Startseite Naturwissenschaften Wafer-scale integration of photonic integrated circuits and atomic vapor cells
Artikel Open Access

Wafer-scale integration of photonic integrated circuits and atomic vapor cells

  • Arieh Grosman ORCID logo , Roy Zektzer , Noa Mazurski , Liron Stern und Uriel Levy EMAIL logo
Veröffentlicht/Copyright: 5. Dezember 2025

Abstract

Atom-based technologies have played a central role in both fundamental research and application-driven developments. For example, devices such as atomic clocks and magnetometers are essential for precision time-keeping, navigation, and sensing. However, many of these demonstrations remain confined to laboratory settings due to their reliance on bulky equipment and centimeter-scale atomic vapor cells. In recent years, significant efforts have been made to miniaturize these vapor cells to enable field-deployable systems. Yet, integrating these cells with the necessary photonic components remains a complex and non-scalable process. To address this challenge, we have introduced the atomic-cladded waveguide (ACWG) architecture, which enables the integration of atomic and photonic functions on the same chip. While the ACWG concept provides a significant step forward toward integration, there is still a significant gap related to wafer scale manufacturability. In particular, previous demonstrations of atomic–photonic integration have relied on manual assembly of vapor cells onto single chips, restricting miniaturization, manufacturability, and thermal robustness. To revolutionize manufacturability of these devices, we hereby demonstrate our new generation of ACWG devices that overcomes these constraints. The approach is based on wafer bonding of a silicon wafer – consisting of multiple photonic chips to a glass wafer with pre-etched atomic chambers. This wafer-scale process yields multiple miniaturized integrated photonic–atomic chips in a single batch. The bonded devices operate reliably at elevated temperatures over an extended period of time, allowing higher atomic densities to be used. The fabrication method consists of well-defined, repeatable steps, paving the way for scalable production of mature integrated photonic–atomic systems for next-generation sensing, metrology, and quantum technologies, inspired by commercial complementary metal-oxide-semiconductor-based processes.

1 Introduction

Alkali vapors interaction with light have been studied over the last several decades due to their multiple applications in metrology and quantum physics, with applications such as atomic clocks [1], magnetometers [2], and RF sensors [3] to name a few. The interaction of light with rubidium (Rb) alkali vapor can serve as a basic building block for atomic clocks [1] that are used as precise frequency and time references in numerous applications such as telecommunications, network synchronization, satellite navigation, and more. Most atomic-based devices such as sensors and clocks require a light source, light modulation, being either frequency, phase, or amplitude modulation, interaction section between light and vapor, and light detection (Figure 1(a)). In order to make quantum sensors an accessible technology, these sensors should be fabricated in large numbers and at low cost. Ideally, all the required optical elements, lasers, modulators, atomic interaction, and photodetection, should be fabricated in the same process and integrated over the same platform. Over the last decade, significant advancements have been made in light manipulation at the chip scale. In fact, the emerging field of silicon photonics is evolving into an enabling technology in creating and successfully exhibiting a range of various nanoscale devices such as lasers [4], modulators [5], and photodetectors [6]. These days, several entities are offering multi-project wafer service (MPW) with several optical materials such as silicon nitride (SIN) for low-loss light propagation [7], lithium niobite for phase modulation [8], and III–V material bonding to silicon for integrated photodetectors and lasers [9]. Over the past decade, significant progress has been made in the miniaturization of vapor cells, evolving from traditional glass-blown (Figure 1(b)), centimeter-scale devices to the microfabrication of microelectromechanical systems-based structures [10], [11], [12] and atomic cladded waveguides (ACWGs) [13]. This progress has enabled reduced power consumption, increased portability, and greater versatility in end-user applications (Figure 1(c)).

Figure 1: 
The evolution of vapor cells. (a) Atomic interaction on a photonic chip requires the integration of various building blocks – such as lasers, modulators, waveguides, and detectors – often fabricated from different materials. (b) Schematic of a traditional centimeter-scale atomic measurement system, featuring a modulated light source directed through a Rb vapor cell and onto a detector. (c) Road map toward miniaturization of Rb cell photonic integration starting from millimeter-sized bonded cells, progresses to glued photonic atomic cladding waveguides, to finally achieving a wafer-scale photonic–atomic platform.
Figure 1:

The evolution of vapor cells. (a) Atomic interaction on a photonic chip requires the integration of various building blocks – such as lasers, modulators, waveguides, and detectors – often fabricated from different materials. (b) Schematic of a traditional centimeter-scale atomic measurement system, featuring a modulated light source directed through a Rb vapor cell and onto a detector. (c) Road map toward miniaturization of Rb cell photonic integration starting from millimeter-sized bonded cells, progresses to glued photonic atomic cladding waveguides, to finally achieving a wafer-scale photonic–atomic platform.

The ACWGs [13], [14] have shown interesting features and already have been used for applications such as strong coupling, extreme nonlinearity [15], [16], optical isolation [17], frequency referencing [18], [19], and quantum optics [20], demonstrating its potential to be a major building block for an atomic–photonic chip system. However, wafer-scale fabrication of these devices remains a challenge in the transition from laboratory experiments to a viable product. While researchers have successfully tackled the challenge of manufacturability and miniaturization for microfabrication vapor cells at the millimeter-scale [21], [22], [23], combining wafer-scale miniaturized vapor cells with silicon complementary metal-oxide-semiconductor (CMOS)-compatible photonic integrated circuits remains a challenge. Mitigating this challenge is expect to enable a new type of devices supporting diverse different applications such as chip-based lasers [24], lithium niobate electro-optic modulators [25], stabilized atomic spectroscopy [26], and photodetectors [6] on low-loss waveguides platform [27], directly connected to standard silicon readout integrated circuit (ROIC).

Motivated by the acute need, this work aims to face the grand challenge of providing wafer-scale photonic–atomic integrated circuits using a CMOS inspired process, without the need for cumbersome glass blowing process and without the need for distillation process. Specifically, we have demonstrated the bonding of a glass wafer to a SIN wafer with several photonic devices separated into four dies. Each dye consists of an Rb dispenser pill that is laser activated. Following fabrication, we have measured and characterized light–atom interactions in these wafer-scale fabricated devices at elevated temperatures and observed their operation over several weeks. These devices were shown to be successful in bridging miniaturization, manufacturability, and integration gap by paving the way for cost effective and mass production of atomic–photonic circuitry.

