardes, F.; Reed, G.T. The emergence of silicon photonics as a flexible technology platform. Proc. IEEE 2018, 106, 2101–2116. [Google Scholar] [CrossRef]
  157. Mujeeb-U-Rahman, M.; Scherer, A. Silicon-on-insulator-based complementary metal oxide semiconductor integrated optoelectronic platform for biomedical applications. J. Biomed. Opt. 2016, 21, 127004. [Google Scholar] [CrossRef]
  158. Brunetti, G.; Olio, F.D.; Conteduca, D.; Armenise, M.N.; Ciminelli, C. Comprehensive mathematical modelling of ultra-high Q grating-assisted ring resonators. J. Opt. 2020, 22, 035802. [Google Scholar] [CrossRef]
  159. Liu, K.; Jin, N.; Cheng, H.; Chauhan, N.; Puckett, M.W.; Nelson, K.D.; Behunin, R.O.; Rakich, P.T.; Blumenthal, D.J. Ultralow 0.034 dB/m loss wafer-scale integrated photonics realizing 720 million Q and 380 μW threshold Brillouin lasing. Opt. Lett. 2022, 47, 1855–1858. [Google Scholar] [CrossRef]
  160. Cheng, Z.; Chen, X.; Wong, C.Y.; Xu, K.; Tsang, H.K. Mid-infrared Suspended Membrane Waveguide and Ring Resonator on Silicon-on-Insulator. IEEE Photonics J. 2012, 4, 1510–1519. [Google Scholar] [CrossRef]
  161. Marris-Morini, D.; Vakarin, V.; Ramirez, J.M.; Liu, Q.; Ballabio, A.; Frigerio, J.; Montesinos, M.; Alonso-Ramos, C.; Roux, X.L.; Serna, S.; et al. Germanium-based integrated photonics from near- to mid-infrared applications. Nanophotonics 2018, 7, 1781–1793. [Google Scholar] [CrossRef]
  162. Kang, J.; Cheng, Z.; Zhou, W.; Xiao, T.-H.; Gopalakrisna, K.-L.; Takenaka, M.; Tsang, H.K.; Goda, K. Focusing subwavelength grating coupler for mid-infrared suspended membrane germanium waveguides. Opt. Lett. 2017, 42, 2094–2097. [Google Scholar] [CrossRef] [PubMed]
  163. Guo, R.; Zhang, S.; Gao, H.; Murugan, G.S.; Liu, T.; Cheng, Z. Blazed subwavelength grating coupler. Photonics Res. 2023, 11, 189–195. [Google Scholar] [CrossRef]
  164. Kazanskiy, N.L.; Khonina, S.N.; Butt, M.A. Subwavelength Grating Double Slot Waveguide Racetrack Ring Resonator for Refractive Index Sensing Application. Sensors 2020, 20, 3416. [Google Scholar] [CrossRef]
  165. Penadés, J.S.; Alonso-Ramos, C.; Khokhar, A.Z.; Nedeljkovic, M.; Boodhoo, L.A.; Ortega-Moñux, A.; Molina-Fernández, I.; Cheben, P.; Mashanovich, G.Z. Suspended SOI waveguide with sub-wavelength grating cladding for mid-infrared. Opt. Lett. 2014, 39, 5661–5664. [Google Scholar] [CrossRef]
  166. Liu, W.; Ma, Y.; Chang, Y.; Dong, B.; Wei, J.; Ren, Z.; Lee, C. Suspended silicon waveguide platform with subwavelength grating metamaterial cladding for long-wave infrared sensing applications. Nanophotonics 2021, 10, 1861–1870. [Google Scholar] [CrossRef]
  167. Kazanskiy, N.L.; Butt, M.A.; Khonina, S.N. Silicon photonic devices realized on refractive index engineered subwavelength grating waveguides—A review—ScienceDirect. Opt. Laser Technol. 2021, 138, 106863. Available online: https://www.sciencedirect.com/science/article/pii/S0030399220314961?via%3Dihub (accessed on 16 August 2023). [CrossRef]
  168. Heideman, R.; Kooyman, R.; Greve, J. Performance of a highly sensitive optical waveguide Mach-Zehnder interferometer immunosensor. Sens. Actuators B Chem. 1993, 10, 209–217. [Google Scholar] [CrossRef]
