Open Access
18 March 2021 Kinetic inductance detectors for the Origins Space Telescope
Author Affiliations +
Abstract

The Origins Space Telescope (Origins) will have a 5.9-m diameter primary mirror cooled to 4.5 K and will be equipped with three instruments, two of which will cover the far-IR (λ  =  25 to 588  μm). These far-IR instruments will require large arrays (∼104 detectors) of ultrasensitive detectors, with noise equivalent powers (NEPs) as low as 3  ×  10  −  20  W Hz  −  1/2. Kinetic inductance detectors (KIDs) have already demonstrated the array format, modularity, and readout multiplexing density requirements for Origins; the only aspect that requires improvement is the per-pixel sensitivity. We show how KIDs can meet the sensitivity target, focusing on two existing architectures that together demonstrate the key necessary attributes. Arrays of antenna-coupled coplanar waveguide resonators have achieved NEPs of 3  ×  10  −  19  W Hz  −  1/2 in laboratory demonstrations at 350  μm; they demonstrate excellent material properties as well as array-level integration and performance. Lumped element detectors such as those under development for balloon-borne spectroscopy at 10 to 350  μm demonstrate flexibility in coupling to shorter-wavelengths, reducing active volume, and providing a means for suppressing capacitor noise. A straightforward combination of the elements of these already-demonstrated devices points to a low-volume design that is expected to meet the Origins sensitivity targets.

1.

Introduction

The Origins Space Telescope (Origins) traces our cosmic history, from the formation of the first galaxies and the rise of metals to the development of habitable worlds and present-day life. Origins does this through exquisite sensitivity to infrared radiation from ions, atoms, molecules, dust, water vapor, and ice, and observations of extra-solar planetary atmospheres, protoplanetary disks, and large-area extragalactic fields. Origins operates in the wavelength range 2.8 to 588  μm and is more than 1000 times more sensitive than its predecessors due to its large, cold (4.5 K) telescope and advanced instruments.

Origins will have two instruments operating at λ>25  μm, a broadband imager (Far-IR Imager and Polarimeter; FIP, Staguhn et al., this volume) and an R=300 spectrometer (Origins Survey Spectrometer; OSS, Bradford et al., this volume). The significant improvement over the scientific capabilities of prior far-IR missions is based on the cold telescope (4.5 K) combined with low-noise far-IR detectors. The detector noise equivalent power (NEP) targets for imaging and spectroscopy are 3×1019  WHz1/2 and 3×1020  WHz1/2, respectively.1 These values render the detector noise negligible compared with the photon noise from the solar-system and galactic dust that ultimately limits far-IR measurements. The spectroscopy NEP target, in particular, has no application outside of a cryogenic space telescope, so improvement over the state-of-the-art is required; initial steps are underway and results are very promising. In addition, the array formats for both the imager and spectrometer are larger than anything flown thus far, though tractable when scaled from ground-based and suborbital experiments. The Origins far-IR detector needs are summarized in Table 1, and additional details can be found in the Origins 2019 Technology Development Plan.3

Table 1

Origins far-IR detector array targets and SPACEKIDs demonstrated performance.

ParameterOrigins-FIP targetOrigins-OSS targetSPACEKIDs demo
λ (μm)50 to 250 (in two bands)25 to 588 (in six bands)350
NEP (WHz1/2)3×10193×10203×1019a
Tile size (pixels)1 kilopixel + mosaicking to 104961
Multiplexing2000 pixels per circuit, 4-GHz BW available961 from 3.9 to 5.6 GHz
Minimum pitch (mm)0.50.4×0.71.6
τdet (ms)<3<31.5
Min. yield70%70%83%
Crosstalk (dB)<17<17<30
1/f knee (Hz)<0.1<0.1<0.05a
Cosmic ray dead time<10%<10%<5%

aFor the SPACEKIDs detectors, the frequency NEP is 3×10−19  W  Hz−1/2 at 60 to 80 Hz audio frequency, but is dominated by 1/f noise; the dissipation NEP is 6×10−19  W Hz−1/2 with a <0.05  Hz knee and is limited by readout electronics.2

A number of superconducting detector technologies are under consideration for the far-IR instruments on Origins, including transition-edge-sensed (TES) bolometers (Sadlier et al., this volume), kinetic inductance detectors (KIDs), and quantum capacitance detectors (Echternach et al, this volume). Here, we discuss the prospects of KIDs to meet the requirements shown in Table 1.

2.