2 Fabrication

In this work, we present a wafer-scale atomic–photonic platform based on anodic bonding between a SIN photonic circuitry wafer and a BF33 glass wafer containing pre-etched vapor chambers. This CMOS-compatible approach enables batch fabrication of multiple atomic–photonic chips in a single process flow, with hermetic sealing, high thermal stability, and precise die-level repeatability. Our prior studies on ACWGs were based on a cumbersome fabrication process. In short, Rb vapor was introduced into the cell devices by attaching a Pyrex cylinder to the chip with thermally cured epoxy adhesive, followed by connecting the cylinder to a turbo-molecular vacuum system (Figure S1). The assembly was subjected to a 24-h bake-out, achieving a base pressure of 10−7 Torr. Natural Rb was then loaded into the cells (Figures S2 and S3), followed by isolation and hermetic sealing [13], [18], [28]. This fabrication approach presents several limitations: First, it is based on a manual processing. Each cell requires individual assembly, limiting scalability. Furthermore, a glass blowing process is needed for the distillation process. Another issue is the need to introduce a bulky and expensive Pyrex cell for each device. And another troubling issue is related to thermal constraints. The epoxy adhesive degrades above 110 °C, compromising vacuum integrity due to out-gassing and loss of sealing properties. Furthermore, the yield of this process is limited.

To overcome these limitations, we hereby describe a wafer-scale fabrication method that overcomes these disadvantages. Our approach is based on bonding on two wafers, the photonic integrated wafer and the vapor cell wafer. This is done by anodic bonding, also known as “field-assisted sealing” [29]. In its most common implementation, anodic bonding enables hermetic sealing between silicon and borosilicate glass (BF33). This technique has been successfully demonstrated for the fabrication of vapor cells with centimeter- and millimeter-scale dimensions [1], [12]. By adopting this approach, we overcome the thermal limitations imposed by epoxy-based sealing methods and enable reliable operation at elevated temperatures. Subsequently, we advance the technology toward CMOS-compatible, wafer-scale mass production – a capability that, to date, has only been demonstrated for millimeter-scale cells [21], [22]. This approach also addresses the challenges of high-cost and manual device fabrication.

Our wafer-scale fabrication process is illustrated in Figure 2. We start with a 4-inch, 500-μm thick silicon wafer covered with a 2-μm thermal oxide and 250-nm low-pressure chemical vapor deposition grown Si3N4 layer. After cleaning in piranha solution (H2SO4:H2O2 3:1), we spin-coated the wafer with ZEP520 and define the desired pattern, which consists of waveguides using an online lithography toolbox [30] and electron beam lithography (Elionix). The exposed regions were etched by reactive ion etching (RIE, Corial) (Figure 2-1). We have fabricated waveguides of different lengths, with the same cross-section of 700 nm width and a 250 nm height, supporting a single TE mode.

Figure 2: 
Fabrication process: (1) Fabricating waveguides by E-beam lithography and RIE of a 250-nm thick SiN layer on top of a 2-μm SiO2 layer on top of a 0.5-mm 4″ Si substrate wafer; (2) adding 2.5 μm PECVD SiO2 cladding; (3) 500 nm of CMP for planarization; (4) etching vias down to the Si substrate; (5) evaporating a 220-nm a-Si layer to short the surface with the Si substrate; (6) creating an interaction region for light with Rb by etching the SiO2 cladding above the waveguides; (7) prepare the second glass wafer made of two welded glasses, a 0.3-mm plane BF33 wafer, and 1.1 mm BF33 with drilled holes and microchannels for the Rb vapor; (8) anodic bonding for the Si and glass wafers with the Rb pill in the reservoir (AML) by applying a voltage of 800 V, a current limitation of 4 mA, and 500 N pressure at a temperature of 370 °C and vacuum of 10−7 mbar.
Figure 2:

Fabrication process: (1) Fabricating waveguides by E-beam lithography and RIE of a 250-nm thick SiN layer on top of a 2-μm SiO2 layer on top of a 0.5-mm 4″ Si substrate wafer; (2) adding 2.5 μm PECVD SiO2 cladding; (3) 500 nm of CMP for planarization; (4) etching vias down to the Si substrate; (5) evaporating a 220-nm a-Si layer to short the surface with the Si substrate; (6) creating an interaction region for light with Rb by etching the SiO2 cladding above the waveguides; (7) prepare the second glass wafer made of two welded glasses, a 0.3-mm plane BF33 wafer, and 1.1 mm BF33 with drilled holes and microchannels for the Rb vapor; (8) anodic bonding for the Si and glass wafers with the Rb pill in the reservoir (AML) by applying a voltage of 800 V, a current limitation of 4 mA, and 500 N pressure at a temperature of 370 °C and vacuum of 10−7 mbar.

We then deposited a 2.5-μm-thick layer of silicon oxide on top using plasma-enhanced chemical vapor deposition (PECVD) (Figure 2-2). The waveguide pattern in the nitride layer was transferred to the upper oxide cladding, forming micro-trenches that can result in leaks inside the vacuum-sealed vapor reservoir. To assess this, we measured the SiO2 cladding surface profile above the waveguides using a mechanical profilometer (Veeco Dektak 150) and observed trench-like channels alongside each waveguide (Figure S4). To eliminate these channels, we deposited an additional 500 nm SiO2 cladding and planarized it by chemical-mechanical polishing (CMP). The thickness of each layer was measured before and after CMP (Mikropack, NanoCalc 2000), confirming the intended material removal and uniformity. Post-CMP profilometry verified that the trench structures vanished, and AFM measurements confirmed a surface roughness below 0.5 nm (Figure 2-3), suitable for reliable anodic bonding. After cleaning the wafer, we used a laser writer to define patterns in an AZ4562 resist for vias openings, followed by ICP-RIE of all three layers (PECVD oxide, nitride, and thermal oxide) down to the silicon substrate (Figure 2-4). Then, we deposited about 220 nm PECVD amorphous silicon (a-Si) (Figure 2-5) to create an electrical contact between the conductive silicon substrate and the top surface (the amorphous silicon) for the anodic bonding [31]. The use of a deposited layer such as amorphous silicon as a binder allow us to bond an oxide cladded wafer to glass and can potentially support many other wafer technologies such as lithium niobite on the insulator. The area of Rb vapor and ACWG interaction was then defined by a laser writer, followed by an oxide wet etching by a buffered HF solution (BOE) (Figure 2-6). At this point, the first wafer for the anodic bonding is ready.