  169. Densmore, A.; Vachon, M.; Xu, D.-X.; Janz, S.; Ma, R.; Li, Y.-H.; Lopinski, G.; Delâge, A.; Lapointe, J.; Luebbert, C. Silicon photonic wire biosensor array for multiplexed real-time and label-free molecular detection. Opt. Lett. 2009, 34, 3598–3600. [Google Scholar] [CrossRef]
  170. Ksendzov, A.; Lin, Y. Integrated optics ring-resonator sensors for protein detection. Opt. Lett. 2005, 30, 3344–3346. [Google Scholar] [CrossRef]
  171. Ciminelli, C.; Dell’Olio, F.; Conteduca, D.; Campanella, C.M.; Armenise, M.N. High performance SOI microring resonator for biochemical sensing. Opt. Laser Technol. 2014, 59, 60–67. [Google Scholar] [CrossRef]
  172. Ouyang, H.; Christophersen, M.; Viard, R.; Miller, B.L.; Fauchet, P.M. Macroporous silicon microcavities for macromolecule detection. Adv. Funct. Mater. 2005, 15, 1851–1859. [Google Scholar] [CrossRef]
  173. Chow, E.; Grot, A.; Mirkarimi, L.W.; Sigalas, M.; Girolami, G. Ultracompact biochemical sensor built with two-dimensional photonic crystal microcavity. Opt. Lett. 2004, 29, 1093–1095. [Google Scholar] [CrossRef]
  174. Fard, S.T.; Grist, S.M.; Donzella, V.; Schmidt, S.A.; Flueckiger, J.; Wang, X.; Shi, W.; Millspaugh, A.; Webb, M.; Ratner, D.M. Label-free silicon photonic biosensors for use in clinical diagnostics. In Proceedings of the Silicon Photonics VIII, San Francisco, CA, USA, 2–7 February 2013; SPIE: Bellingham, WA, USA, 2013; Volume 8629, pp. 49–62. [Google Scholar]
  175. Heideman, R.G.; Kooyman, R.P.H.; Greve, J. Development of an optical waveguide interferometric immunosensor. Sens. Actuators B Chem. 1991, 4, 297–299. [Google Scholar] [CrossRef]
  176. Brandenburg, A.; Krauter, R.; Künzel, C.; Stefan, M.; Schulte, H. Interferometric sensor for detection of surface-bound bioreactions. Appl. Opt. 2000, 39, 6396–6405. [Google Scholar] [CrossRef]
  177. Tu, X.; Song, J.; Liow, T.-Y.; Park, M.K.; Yiying, J.Q.; Kee, J.S.; Yu, M.; Lo, G.-Q. Thermal independent Silicon-Nitride slot waveguide biosensor with high sensitivity. Opt. Express 2012, 20, 2640–2648. [Google Scholar] [CrossRef]
  178. Selmi, M.; Gazzah, M.H.; Belmabrouk, H. Optimization of microfluidic biosensor efficiency by means of fluid flow engineering. Sci. Rep. 2017, 7, 5721. [Google Scholar] [CrossRef]
  179. Kumaar, P.; Sivabramanian, A. Design and Bulk Sensitivity Analysis of a Silicon Nitride Photonic Biosensor for Cancer Cell Detection. Int. J. Opt. 2022, 2022, 6085833. [Google Scholar] [CrossRef]
  180. Kundal, S.; Khandelwal, A. Optimization of Sensitivity and Q-Factor of Ring Resonator Based Label Free Biosensor. In Proceedings of the 2022 Workshop on Recent Advances in Photonics (WRAP), Mumbai, India, 4–6 March 2022; pp. 1–2. [Google Scholar]
  181. Bahadoran, M.; Seyfari, A.K.; Sanati, P.; Chua, L.S. Label free identification of the different status of anemia disease using optimized double-slot cascaded microring resonator. Sci. Rep. 2022, 12, 5548. [Google Scholar] [CrossRef]
  182. Lin, V.S.Y.; Motesharei, K.; Dancil, K.P.S.; Sailor, M.J.; Ghadiri, M.R. A porous silicon-based optical interferometric biosensor. Science 1997, 278, 840–843. [Google Scholar] [CrossRef] [PubMed]