KIDs: A Brief Introduction

The technology for semiconductor infrared array detectors is well developed, with silicon impurity band conduction detectors offering excellent performance in 1  k×1  k formats out to λ=28  μm.4 Stressed gallium-doped germanium photoconductors have been used out to λ=200  μm, e.g., on Spitzer/MIPS5 and Herschel/PACS;6 however, the difficulty of scaling this technology and the need for detectors beyond 200  μm has driven strong interest in superconducting detectors.79 At present, two superconducting detector technologies have been demonstrated in large array formats at millimeter through far-IR wavelengths: TES bolometers1012 and KIDs.13,14

In contrast to TES bolometers, which operate at the superconducting transition temperature Tc, KIDs operate at temperatures well below Tc. As a result, almost all of the conduction electrons in a KID are in the form of Cooper pairs. Unlike electrons in a normal metal, the Cooper pairs do not suffer scattering and can therefore carry current with no dissipation, leading to the hallmark property of superconductivity, the vanishing of the DC resistance. Indeed, the DC resistance of a KID is identically zero; a KID functions by sensing the (nonzero) AC impedance of the superconductor15 – the “kinetic inductance” – at RF or microwave frequencies. Meanwhile, a TES bolometer operates by sensing the temperature-dependent DC resistance at the onset of superconductivity, where the resistance is an appreciable fraction of its normal-state value.

Superconductivity has an energy scale and therefore a limited frequency range. The pairs have a binding or “gap” energy of Eg=2Δ3.5kBTc, corresponding to a photon frequency of νg=2Δ/h75  GHz×(Tc/1  K). This “gap frequency” is the minimum energy required for a photon to break a Cooper pair. At frequencies below νg, a single photon cannot break a Cooper pair, so the material cannot efficiently absorb energy and thus behaves as a superconductor. However, for frequencies above νg the material can readily absorb energy and behaves as an ordinary (dissipative) normal metal.

KIDs are microwave resonators that use both of the behaviors described above to detect optical radiation. One approach is to construct the resonator using a lumped-element circuit (the lumped element KID; LEKID),16 as shown in Fig. 1. As illustrated in Fig. 1(a), a simple meandered thin-film superconducting trace can serve as an efficient radiation absorber for frequencies above the gap, νg, provided the meander spacing, area filling factor, and normal-state sheet resistance of the superconductor are chosen appropriately. Simultaneously, the meander behaves as a low-loss superconducting inductor for frequencies below the gap, νg. For an ordinary inductor, the stored energy is given by LI2/2, where I is the current. For a KID, the inductance L has two components: the ordinary magnetic inductance Lmagnetic corresponding to energy in the magnetic field and the kinetic inductance Lkinetic corresponding to the kinetic energy of the Cooper pairs carrying the current. Like the energy in the magnetic field, the kinetic energy represents stored energy that can be recovered because the pairs do not scatter and the kinetic energy is not dissipated. By combining the meandered inductor with a simple interdigitated capacitor, an LC circuit is formed whose resonant frequency is given by νr=1/(2πLC). The resonance can be very sharp (high Q) because the inductor is superconducting and has very low loss; the resonance frequency, typically chosen to be a few GHz or lower, is far below the gap frequency, νrνg.

Fig. 1

(a) Close-up of a simple LEKID consisting of a meandered inductor (L) that doubles as a radiation absorber, and an interdigitated capacitor (CIDC). These form an LC resonant circuit that is weakly coupled to a common readout line. Reprinted with permission from Ref. 17. (b) The 1-GHz resonant frequency of the KID shifts downward in response to the intensity of the radiation. These curves show measured data for a KID illuminated with blackbody radiation passed through a λ=200  μm bandpass filter. The intensity is controlled by the blackbody temperature, which was varied from 5 K to 27.5 K. Note that a 10-kHz shift represents a fractional change of 10  kHz/1  GHz=105. Credit: P. K. Day. (c) A 432-pixel, λ=350  μm KID array, an early prototype for the MAKO instrument.18,19 The array is read out using frequency multiplexing. Readout signals are injected and extracted using two coaxial connectors; the readout line is highlighted in white. Credit: C. McKenney and H. G. Leduc. (d) A frequency sweep of the MAKO prototype array reveals 415 of 432 resonances with >6  dB depth, corresponding to a fabrication yield of 96%. Credit: C. McKenney. (e) Example of first-generation (ROACH-1) KID readout electronics.20 The circuit boards in the foreground contain fast digital-to-analog converters that generate the multifrequency analog signal that excites the KID array, and the analog-to-digital converters that digitize the return signal from the KID array. The FPGA, illuminated and underneath the cooling fan in the background, performs the digital signal processing needed to separate the individual KID frequencies.