The second glass wafer must satisfy several requirements: First, it must be transparent at the activation illumination wavelength to generate sufficient heat corresponding to the operational requirements of a commercially available alkali dispenser (SAES Rb/AMAX/Pill1-0.6). Second, the wafer must have sufficient thickness to allow for the fabrication of a reservoir capable of accommodating the Rb dispenser (Rb dispenser dimensions are 1 × 0.6 mm). Third, the thermal expansion should be well matched to the silicon thermal expansion. Finally, the contact surface should be polished to a minimal roughness to facilitate anodic bonding sealing.

Indeed, there are several established approaches for fabricating reservoir cavities in glass, each with distinct limitations that we evaluated experimentally. Mechanical drilling, for example, often results in poor surface quality and can introduce cracks due to the deformation of glass under the thrust force of the drill, particularly at the exit surface [32]. Additionally, this method does not meet the transparency requirements for our application. Water jet micromachining was also tested; however, the resulting holes exhibited highly irregular surfaces, non-uniform inner walls, and residual particles adhering to the glass, most likely due to the high operating pressures involved. These surface imperfections complicate subsequent bonding steps [12]. Wet etching, typically using hydrofluoric acid (HF), is another option, but for 1-mm thick glass, the process is very slow – typical etch rates range from 0.07 μm/s to 0.2 μm/s, requiring up to 4 h for complete etching. Furthermore, the isotropic nature of wet etching in amorphous glass leads to rounded sidewalls and low aspect ratios, and the process is prone to defects such as pinholes and notching, primarily due to residual stress in the masking materials [33]. Deep reactive ion etching (DRIE) was also considered, but it exhibits a low etch rate making it impractical for fabricating millimeter-scale features, as etching a 1 mm-thick glass would require many hours.

Given those limitations, we adopted a hybrid approach: bonding two separate Borofloat 33 (BF33) glass wafers [34] of different thicknesses. The first, a 1.1-mm-thick wafer, was used to fabricate through-holes for the dispenser reservoirs, while the second, a 0.3-mm-thick wafer, served as a transparent window.

Glass-to-glass bonding can be achieved using several techniques, including thermocompression [35] or plasma-assisted hydrophilic glass-to-glass direct bonding [36], [37]. However, to minimize the risk of contamination from bonding agents or residual molecules that could interact with the Rb vapor, we elected to bond the two BF33 glass wafers using laser welding technology (Figure 2-7). This approach enables a robust, hermetic seal without introducing additional materials into the device environment.

Subsequently, anodic bonding was carried out using a wafer bonder (AML AWB-04). Prior to bonding, both wafer surfaces were thoroughly cleaned with piranha solution, and the Rb dispenser was carefully positioned within the reservoir of each device. The wafers were then loaded into the bonding chamber, which was evacuated to a base pressure of 10−7 Torr and simultaneously baked at 370 °C for 24 h to remove residual moisture. The anodic bonding process was conducted at 370 °C, with an applied voltage of 800 V, a current limit of 4 mA, and an applied force of 500 N (Figure 2-8). Figure 3-a shows the lithographically patterned silicon wafer, with an a-Si top layer, positioned inside the wafer bonder chamber prior to anodic bonding. Rb dispensers are aligned atop the designated reservoir sites.

Figure 3: 
From wafer-level fabrication to chip-scale ACWG devices. (a) Photograph of the lithographically patterned silicon wafer, with an a-Si top layer, positioned inside the wafer bonder chamber prior to anodic bonding. Rb dispenser pills are aligned atop the designated reservoir sites. (b) Photograph of the bonded wafer after dicing showing four individual devices. Key features are annotated: reservoir (Rb dispenser pill), ACWG interaction region, microchannel, and optical coupling areas. Inset (not in scale): Cross-sectional schematic of layer stack (from bottom-up): Si substrate (500 μm), SiO2 (2 μm), SiN waveguide (250 nm), SiO2 (2 μm), a-Si (220 nm), and two layers of BF33 glass (1.4 mm). (c) Illustration of a diced photonic ACWG device during Rb dispenser pill activation.
Figure 3:

From wafer-level fabrication to chip-scale ACWG devices. (a) Photograph of the lithographically patterned silicon wafer, with an a-Si top layer, positioned inside the wafer bonder chamber prior to anodic bonding. Rb dispenser pills are aligned atop the designated reservoir sites. (b) Photograph of the bonded wafer after dicing showing four individual devices. Key features are annotated: reservoir (Rb dispenser pill), ACWG interaction region, microchannel, and optical coupling areas. Inset (not in scale): Cross-sectional schematic of layer stack (from bottom-up): Si substrate (500 μm), SiO2 (2 μm), SiN waveguide (250 nm), SiO2 (2 μm), a-Si (220 nm), and two layers of BF33 glass (1.4 mm). (c) Illustration of a diced photonic ACWG device during Rb dispenser pill activation.

Alignment marks patterned at identical positions on both the glass and silicon wafers enabled accurate alignment (∼+/−2.5 μm accuracy) prior to the bonding step. Figure 3-b presents the post-dicing bonded wafer containing four individual devices fabricated from the 4-inch substrate. The image delineates the following functional regions within each device: reservoir housing the Rb dispenser pill, ACWG interaction region where atom–light interactions occur, a microchannel connecting the reservoir to the waveguide region (microchannel dimensions – width: 500 μm, length: 2,400 μm, height: 1.1 mm), optical coupling interfaces consisting of input/output grating couplers for light injection and collection. Inset describing (not in scale) cross-sectional schematic of the layered structure (from bottom-up): Si substrate (500 μm), SiO2 (2 μm), SiN waveguide (250 nm), SiO2 (2 μm), a-Si (220 nm), and two welded layers of BF33 (1.4 mm). We performed several wafer-bonding runs and obtained repeatable results with no evident cracks or stitching-related defects.