  183. Harraz, F.A. Porous silicon chemical sensors and biosensors: A review. Sens. Actuators B Chem. 2014, 202, 897–912. [Google Scholar] [CrossRef]
  184. Lee, M.; Fauchet, P.M. Two-dimensional silicon photonic crystal based biosensing platform for protein detection. Opt. Express 2007, 15, 4530–4535. [Google Scholar] [CrossRef] [PubMed]
  185. Benelarbi, D.; Bouchemat, T.; Bouchemat, M. Study of photonic crystal microcavities coupled with waveguide for biosensing applications. Opt. Quantum Electron. 2017, 49, 347. [Google Scholar] [CrossRef]
  186. Konopsky, V.; Mitko, T.; Aldarov, K.; Alieva, E.; Basmanov, D.; Moskalets, A.; Matveeva, A.; Morozova, O.; Klinov, D. Photonic crystal surface mode imaging for multiplexed and high-throughput label-free biosensing. Biosens. Bioelectron. 2020, 168, 112575. [Google Scholar] [CrossRef]
  187. Chupradit, S.; Ashfaq, S.; Bokov, D.; Suksatan, W.; Jalil, A.T.; Alanazi, A.M.; Sillanpaa, M. Ultra-Sensitive Biosensor with Simultaneous Detection (of Cancer and Diabetes) and Analysis of Deformation Effects on Dielectric Rods in Optical Microstructure. Coatings 2021, 11, 1564. [Google Scholar] [CrossRef]
  188. Baraty, F.; Hamedi, S. Label-Free cancer cell biosensor based on photonic crystal ring resonator. Results Phys. 2023, 46, 106317. [Google Scholar] [CrossRef]
  189. Mohammed, N.A.; Khedr, O.E.; El-Rabaie, E.-S.M.; Khalaf, A.A.M. Brain tumors biomedical sensor with high-quality factor and ultra-compact size based on nanocavity 2D photonic crystal. Alex. Eng. J. 2023, 64, 527–540. [Google Scholar] [CrossRef]
  190. Voronkov, G.S.; Aleksakina, Y.V.; Ivanov, V.V.; Zakoyan, A.G.; Stepanov, I.V.; Grakhova, E.P.; Butt, M.A.; Kutluyarov, R.V. Enhancing the Performance of the Photonic Integrated Sensing System by Applying Frequency Interrogation. Nanomaterials 2023, 13, 193. [Google Scholar] [CrossRef]
  191. Butt, M.A.; Piramidowicz, R. Standard slot waveguide and double hybrid plasmonic waveguide configurations for enhanced evanescent field absorption methane gas sensing. Photonics Lett. Pol. 2022, 14, 10–12. [Google Scholar] [CrossRef]
  192. Wu, J.; Yue, G.; Chen, W.; Xing, Z.; Wang, J.; Wong, W.R.; Cheng, Z.; Set, S.Y.; Senthil Murugan, G.; Wang, X.; et al. On-Chip Optical Gas Sensors Based on Group-IV Materials. ACS Photonics 2020, 7, 2923–2940. [Google Scholar] [CrossRef]
  193. Hodgkinson, J.; Tatam, R.P. Optical gas sensing: A review. Meas. Sci. Technol. 2012, 24, 012004. [Google Scholar] [CrossRef]
  194. Yebo, N.A.; Taillaert, D.; Roels, J.; Lahem, D.; Debliquy, M.; Van Thourhout, D.; Baets, R. Silicon-on-insulator (SOI) ring resonator-based integrated optical hydrogen sensor. IEEE Photonics Technol. Lett. 2009, 21, 960–962. [Google Scholar] [CrossRef]
  195. Dullo, F.T.; Lindecrantz, S.; Jágerská, J.; Hansen, J.H.; Engqvist, M.; Solbø, S.A.; Hellesø, O.G. Sensitive on-chip methane detection with a cryptophane-A cladded Mach-Zehnder interferometer. Opt. Express 2015, 23, 31564–31573. [Google Scholar] [CrossRef] [PubMed]