JATIS_7_1_011015_f001.png

An alternative to the LEKID is the transmission-line resonator.13 One common implementation is the quarter-wave coplanar waveguide (CPW) design, an example of which is shown in Fig. 2. The inductance and capacitance per unit length of the CPW transmission line are L and C, respectively. The resonance frequency is then set by the length of the segment, l, as νr=1/(4lLC). Unlike a LEKID, the CPW resonator does not directly absorb optical radiation but uses an antenna to deposit optical photons at the shorted end of the resonator. An advantage of this optical coupling scheme is that the antenna and detector are decoupled and can therefore be optimized individually. However, designs that require a transmission line to couple the radiation from the antenna to the KID become unacceptably lossy at optical frequencies above the gap frequency (e.g., νg=1.1  THz for NbTiN). A design that directly couples the antenna to the KID has been shown to extend the optical frequency range up to 2.8 THz, at the expense of an increased antenna complexity and tighter fabrication tolerances.21 The typical dimensions of CPW resonators result in the CPW KIDs being larger and operating at higher readout frequency than LEKIDs.

Fig. 2

SPACEKIDs detector figures from Ref. 2. Top: Photographs of the 961 pixel demonstration array highlighting (a) the TiN absorbing mesh located between the detector wafer and lens array, (b) the quarter-wave CPW hybrid NbTiN-Al KID and the coupled twin-slot antenna, and (c) the fully assembled detector package. Bottom: Measured NEP of a representative detector for phase and amplitude readout. Credit: Ref. 2, reproduced with permission © ESO.

JATIS_7_1_011015_f002.png

Because the kinetic inductance derives from the motion of Cooper pairs, it is sensitive to the pair density. Meanwhile, absorption of above-gap radiation causes the breaking of Cooper pairs, which reduces their density and increases the kinetic inductance. In turn, the increased kinetic inductance causes the resonance frequency to shift downward. This is the fundamental detection mechanism for KIDs and is illustrated using experimental data in Fig. 1. In addition, the breaking of pairs by the absorbed radiation also increases the density of single electrons, which do scatter, increasing the (AC) resistive dissipation. This broadens the resonances and reduces their depth, as evident in Fig. 1. Either of these two responses—the shift in the frequency νr or the dissipation Qr1 of the resonance—may be used for KID readout. This is the origin of terms such as “frequency readout” or “dissipation NEP” used in this paper and in the literature; the terms “phase readout” and “amplitude NEP” are used synonymously.

The narrow resonance produced by a single KID leads to the simple scheme for frequency-domain multiplexing shown in Fig. 1. For LEKIDs, the capacitor geometry is adjusted so each KID has a unique resonance frequency; for CPW KIDs, this is achieved by adjusting the length of each resonator. The readout electronics generate a superposition of frequencies to excite the array and then separate these frequencies in the output signal from the array to isolate the response of individual pixels. This task is now tractable as a result of advances in electronics, especially digital signal processing using field programmable gate arrays (FGPAs). Figure 1(e) shows an early example of KID readout electronics.

The deployments of DemoCam,22,23 MUSIC,2426 and MAKO18 (Fig. 1) at the Caltech Submillimeter Observatory, as well as NIKA27 on the IRAM 30-m telescope, represent early demonstrations of prototype KID arrays. Since then, KIDs have enabled a number of large (multi-kilopixel) ground-based millimeter-wave and submillimeter-wave cameras, including NIKA2,2830 and A-MKID on the APEX 12-m telescope,31,32 with TolTEC33,34 soon to be deployed on the 50-m Large Millimeter Telescope (LMT) in Mexico. A new 850-μm camera employing TolTEC-style KIDs is under development for the EAO/JCMT 15-m telescope on Mauna Kea,35,36 to replace the SCUBA-2 TES camera37 in several years. Intensity mapping38 using the [C II] 158  μm fine structure line (redshifted to 200 to 300 GHz at z=58) is a major goal for KID-based millimeter-wave, ground-based imaging spectrometers under development, including SuperSpec,39 DESHIMA,40 and CONCERTO.41 Balloon payloads using KIDs are being developed, e.g., for the Terahertz Intensity Mapper (TIM)42 and EXCLAIM,43 and have now been flown on OLIMPO44 and BLAST-TNG.45 KID-based millimeter-wave cameras are now also being developed for other applications, e.g., security46 and maritime surveillance.47

The NEP requirements for Origins are more demanding than for any existing instruments and necessitate dedicated detector development. A European consortium has executed a EU Seventh Framework Programme (FP7) project, named SPACEKIDs, explicitly targeting the very low NEP for space astrophysics missions. The SPACEKIDs demonstration meets the sensitivity requirement of the FIP instrument (Table 1) and achieves a high level of system maturity.2 A separate effort is underway to develop KIDs for Origins based on an extension of the TIM detector design.48 The prospects of meeting the Origins detector goals with these two approaches are discussed below.

3.