A fiber-coupled diode laser (830 nm, K830F02FN-2.000W) was used to activate the Rb dispenser. The beam was focused onto the pill surface with a 50-mm focal length lens. Figure 3-c illustrates a photonic, diced, ACWG device during Rb dispenser pill activation. During activation, the diode laser was driven at 1 A, delivering 0.38 mW of optical power for 30 s. The chip was maintained at 120 °C throughout the process to prevent condensation of atoms on the waveguides.

3 Atomic clad waveguides (ACWGs) – experimental characterization

After device fabrication and Rb dispenser activation, we characterize the optical and atomic properties of the integrated chip devices. The experimental setup was designed to simultaneously enable precise alignment of optical lens fibers to the chip facets, temperature control, and monitoring of spectroscopic optical signals. In the following subsections, we detail the measurement configurations, procedures, and data analysis methods used to evaluate device performance. A narrow-linewidth 780 nm laser (Toptica) was fiber-coupled to a 90:10 splitter. The low-power output was directed to a reference Rb cell for reference and calibration, while the higher-power arm was sent through an optical attenuator and then to a 50:50 splitter. One output of this splitter was connected to an optical switch, which routed the light either to a power meter or to a fiber collimator for free-space spectroscopy measurements. During Rb dispenser activation, the output from the chip’s reservoir area was reflected onto a photodetector (PD) equipped with a 780-nm bandpass filter. This configuration enabled in situ saturated absorption free space spectroscopy, allowing us to monitor the Rb vapor density and try to avoid overactivation of the dispenser [12]. Figure S5 shows the initial Rb free space absorption signal recorded immediately after pill activation. Following the detection of free-space Rb spectroscopy via reflected light from the chip, the switch was set to direct the output to a power meter. The second output of the 50:50 splitter was fiber-coupled to the input facet of the device, injecting light into the waveguide via edge coupling to an inverse taper. The guided light then interacted with the Rb atoms in the interaction region through the evanescent field and was subsequently collected at the output facet by another inverse taper and coupled via lensed fiber to a photodetector. Since the start of the activation process, the device under test was placed on top of an oven to maintain the desired temperature, which was applied to both the reservoir and the interaction region. In Figure 4-a, we present a schematic of the simultaneous Rb dispenser activation and measurement setup. Figure 4-b shows a photograph of the device during activation, illustrating the configuration for concurrent free-space and waveguide measurements.

Figure 4: 
Measurement setup. (a) Schematic of the experimental setup for simultaneous Rb dispenser activation and measurement. The setup enables in situ saturated absorption spectroscopy during activation, as well as optical characterization via both free-space and waveguide-coupled configurations. (b) Photograph of the device under activation, showing the arrangement for concurrent free-space and waveguide measurements. The laser-heated pill is emitting bright red.
Figure 4:

Measurement setup. (a) Schematic of the experimental setup for simultaneous Rb dispenser activation and measurement. The setup enables in situ saturated absorption spectroscopy during activation, as well as optical characterization via both free-space and waveguide-coupled configurations. (b) Photograph of the device under activation, showing the arrangement for concurrent free-space and waveguide measurements. The laser-heated pill is emitting bright red.

The fabricated device contains several SIN waveguides. The waveguides have a different length of exposure to the Rb vapor, which is predefined in the SiO2 etching mask step of the fabrication (Figure 2-6). The overall coupling loss, measured at 170 °C, was found to be ∼50 dB, including coupling loss from the fiber to the waveguide (two facets) and from the oxide-clad waveguide to the ACWG (two facets). We attribute the high losses also to the condensation of Rb atoms or other materials ejected from the Rb pill during activation. Other possible reasons for the high loss are the instability of the edge coupling measurement at high temperature due to thermal expansion and hot air flow, as well as low dicing quality of the facets. The loss mechanisms are now being studied, with the goal of finding the exact loss mechanisms and achieving significant loss reduction in future devices. We have fabricated waveguide with interaction lengths varying from 200 μm to several millimeters but were only able to observe reasonable transmission for devices with interaction lengths of 200 and 500 μm. After the activation, we kept the device heated to 100 °C permanently to generate a temperature gradient between the Si chip and the glass cover and prevent any further condensation of Rb on the surface [38]. In Figure 5, we present measurements of evanescent absorption in the ACWG device with a 200 μm interaction length using a constant input power of 0.2 mW at 780 nm. The measurements were performed at various temperatures, ranging from 125 °C to 195 °C in 10 °C increments, under conditions where the optical transitions are not saturated. As expected, the absorption contrast increases with the temperature. Additionally, when compared to the reference signal from the Rb cell (Cosy), the broadening of the Rb resonance becomes more pronounced at higher temperatures. This broadening can be attributed to both the finite transit and the Doppler effect, with Doppler broadening being the dominant mechanism in this case [13], [14], [39].

Figure 5: 
Evanescent absorption spectra measured in a 200-μm interaction region of the ACWG device with Rb vapor, using a constant input power of 50 μW at 780 nm. Measurements were performed from 125 °C to 195 °C in 10 °C increments. Increased absorption contrast observed at higher temperature, consistent with an enhanced atomic density. Resonance broadening increases by transit time and Doppler broadening effects.
Figure 5:

Evanescent absorption spectra measured in a 200-μm interaction region of the ACWG device with Rb vapor, using a constant input power of 50 μW at 780 nm. Measurements were performed from 125 °C to 195 °C in 10 °C increments. Increased absorption contrast observed at higher temperature, consistent with an enhanced atomic density. Resonance broadening increases by transit time and Doppler broadening effects.

We repeated the measurements for a longer ACWG interaction length of 500 μm, as shown in Figure S6. Like the results obtained with the shorter interaction length, we observed consistent behavior both in the saturation region and at maximum absorption. Notably, the longer interaction length resulted in improved absorption contrast at lower input power, which can be attributed to the increased interaction between the evanescent guided light and the Rb atoms along the extended ACWG region.

Next, we investigated the nonlinear interaction in the ACWG by measuring the absorption spectrum as a function of incident light intensity, which was controlled using an attenuator placed before the 50:50 splitter. In Figure 6, we present the transmission spectra obtained for a 200-μm interaction length at a temperature of 185 °C. Under these conditions, we observed a high absorption contrast at a low input optical power (10 nW) and clear saturation behavior at higher input power.