  196. Kishikawa, H.; Sato, M.; Goto, N.; Yanagiya, S.; Kaito, T.; Liaw, S.K. An optical ammonia gas sensor with adjustable sensitivity using a silicon microring resonator covered with monolayer graphene. Jpn. J. Appl. Phys. 2019, 58, SJJD05. [Google Scholar] [CrossRef]
  197. Gervais, A.; Jean, P.; Shi, W.; LaRochelle, S. Design of slow-light subwavelength grating waveguides for enhanced on-chip methane sensing by absorption spectroscopy. IEEE J. Sel. Top. Quantum Electron. 2018, 25, 5200308. [Google Scholar] [CrossRef]
  198. Kazanskiy, N.L.; Butt, M.A.; Khonina, S.N. Carbon Dioxide Gas Sensor Based on Polyhexamethylene Biguanide Polymer Deposited on Silicon Nano-Cylinders Metasurface. Sensors 2021, 21, 378. [Google Scholar] [CrossRef]
  199. Hänsel, A.; Heck, M.J.R. Opportunities for photonic integrated circuits in optical gas sensors. J. Phys. Photonics 2020, 2, 012002. [Google Scholar] [CrossRef]
  200. Prieto, F.; Sepúlveda, B.; Calle, A.; Llobera, A.; Domínguez, C.; Abad, A.; Montoya, A.; Lechuga, L.M. An integrated optical interferometric nanodevice based on silicon technology for biosensor applications. Nanotechnology 2003, 14, 907. [Google Scholar] [CrossRef]
  201. Yang, R.; Yu, Y.S.; Xue, Y.; Chen, C.; Chen, Q.-D.; Sun, H.-B. Single S-tapered fiber Mach–Zehnder interferometers. Opt. Lett. 2011, 36, 4482–4484. [Google Scholar] [CrossRef]
  202. El Shamy, R.S.; Swillam, M.A.; Khalil, D.A. Mid Infrared Integrated MZI Gas Sensor Using Suspended Silicon Waveguide. J. Light. Technol. 2019, 37, 4394–4400. [Google Scholar] [CrossRef]
  203. El Shamy, R.S.; Swillam, M.A.; ElRayany, M.M.; Sultan, A.; Li, X. Compact Gas Sensor Using Silicon-on-Insulator Loop-Terminated Mach–Zehnder Interferometer. Photonics 2022, 9, 8. [Google Scholar] [CrossRef]
  204. Taha, A.M.; Yousuf, S.; Dahlem, M.S.; Viegas, J. Highly-Sensitive Unbalanced MZI Gas Sensor Assisted With a Temperature-Reference Ring Resonator. IEEE Photonics J. 2022, 14, 6657709. [Google Scholar] [CrossRef]
  205. Troia, B.; Passaro, V.M.N. Investigation of a novel silicon-on-insulator Rib-Slot photonic sensor based on the vernier effect and operating at 3.8 µm. J. Eur. Opt. Soc. Rapid Publ. 2014, 9, 14005. Available online: http://www.jeos.org/index.php/jeos_rp/article/view/951 (accessed on 18 July 2023). [CrossRef]
  206. Zhang, W.; Zhang, X.; Zhang, X.; Jin, H.; Jin, Q.; Jian, J. A single slot micro-ring structure for simultaneous CO2 and CH4 gas sensing. Eur. Phys. J. Appl. Phys. 2018, 82, 30502. [Google Scholar] [CrossRef]
  207. Tomono, Y.; Hoshi, H.; Shimizu, H. CO2 Detection with Si Slot Waveguide Ring Resonators toward On-chip Specific Gas Sensing. In Proceedings of the Conference on Lasers and Electro-Optics, San Jose, CA, USA, 5–10 May 2019; Optica Publishing Group: San Jose, CA, USA, 2019; p. JTh2A.96. [Google Scholar]
  208. Yebo, N.A.; Bogaerts, W.; Hens, Z.; Baets, R. On-Chip Arrayed Waveguide Grating Interrogated Silicon-on-Insulator Microring Resonator-Based Gas Sensor. IEEE Photonics Technol. Lett. 2011, 23, 1505–1507. [Google Scholar] [CrossRef]
  209. Robinson, J.T.; Chen, L.; Lipson, M. On-chip gas detection in silicon optical microcavities. Opt. Express 2008, 16, 4296–4301. [Google Scholar] [CrossRef]
  210. Liu, C.; Sang, C.; Wu, X.; Cai, J.; Wang, J. Grating double-slot micro-ring resonator for sensing. Opt. Commun. 2021, 499, 127280. [Google Scholar] [CrossRef]
  211. Li, H.; Wu, L.; Jin, Y.; Wu, A. High Sensitivity Refractive Index Sensor Based on Trapezoidal Subwavelength Grating Slot Microring Resonator. IEEE Photonics J. 2023, 15, 6801706. [Google Scholar] [CrossRef]
  212. Saeidi, P.; Jakoby, B.; Pühringer, G.; Tortschanoff, A.; Stocker, G.; Dubois, F.; Spettel, J.; Grille, T.; Jannesari, R. Designing Mid-Infrared Gold-Based Plasmonic Slot Waveguides for CO2-Sensing Applications. Sensors 2021, 21, 2669. [Google Scholar] [CrossRef]
  213. Khonina, S.N.; Kazanskiy, N.L.; Butt, M.A. Evanescent Field Ratio Enhancement of a Modified Ridge Waveguide Structure for Methane Gas Sensing Application. IEEE Sens. J. 2020, 20, 8469–8476. [Google Scholar] [CrossRef]
  214. Butt, M.A.; Khonina, S.N.; Kazanskiy, N.L. Silicon on silicon dioxide slot waveguide evanescent field gas absorption sensor. J. Mod. Opt. 2018, 65, 174–178. [Google Scholar] [CrossRef]
  215. Butt, M.A.; Degtyarev, S.A.; Khonina, S.N.; Kazanskiy, N.L. An evanescent field absorption gas sensor at mid-IR 3.39 μm wavelength. J. Mod. Opt. 2017, 64, 1892–1897. [Google Scholar] [CrossRef]
  216. Kazanskiy, N.L.; Khonina, S.N.; Butt, M.A. Polarization-Insensitive Hybrid Plasmonic Waveguide Design for Evanescent Field Absorption Gas Sensor. Photonic Sens. 2021, 11, 279–290. [Google Scholar] [CrossRef]
  217. Liu, J.; Sun, Y.; Fan, X. Highly versatile fiber-based optical Fabry-Pérot gas sensor. Opt. Express 2009, 17, 2731–2738. [Google Scholar] [CrossRef]
  218. Hossein-Babaei, F.; Paknahad, M.; Ghafarinia, V. A miniature gas analyzer made by integrating a chemoresistor with a microchannel. Lab Chip 2012, 12, 1874–1880. [Google Scholar] [CrossRef]
  219. Metehri, F.; Youcef Mahmoud, M.; Bassou, G.; Richalot, E.; Bourouina, T. Stable optofluidic Fabry-Pérot resonator for liquid and gas sensing. Sens. Actuators Phys. 2018, 281, 95–99. [Google Scholar] [CrossRef]
  220. Tao, J.; Zhang, Q.; Xiao, Y.; Li, X.; Yao, P.; Pang, W.; Zhang, H.; Duan, X.; Zhang, D.; Liu, J. A Microfluidic-Based Fabry-Pérot Gas Sensor. Micromachines 2016, 7, 36. [Google Scholar] [CrossRef]
  221. Zhang, Y.; Zhao, Y.; Wang, Q. Optimizing the slow light properties of slotted photonic crystal waveguide and its application in a high-sensitivity gas sensing system. Meas. Sci. Technol. 2013, 24, 105109. [Google Scholar] [CrossRef]
  222. Zhao, Y.; Zhang, Y.N.; Wang, Q. Research advances of photonic crystal gas and liquid sensors. Sens. Actuators B Chem. 2011, 160, 1288–1297. [Google Scholar] [CrossRef]
  223. Feng, C.; Feng, G.-Y.; Zhou, G.R.; Chen, N.J.; Zhou, S.-H. Design of an ultracompact optical gas sensor based on a photonic crystal nanobeam cavity. Laser Phys. Lett. 2012, 9, 875. [Google Scholar] [CrossRef]
  224. Zhang, Y.; Zhao, Y.; Wang, Q. Measurement of methane concentration with cryptophane E infiltrated photonic crystal microcavity. Sens. Actuators B Chem. 2015, 209, 431–437. [Google Scholar] [CrossRef]