Figure 6: 
Minimal transimission of the 85Rb F = 2 → F′ = 2/3 transition as a function of the optical power within the ACWG for 200-μm interaction region at 185 °C. The estimated saturation power is 81 nW.
Figure 6:

Minimal transimission of the 85Rb F = 2 → F′ = 2/3 transition as a function of the optical power within the ACWG for 200-μm interaction region at 185 °C. The estimated saturation power is 81 nW.

To estimate the saturation power, we analyzed the contrast data for the 85Rb F = 2 → F′ = 2/3 transition at 185 °C as a function of input power. The data were fitted to the exponential function: T P = exp α 0 1 + P P sat 0.5 , where α 0 is the optical density corresponding to the unsaturated level measured, P is the optical power level within the ACWG at resonance (obtained by a linear fit around the resonance), and P sat is the saturation optical power [13]. The fit, shown in (Figure 7), yields an estimated onset of saturation of 81 nW.

Figure 7: 
Measurement of devices of the 200-μm interaction region of Rb at 185 °C.
Figure 7:

Measurement of devices of the 200-μm interaction region of Rb at 185 °C.

We repeat these 85Rb F = 2 transition measurements for different temperatures. In Figure S7, we present a series of measurements for a 200-μm long waveguide showing the effects of varying both temperature and input power. Similarly, Figure S8 displays corresponding measurements for a 500-μm interaction length.

4 Discussion

In summary, we have demonstrated a new platform of wafer atomic photonic integrated circuits that are integrated with a wafer of hot vapor cells filled with Rb pills as the vapor source. We have used this platform to fabricate and demonstrate wafer-scale ACWGs operating at unprecedently high temperatures of up to 195 °C.

Temperature- and power-dependent spectroscopic measurements were used to characterize atom–light interactions, revealing Doppler broadening as the dominant spectral broadening mechanism. Importantly, no significant spectral broadening attributable to contamination or residual byproducts from the anodic bonding process or the Rb dispenser was observed. The device was measured over several weeks of nearly continuous operation without showing noticeable degradation.

Waveguide losses are currently noticeable and are attributed to the condensation of Rb vapor on the waveguide. The issue of condensation is known [38], [40] and seems to be more severe in microfabricated cells. We intend to take several steps to tackle this effect, including the use of light-induced atomic desorption [40], [41], [42] and alumina deposition [43]. A complementary solution will be to minimize the amount of Rb being sputtered at the moment of activation, e.g., by precisely control the optical power required for activation, blocking any direct line of sight between the pill and the photonic devices, and studying different Rb placement technics [21], [44] where devices are isolated at the moment of activation.

To further reduce the overall optical loss, future device iterations will focus on improving facet quality by enhancing facet smoothness through optimized dicing and polishing procedures, or alternatively by implementing a deep RIE facet-etch process to define clean, vertical facets. In addition, replacing the current oxide wet-etch with an ICP-PECVD oxide lift-off approach is expected to improve the quality of the interface between the region with top oxide cladding to the region where the top oxide is removed, thereby reducing scattering losses. It may also be that the silicon–glass anodic bonding at 370 °C produces mechanical stress in the waveguides, leading to loss and perhaps also refractive index modulation and birefringence. This will be studied in more detail in future devices.

While challenges remain, the demonstrated platform holds an enormous potential to become an enabling technology in wafer-scale integrated atomic devices, allowing miniaturization, high level of integration, manufacturability, and low-cost devices. This will pave the way for the use of such devices in applications ranging from frequency referencing, wavelength meters, laser stabilization, and atomic clocks to magnetic and electric field sensing.


Corresponding author: Uriel Levy, Institute of Applied Physics, The Faculty of Science, The Center for Nanoscience and Nanotechnology, The Hebrew University of Jerusalem, Jerusalem, Israel, E-mail: 

Acknowledgments

The devices were fabricated at the Center for Nanoscience and Nanotechnology of the Hebrew University.

  1. Research funding: The research was supported by the Israeli Science Foundation.

  2. Author contributions: AG was the leading student on the project, contributing to device design, fabrication and characterization. NM fabricated the devices. RZ, LS, and UL proposed the concept and contributed to the various stages of the work. UL supervised the project. All authors have accepted responsibility for the entire content of this manuscript and consented to its submission to the journal, reviewed all the results, and approved the final version of the manuscript.

  3. Conflict of interest: Authors state no conflict of interest.

  4. Informed consent: Informed consent was obtained from all individuals included in this study.

  5. Ethical approval: The conducted research is not related to either human or animals use.

  6. Data availability: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.

References

[1] F. D. Martinez, C. Li, A. Staron, J. Kitching, C. Raman, and W. R. Mcflehee, “A chip-scale atomic beam clock,” Nat. Commun., vol. 14, no. 1, pp. 1–7, 2023. https://doi.org/10.1038/s41467-023-39166-1.Suche in Google Scholar PubMed PubMed Central

[2] I. K. Kominis, T. W. Kornack, J. C. Allred, and M. V. Romalis, “A subfemtotesla multichannel atomic magne-tometer,” Nature, vol. 422, no. 6932, pp. 596–599, 2003. https://doi.org/10.1038/nature01484.Suche in Google Scholar PubMed

[3] A. fliat, K. Levi, O. Nefesh, and L. Stern, “Sub-wavelength micromachined vapor-cell based rydberg sensing,” arxiv, 2025.Suche in Google Scholar

[4] D. Liang and J. E. Bowers, “Recent progress in lasers on silicon,” Nat. Photonics, vol. 4, no. 8, pp. 511–517, 2010. https://doi.org/10.1038/nphoton.2010.167.Suche in Google Scholar

[5] B. Desiatov, A. Shams-Ansari, M. Zhang, C. Wang, and M. Lončar, “Ultra-low-loss integrated visible photonics using thin-film lithium niobate,” Optica, vol. 6, no. 3, p. 380, 2019. https://doi.org/10.1364/OPTICA.6.000380.Suche in Google Scholar