  225. Kassa-Baghdouche, L.; Cassan, E. Mid-infrared gas sensor based on high-Q/V point-defect photonic crystal nanocavities. Opt. Quantum Electron. 2020, 52, 260. [Google Scholar] [CrossRef]
  226. Wang, X.; Lv, J.; E, S.; Han, B.; Zhang, Y. Theoretical Design and Simulation Optimization of Photonic Crystal Cavity for Tetrahydrofuran Vapor Sensing. Phys. Status Solidi B 2019, 256, 1900221. [Google Scholar] [CrossRef]
  227. Jágerská, J.; Zhang, H.; Diao, Z.; Thomas, N.L.; Houdré, R. Refractive index sensing with an air-slot photonic crystal nanocavity. Opt. Lett. 2010, 35, 2523–2525. [Google Scholar] [CrossRef]
  228. Li, K.; Li, J.; Song, Y.; Fang, G.; Li, C.; Feng, Z.; Su, R.; Zeng, B.; Wang, X.; Jin, C. Ln Slot Photonic Crystal Microcavity for Refractive Index Gas Sensing. IEEE Photonics J. 2014, 6, 6802509. [Google Scholar] [CrossRef]
Figure 1. Different applications of Si photonics discussed in this review.
Figure 1. Different applications of Si photonics discussed in this review.
Micromachines 14 01637 g001
Figure 2. Si-based super chip. Inspired by [42].
Figure 2. Si-based super chip. Inspired by [42].
Micromachines 14 01637 g002
Figure 3. (a) Cross-sectional SEM image of the SOI p+-i-n+ diode nanophotonic rib WG [68]. (b) Schematic representation of proposed MZI device for all-optical modulation [74].
Figure 3. (a) Cross-sectional SEM image of the SOI p+-i-n+ diode nanophotonic rib WG [68]. (b) Schematic representation of proposed MZI device for all-optical modulation [74].
Micromachines 14 01637 g003
Figure 4. (a) A bird’s-eye view of the microscopic image of the manufactured cascaded micro-ring modulator and the schematic 3D view of the interleaved PN junctions [88], (b) RR modulator top view micrograph, (c) WG cross section, and (d) coupling scheme for the Si photonic chip [89].
Figure 4. (a) A bird’s-eye view of the microscopic image of the manufactured cascaded micro-ring modulator and the schematic 3D view of the interleaved PN junctions [88], (b) RR modulator top view micrograph, (c) WG cross section, and (d) coupling scheme for the Si photonic chip [89].
Micromachines 14 01637 g004
Figure 6. (a) Schematic views of pin germanium photodetector integrated into SOI WG, (b) cross section of the pin diode [133], (c) cross-sectional schematic of the fabricated vertical N(Ge)-I-P(Si) photodetector, and (d) SEM image of the fabricated device [134].
Figure 6. (a) Schematic views of pin germanium photodetector integrated into SOI WG, (b) cross section of the pin diode [133], (c) cross-sectional schematic of the fabricated vertical N(Ge)-I-P(Si) photodetector, and (d) SEM image of the fabricated device [134].
Micromachines 14 01637 g006
Figure 7. (a) Three-dimensional structure schematic of the photodetector, (b) top-view SEM image of the device [137], (c) schematic diagram, and (d) SEM image of the fabricated device [138].
Figure 7. (a) Three-dimensional structure schematic of the photodetector, (b) top-view SEM image of the device [137], (c) schematic diagram, and (d) SEM image of the fabricated device [138].
Micromachines 14 01637 g007
Figure 8. (a) Cross-sectional schematic of a Si WG detector, (b) cross-sectional schematic WG segment of the Si WG detector, (c) top view micrograph of Si photodetector [150], (d) schematic diagram of all-Si MRR APD, (e) eye diagrams of 80 Gb/s NRZ, and (f) 100 Gb/s PAM4 modulations [151].
Figure 8. (a) Cross-sectional schematic of a Si WG detector, (b) cross-sectional schematic WG segment of the Si WG detector, (c) top view micrograph of Si photodetector [150], (d) schematic diagram of all-Si MRR APD, (e) eye diagrams of 80 Gb/s NRZ, and (f) 100 Gb/s PAM4 modulations [151].
Micromachines 14 01637 g008
Figure 9. Plan-sight and cross-sectional schematic representations of the athermal MZI biosensor. Cross-sectional schematic representations and TEM figures of separate arms are also presented [177].
Figure 9. Plan-sight and cross-sectional schematic representations of the athermal MZI biosensor. Cross-sectional schematic representations and TEM figures of separate arms are also presented [177].
Micromachines 14 01637 g009
Figure 10. The setup of a Si3N4 biosensor with MZI and a grating coupler [179].
Figure 10. The setup of a Si3N4 biosensor with MZI and a grating coupler [179].
Micromachines 14 01637 g010
Figure 11. (a) The WG layout of an ICCRR, (b) 3D structure of ICCRR sensor, and (c) the WG cross section of the ICCRR sensor [181].
Figure 11. (a) The WG layout of an ICCRR, (b) 3D structure of ICCRR sensor, and (c) the WG cross section of the ICCRR sensor [181].
Micromachines 14 01637 g011
Figure 12. (a) An isometric view of the proposed structure of an optical microstructure biosensor [187]. (b) Schematic structure of the proposed 2D PhC biosensor [188]. (c) Suggested configuration for the improved design, incorporating optimized parameters [189].
Figure 12. (a) An isometric view of the proposed structure of an optical microstructure biosensor [187]. (b) Schematic structure of the proposed 2D PhC biosensor [188]. (c) Suggested configuration for the improved design, incorporating optimized parameters [189].
Micromachines 14 01637 g012
Figure 13. (a) Three-dimensional schematic representation of the LT-MZI gas sensor. Insets: cross-sectional representations of the strip (reference arm) and slot (sensing arm) WGs [203], (b) 3D schematic of an unbalanced MZI gas sensor supported with a temperature reference RR [204].
Figure 13. (a) Three-dimensional schematic representation of the LT-MZI gas sensor. Insets: cross-sectional representations of the strip (reference arm) and slot (sensing arm) WGs [203], (b) 3D schematic of an unbalanced MZI gas sensor supported with a temperature reference RR [204].
Micromachines 14 01637 g013
Figure 14. (a) SEM picture of a Si-slotted µ-RR. Inset shows the slot WG in the ring [209], (b) 3D representation of the proposed T-SWGS-µ-RR sensor, and (c) structural drawing of the two cascaded T-SWGS-µ-RR sensor [211].
Figure 14. (a) SEM picture of a Si-slotted µ-RR. Inset shows the slot WG in the ring [209], (b) 3D representation of the proposed T-SWGS-µ-RR sensor, and (c) structural drawing of the two cascaded T-SWGS-µ-RR sensor [211].
Micromachines 14 01637 g014
Figure 15. (a) Schematic representation of dual hybrid plasmonic WG. (b) Dual hybrid plasmonic WG employed for CH4 gas sensing. Inspired by [213].
Figure 15. (a) Schematic representation of dual hybrid plasmonic WG. (b) Dual hybrid plasmonic WG employed for CH4 gas sensing. Inspired by [213].
Micromachines 14 01637 g015
Figure 16. Graphical interpretation of the polarization-independent HPWG scheme. Inset exhibits the 3D E-field mapping of TE (bottom-right) and TM (top-left) polarized lights in the WG geometry [216].
Figure 16. Graphical interpretation of the polarization-independent HPWG scheme. Inset exhibits the 3D E-field mapping of TE (bottom-right) and TM (top-left) polarized lights in the WG geometry [216].