[6] I. Floykhman, B. Desiatov, J. Khurgin, J. Shappir, and U. Levy, “Locally oxidized silicon surface-plasmon schottky detector for telecom regime,” Nano Lett., vol. 11, no. 6, pp. 2219–2224, 2011. https://doi.org/10.1021/NL200187V/ASSET/IMAflES/MEDIUM/NL-2011-00187V_0008.flIF.Suche in Google Scholar

[7] A. Diakonov, K. Khrizman, E. Zano, and L. Stern, “Broadband cavity-enhanced Kerr comb spectroscopy on chip,” NPJ Nanophoton., vol. 1, no. 1, p. 47, 2024, https://doi.org/10.1038/s44310-024-00047-0.Suche in Google Scholar

[8] A. Honardoost, “HyperLight’s thin-film lithium niobate (TFLN) platform: Unlocking the future of integrated photonics,” in Quantum Information Science, Sensing, and Computation XVII, M. L. Fanto and M. Hayduk, Eds., SPIE, 2025, p. 37.10.1117/12.3069394Suche in Google Scholar

[9] N. M. Fahrenkopf, C. McDonough, F. L. Leake, Z. Su, E. Timurdogan, and D. D. Cool-Baugh, “The AIM photonics MPW: A highly accessible cutting edge technology for rapid prototyping of pho-tonic integrated circuits,” IEEE J. Sel. Top. Quantum Electron., vol. 25, no. 5, pp. 1–6, 2019, https://doi.org/10.1109/JSTQE.2019.2935698.Suche in Google Scholar

[10] J. Kitching, “Chip-scale atomic devices,” Appl. Phys. Rev., vol. 5, no. 3, p. 031302, 2018. https://doi.org/10.1063/1.5026238.Suche in Google Scholar

[11] S. Liron, D. F. Bopp, S. A. Schima, V. N. Maurice, and J. E. Kitching, “Chip-scale atomic diffractive optical elements,” Nat. Commun., vol. 10, no. 1, pp. 1–7, 2019. https://doi.org/10.1038/s41467-019-11145-5.Suche in Google Scholar PubMed PubMed Central

[12] S. Dyer et al.., “Micro-machined deep silicon atomic vapor cells,” J. Appl. Phys., vol. 132, no. 13, p. 134401, 2022. https://doi.org/10.1063/5.0114762.Suche in Google Scholar

[13] L. Stern, B. Desiatov, I. floykhman, and U. Levy, “Nanoscale light-matter interactions in atomic cladding waveguides,” Nat. Commun., vol. 4, no. 1, p. 1548, 2013. https://doi.org/10.1038/ncomms2554.Suche in Google Scholar PubMed PubMed Central

[14] R. Ritter, N. flruhler, W. Pernice, H. Kübler, T. Pfau, and R. Löw, “Atomic vapor spectroscopy in integrated photonic structures,” Appl. Phys. Lett., vol. 107, no. 4, 2015. https://doi.org/10.1063/1.4927172.Suche in Google Scholar

[15] A. Skljarow et al.., “Integrating two-photon nonlinear spectroscopy of rubidium atoms with silicon photonics,” Opt. Express, vol. 28, no. 13, pp. 19593–19607, 2020. https://doi.org/10.1364/OE.389644.Suche in Google Scholar PubMed

[16] L. Stern, R. Zektzer, N. Mazurski, and U. Levy, “Enhanced light-vapor interactions and all optical switching in a chip scale micro-ring resonator coupled with atomic vapor: Enhanced light-vapor interactions,” Laser Photonics Rev., vol. 10, no. 6, pp. 1016–1022, 2016, https://doi.org/10.1002/lpor.201600176.Suche in Google Scholar

[17] R. Zektzer, E. Talker, Y. Barash, N. Mazurski, and U. Levy, “Chiral light–matter interactions in hot vapor-cladded waveguides,” Optica, vol. 6, no. 1, p. 15, 2019. https://doi.org/10.1364/OPTICA.6.000015.Suche in Google Scholar

[18] R. Zektzer, N. Mazurski, Y. Barash, and U. levy, “Nanoscale atomic suspended waveguides for improved vapour coherence times and optical frequency referencing,” Nat. Photonics, vol. 15, no. 10, pp. 772–779, 2021. https://doi.org/10.1038/s41566-021-00853-4.Suche in Google Scholar

[19] R. Zektzer, E. Talker, Y. Barash, N. Mazurski, L. Stern, and U. Levy, “Atom–photon interactions in atomic cladded waveguides: Bridging atomic and tele-com technologies,” ACS Photonics, vol. 8, no. 3, pp. 879–886, 2021. https://doi.org/10.1021/acsphotonics.0c01895.Suche in Google Scholar

[20] R. Zektzer et al.., “Strong interactions between integrated mi-croresonators and alkali atomic vapors: Towards single-atom, single-photon operation,” Optica, vol. 11, no. 10, p. 1376, 2024. https://doi.org/10.1364/OPTICA.525689.Suche in Google Scholar

[21] Y. Li, M. R. Slot, M. T. Hummon, S.-S. Schima, and J. Kitching, “Wafer-scale fabrication of temperature-compensated alkali vapor cells,” Appl. Phys. Lett., vol. 126, no. 22, p. 224001, 2025. https://doi.org/10.1063/5.0264409.Suche in Google Scholar

[22] T. Overstolz et al.., “Wafer scale fabrication of highly integrated rubidium vapor cells,” in Pro-ceedings of the IEEE International Conference on Micro Electro Mechanical Systems (MEMS), 2014, pp. 552–555.10.1109/MEMSYS.2014.6765700Suche in Google Scholar

[23] Y. Pétremand et al.., “Microfabricated rubidium vapour cell with a thick glass core for small-scale atomic clock applications,” J. Micromech. Microeng., vol. 22, no. 2, 2012, Art. no. 025013. https://doi.org/10.1088/0960-1317/22/2/025013.Suche in Google Scholar

[24] J. Fluo et al.., “Chip-based laser with 1-hertz integrated linewidth,” Sci. Adv., vol. 8, no. 43, p. 9006, 2022. https://doi.org/10.1126/SCIADV.ABP9006/ASSET/B5EC91F9-8F70-45F7-8591-A65F66AE3777/ASSETS/IMAGES/LARGE/SCIADV.ABP9006-F4.JPG.Suche in Google Scholar