Micromachines 14 01637 g016
Table 1. Performance analysis of Si-based modulators employing MZI and ring structures.
Table 1. Performance analysis of Si-based modulators employing MZI and ring structures.
Modulation PrincipleStructureEbit (fJ/bit)ER (dB)References
Electro-refractiveMZI13.2120.36[73]
Electro-opticMZI30.18-[71]
Electro-opticMZI-30[72]
Electro-opticMZI30-[70]
Carrier-depletionRing-3.9[88]
Carrier-depletionRing506.5[87]
Carrier-injectionRing1207[79]
Carrier-depletionRing6808[81]
Carrier-depletionRing->10[83]
Table 2. Performance Comparison and Summary of Different Optical Photodetectors.
Table 2. Performance Comparison and Summary of Different Optical Photodetectors.
MaterialWL
(nm)
Dark Current
(nA)
Dark Current
Density
(mA/cm2)
Responsivity
(A/W)
BW
(GHz)
Speed
(Gbps)
Refs
Ge155035-0.817564[119]
Ge15504218.50.473640[120]
Ge1550150.822950[124]
GeSn1887-730.017--[131]
GeSn20000.0014-0.016--[132]
GeSn2000-1250.01430-[133]
InGaAs/InAlAs/InP1550100.80.799-[141]
InP/InGaAs15500.55-0.3>4040[137]
All Si1550--0.0728~715[148]
All Si1550--0.0033-30[149]
All Si1310--0.5325.5100[151]
Table 3. Comparative analysis of different optical biosensors.
Table 3. Comparative analysis of different optical biosensors.
Sensor ConfigurationSensitivity (nm/RIU)Q-FactorLOD References
Si3N4-based MZI568--[179]
Silica-based RR2002000-[180]
Si/SiO2 double-slot-WG, MRR1024-4.88 × 10−6 RIU[181]
SOI-based PhC WG, microcavity97-0.0055 fg *[185]
PhCs, nanocavities on Si20,393--[187]
Si PhCs, RR308.53803.55-[188]
Si-based PhCs133216,2549.08 × 10−6 RIU[189]
Table 4. Comparative overview of cutting-edge optical sensors designed for detecting different types of gases.
Table 4. Comparative overview of cutting-edge optical sensors designed for detecting different types of gases.
SensorConfigurationGasSensitivity (nm/RIU)Q-FactorLOD (RIU)Refs.
MZIRR-MZIHe, N21458
5500
(Suspended MZI)
-8.5 × 10−5[204]
RRRIB-slotted RRCO2, CH420,600-3.675 × 10−4[205]
RRSlotted MRRCH4, CO22308--[206]
RRSlotted RRCO2300--[207]
RRSlotted MRRAcetylene490500010−5[209]
SPPPSWGCO2--274.6
(Free-standing structure)
70.1
(Asymmetric structure)
[212]
PhCCryptophane-E-infiltrated PhC microcavityCH4363.812,923-[224]
PhCPhC cavityTetrahydrofuran (THF) vapor1942 × 1054 × 10−5[226]
PhCPhC air-slot cavityCO2, N2, He5102.6 × 1041 × 10−5[227]
PhCSlot PhC microcavitiesN2, CO2, He421>3.0 × 1041 × 10−5[228]
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

Shahbaz, M.; Butt, M.A.; Piramidowicz, R. Breakthrough in Silicon Photonics Technology in Telecommunications, Biosensing, and Gas Sensing. Micromachines 2023, 14, 1637. https://doi.org/10.3390/mi14081637

AMA Style

Shahbaz M, Butt MA, Piramidowicz R. Breakthrough in Silicon Photonics Technology in Telecommunications, Biosensing, and Gas Sensing. Micromachines. 2023; 14(8):1637. https://doi.org/10.3390/mi14081637

Chicago/Turabian Style

Shahbaz, Muhammad, Muhammad A. Butt, and Ryszard Piramidowicz. 2023. "Breakthrough in Silicon Photonics Technology in Telecommunications, Biosensing, and Gas Sensing" Micromachines 14, no. 8: 1637. https://doi.org/10.3390/mi14081637

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