[25] C. Wang et al.., “Integrated lithium niobate electro-optic modulators operating at CMOS-compatible voltages,” Nature, vol. 562, pp. 101–104, 2018. https://doi.org/10.1038/s41586-018-0551-y.Suche in Google Scholar PubMed

[26] M. H. Idjadi and F. Aflatouni, “Integrated Pound−Drever−Hall laser stabilization system in silicon,” Nat. Commun., vol. 8, no. 1, p. 1209, 2017. https://doi.org/10.1038/s41467-017-01303-y.Suche in Google Scholar PubMed PubMed Central

[27] J. Cardenas et al.., “Low loss etchless silicon photonic waveguides,” Opt. Express, vol. 17, no. 6, pp. 4752–4757, 2009, https://doi.org/10.1364/OE.17.004752.Suche in Google Scholar PubMed

[28] L. Stern, B. Desiatov, N. Mazurski, and U. Levy, “Strong coupling and high-contrast all-optical modulation in atomic cladding waveguides,” Nat. Commun., vol. 8, no. 1, p. 14461, 2017. https://doi.org/10.1038/ncomms14461.Suche in Google Scholar PubMed PubMed Central

[29] F. Wallis and D. I. Pomerantz, “Field assisted glass-metal sealing,” J. Appl. Phys., vol. 40, no. 10, pp. 3946–3949, 1969. https://doi.org/10.1063/1.1657121.Suche in Google Scholar

[30] K. C. Balram et al.., “The nanolithography toolbox,” J. Res. Natl. Inst. Stand. Technol., vol. 121, no. 1, pp. 464–475, 2016. https://doi.org/10.6028/jres.121.024.Suche in Google Scholar PubMed PubMed Central

[31] E. Talker, R. Zektzer, Y. Barash, N. Mazurski, and U. Levy, “Atomic spectroscopy and laser frequency stabilization with scalable micrometer and sub-micrometer vapor cells,” J. Vac. Sci. Technol. B, vol. 38, no. 5, p. 050601, 2020. https://doi.org/10.1116/6.0000416.Suche in Google Scholar

[32] L. Hof and J. Abou Ziki, “Micro-hole drilling on glass substrates – a review,” Micromachines, vol. 8, no. 2, p. 53, 2017. https://doi.org/10.3390/mi8020053.Suche in Google Scholar

[33] C. Iliescu, B. Chen, and J. Miao, “Deep wet etching-through 1 mm pyrex glass wafer for microfluidic applications,” in 2007 IEEE 20th International Conference on Micro Electro Mechanical Systems (MEMS), IEEE, 2007, pp. 393–396.10.1109/MEMSYS.2007.4433150Suche in Google Scholar

[34] schott, “Borofloat 33 datasheet,” 2025, Available at: https://www.schott.com/en-il/products/borofloat-p1000314/downloads.Suche in Google Scholar

[35] S. Karlen, J. Haesler, T. Overstolz, F. Bergonzi, and S. Lecomte, “Sealing of mems atomic vapor cells using cu-cu thermocompression bonding,” J. Microelectromech. Syst., vol. 29, no. 1, pp. 95–99, 2020. https://doi.org/10.1109/JMEMS.2019.2949349.Suche in Google Scholar

[36] F. Kalkowski, S. Risse, U. Zeitner, F. Fuchs, R. Eberhardt, and A. Tünnermann, “Glass-glass direct bonding,” ECS Trans., vol. 64, no. 5, pp. 3–11, 2014. https://doi.org/10.1149/06405.0003ecst.Suche in Google Scholar

[37] J. P. Mcflilligan, K. Flallacher, P. F. Flriffin, D. J. Paul, A. S. Arnold, and E. Riis, “Micro-fabricated components for cold atom sensors,” Rev. Sci. Instrum., vol. 93, no. 9, 2022. https://doi.org/10.1063/5.0101628.Suche in Google Scholar PubMed

[38] M. Lai, J. D. Franson, and T. B. Pittman, “Transmission degradation and preservation for tapered optical fibers in rubidium vapor,” Appl. Opt., vol. 52, no. 12, p. 2595, 2013. https://doi.org/10.1364/AO.52.002595.Suche in Google Scholar PubMed

[39] L. Stern and U. Levy, “Transmission and time delay properties of an integrated system consisting of atomic vapor cladding on top of a micro ring resonator,” Opt. Express, vol. 20, pp. 28082–28093, 2012.10.1364/OE.20.028082Suche in Google Scholar PubMed

[40] S. M. Hendrickson, T. B. Pittman, and J. D. Franson, “Non-linear transmission through a tapered fiber in rubidium vapor,” J. Opt. Soc. Am. B, vol. 26, no. 2, p. 267, 2009. https://doi.org/10.1364/JOSAB.26.000267.Suche in Google Scholar

[41] V. Venkataraman, K. Saha, P. Londero, and A. L. flaeta, “Few-photon all-optical mod-ulation in a photonic band-gap fiber,” Phys. Rev. Lett., vol. 107, no. 19, p. 193902, 2011. https://doi.org/10.1103/PhysRevLett.107.193902.Suche in Google Scholar PubMed

[42] E. Talker, P. Arora, R. Zektzer, Y. Sebbag, M. Dikopltsev, and U. Levy, “Light-induced atomic desorption in microfabricated vapor cells for demon-strating quantum optical applications,” Phys. Rev. Appl., vol. 15, no. 5, p. L051001, 2021. https://doi.org/10.1103/PhysRevApplied.15.L051001.Suche in Google Scholar

[43] R. Ritter, N. Flruhler, W. H. P. Pernice, H. Kübler, T. Pfau, and R. Löw, “Coupling thermal atomic vapor to an inte-grated ring resonator,” New J. Phys., vol. 18, no. 10, 2016, Art. no. 103031, https://doi.org/10.1088/1367-2630/18/10/103031.Suche in Google Scholar

[44] V. Maurice et al.., “Wafer-level vapor cells filled with laser-actuated hermetic seals for integrated atomic devices,” Microsyst. Nanoeng., vol. 8, no. 1, p. 129, 2022. https://doi.org/10.1038/s41378-022-00468-x.Suche in Google Scholar PubMed PubMed Central


Supplementary Material

This article contains supplementary material (https://doi.org/10.1515/nanoph-2025-0500).


Received: 2025-09-24
Accepted: 2025-11-21
Published Online: 2025-12-05

© 2025 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Artikel in diesem Heft

  1. Frontmatter
  2. Reviews
  3. Light-driven micro/nanobots
  4. Tunable BIC metamaterials with Dirac semimetals
  5. Large-scale silicon photonics switches for AI/ML interconnections based on a 300-mm CMOS pilot line
  6. Perspective
  7. Density-functional tight binding meets Maxwell: unraveling the mysteries of (strong) light–matter coupling efficiently
  8. Letters
  9. Broadband on-chip spectral sensing via directly integrated narrowband plasmonic filters for computational multispectral imaging
  10. Sub-100 nm manipulation of blue light over a large field of view using Si nanolens array
  11. Tunable bound states in the continuum through hybridization of 1D and 2D metasurfaces
  12. Integrated array of coupled exciton–polariton condensates
  13. Disentangling the absorption lineshape of methylene blue for nanocavity strong coupling
  14. Research Articles
  15. Demonstration of multiple-wavelength-band photonic integrated circuits using a silicon and silicon nitride 2.5D integration method
  16. Inverse-designed gyrotropic scatterers for non-reciprocal analog computing
  17. Highly sensitive broadband photodetector based on PtSe2 photothermal effect and fiber harmonic Vernier effect
  18. Online training and pruning of multi-wavelength photonic neural networks
  19. Robust transport of high-speed data in a topological valley Hall insulator
  20. Engineering super- and sub-radiant hybrid plasmons in a tunable graphene frame-heptamer metasurface
  21. Near-unity fueling light into a single plasmonic nanocavity
  22. Polarization-dependent gain characterization in x-cut LNOI erbium-doped waveguide amplifiers
  23. Intramodal stimulated Brillouin scattering in suspended AlN waveguides
  24. Single-shot Stokes polarimetry of plasmon-coupled single-molecule fluorescence
  25. Metastructure-enabled scalable multiple mode-order converters: conceptual design and demonstration in direct-access add/drop multiplexing systems
  26. High-sensitivity U-shaped biosensor for rabbit IgG detection based on PDA/AuNPs/PDA sandwich structure
  27. Deep-learning-based polarization-dependent switching metasurface in dual-band for optical communication
  28. A nonlocal metasurface for optical edge detection in the far-field
  29. Coexistence of weak and strong coupling in a photonic molecule through dissipative coupling to a quantum dot
  30. Mitigate the variation of energy band gap with electric field induced by quantum confinement Stark effect via a gradient quantum system for frequency-stable laser diodes
  31. Orthogonal canalized polaritons via coupling graphene plasmon and phonon polaritons of hBN metasurface
  32. Dual-polarization electromagnetic window simultaneously with extreme in-band angle-stability and out-of-band RCS reduction empowered by flip-coding metasurface
  33. Record-level, exceptionally broadband borophene-based absorber with near-perfect absorption: design and comparison with a graphene-based counterpart
  34. Generalized non-Hermitian Hamiltonian for guided resonances in photonic crystal slabs
  35. A 10× continuously zoomable metalens system with super-wide field of view and near-diffraction–limited resolution
  36. Continuously tunable broadband adiabatic coupler for programmable photonic processors
  37. Diffraction order-engineered polarization-dependent silicon nano-antennas metagrating for compact subtissue Mueller microscopy
  38. Lithography-free subwavelength metacoatings for high thermal radiation background camouflage empowered by deep neural network
  39. Multicolor nanoring arrays with uniform and decoupled scattering for augmented reality displays
  40. Permittivity-asymmetric qBIC metasurfaces for refractive index sensing
  41. Theory of dynamical superradiance in organic materials
  42. Second-harmonic generation in NbOI2-integrated silicon nitride microdisk resonators
  43. A comprehensive study of plasmonic mode hybridization in gold nanoparticle-over-mirror (NPoM) arrays
  44. Foundry-enabled wafer-scale characterization and modeling of silicon photonic DWDM links
  45. Rough Fabry–Perot cavity: a vastly multi-scale numerical problem
  46. Classification of quantum-spin-hall topological phase in 2D photonic continuous media using electromagnetic parameters
  47. Light-guided spectral sculpting in chiral azobenzene-doped cholesteric liquid crystals for reconfigurable narrowband unpolarized light sources
  48. Modelling Purcell enhancement of metasurfaces supporting quasi-bound states in the continuum
  49. Ultranarrow polaritonic cavities formed by one-dimensional junctions of two-dimensional in-plane heterostructures
  50. Bridging the scalability gap in van der Waals light guiding with high refractive index MoTe2
  51. Ultrafast optical modulation of vibrational strong coupling in ReCl(CO)3(2,2-bipyridine)
  52. Chirality-driven all-optical image differentiation
  53. Wafer-scale CMOS foundry silicon-on-insulator devices for integrated temporal pulse compression
  54. Monolithic temperature-insensitive high-Q Ta2O5 microdisk resonator
  55. Nanogap-enhanced terahertz suppression of superconductivity
  56. Large-gap cascaded Moiré metasurfaces enabling switchable bright-field and phase-contrast imaging compatible with coherent and incoherent light
  57. Synergistic enhancement of magneto-optical response in cobalt-based metasurfaces via plasmonic, lattice, and cavity modes
  58. Scalable unitary computing using time-parallelized photonic lattices
  59. Diffusion model-based inverse design of photonic crystals for customized refraction
  60. Wafer-scale integration of photonic integrated circuits and atomic vapor cells
  61. Optical see-through augmented reality via inverse-designed waveguide couplers
  62. One-dimensional dielectric grating structure for plasmonic coupling and routing
  63. MCP-enabled LLM for meta-optics inverse design: leveraging differentiable solver without LLM expertise
  64. Broadband variable beamsplitter made of a subwavelength-thick metamaterial
  65. Scaling-dependent tunability of spin-driven photocurrents in magnetic metamaterials
  66. AI-based analysis algorithm incorporating nanoscale structural variations and measurement-angle misalignment in spectroscopic ellipsometry
Heruntergeladen am 22.1.2026 von https://www.degruyterbrill.com/document/doi/10.1515/nanoph-2025-0500/html
Button zum nach oben scrollen