Low NEP Demonstration of the SPACEKIDs Array

The KID array that most closely meets the Origins requirements is the array developed in the EU-funded SPACEKIDs program,2 a 5-year multinational research activity specifically targeting the needs of future far-IR space missions in astrophysics and earth science. The demonstration array is a 31×31 (961) pixel detector array optimized for imaging at 850 GHz and is coupled to a readout electronics system49 developed under the SPACEKIDs program.

The basic pixel is an antenna-coupled hybrid NbTiN-Al device implemented as a quarterwave CPW resonator, shown in detail in Fig. 2. The hybrid design combines the low noise properties of a NbTiN resonator with the excellent responsivity of aluminum, which has been shown to produce quasiparticle lifetimes >1.5  ms.50,51 The CPW resonator is 6.3  mm long, producing a resonant frequency fr4.7  GHz. For 75% of its length the resonator is fabricated from 500-nm thick NbTiN with large features (the linewidth and slotwidth are both 20  μm), where both the choice of NbTiN52 and the large features are designed to minimize noise from two-level system (TLS) fluctuators on the surfaces of the capacitive portions of the device.53,54 Due to the high Tc (Tc=15  K), the NbTiN film provides only a reactive load to optical frequencies <1.1  THz. For the rest of the length, the central conductor is a narrow (2  μm linewidth) section of aluminum, which is expected to absorb >95% of the optical power provided by a twin-slot antenna. The antenna is fed by an antireflection-coated silicon microlens, and the end-to-end single-polarization optical efficiency is estimated to be 74%.

All 961 resonators in the array are capacitively coupled to a single CPW readout line at the open end of the resonator. The SPACEKIDs readout system is capable of reading up to 4000 pixels in a 2-GHz bandwidth centered anywhere between 4.5 and 7.5 GHz.49 Crosstalk levels (both optical and electrical) <30  dB are achieved by implementing aluminum bridges over the CPW readout line, intelligent positioning of neighboring resonators in frequency space, and a substoichiometric TiN (Tc=0.6  K) absorbing mesh on the backside of the sapphire wafer. The latter also assists in the suppression of cosmic ray effects,55 by absorbing the athermal phonons generated by cosmic ray events in the substrate.

The NEP of a typical KID in the array is shown in Fig. 2. At low temperature (120 mK) and under low optical load (10 aW), the frequency noise is dominated by TLS noise. The noise PSD has a nonwhite spectrum, and at high audio frequency (60 to 80 Hz), the NEP is 3×1019  WHz1/2. In the amplitude quadrature, the white NEP is 6×1019  WHz1/2 and is dominated by a combination of the cold and warm readout electronics. The 1/f knee in the amplitude noise PSD is <0.05  Hz.

The NEP of these detectors could be further reduced through an increase in the responsivity or a decrease in the electrical noise. Increasing the responsivity through a longer quasiparticle lifetime is not a practical solution, given the (τdet<3  ms) requirement (Table 1). A more plausible path forward50 would be to decrease the VL130  μm3 active volume of the detector. Reductions to the TLS noise could potentially be achieved by optimizing the capacitor geometry and choice of substrate material52 and/or the removal of exposed substrate surfaces in high-field regions.56 Reducing the dissipation NEP could be achieved by optimizing the coupling to the readout line, which increases the KID amplitude noise with respect to the readout noise. Reference 2 suggested such changes could reduce the NEP down to a few 1019  WHz1/2 down to 0.1 Hz, but further reduction would demand more significant measures.

In Table 1, we compare the performance of the SPACEKIDs system with some of the key goals for the Origins detectors. The SPACEKIDs system achieves the desired NEP for the FIP instrument, as well as many other key requirements including the achieved tile size, multiplexing density, and cosmic ray dead time.

4.

Low NEP Lumped Element KIDs

Direct-absorbing lumped element KIDs are promising candidates for the far-IR instruments on Origins. As described below, a model for an Origins appropriate detector leverages the demonstration of a feedhorn-coupled, low-volume aluminum device at 350  μm. Further reductions to the inductor volume may be implemented to reduce the NEP below 1019  WHz1/2, and modifications to the absorber will allow operation over the full far-IR band.

4.1.

Aluminum LEKIDs for TIM

TIM is a balloon-borne spectrometer that will operate at 240 to 420  μm, with a required detector NEP of 1×1017  WHz1/2.42 The detectors will be horn-coupled LEKIDs, shown in Fig. 3, and described more fully in Ref. 48.

Fig. 3

TIM prototype detector. (a) Microscope image of a single pixel within a 45-pixel prototype array. The meandered inductor is surrounded by a set of optical choke rings and connected to a large interdigitated capacitor to form a 250-MHz LEKID. The inductor/absorber is illuminated by a conical feedhorn coupled to a circular waveguide. (b) Amplifier-subtracted Sxx for a representative detector measured under various optical loads at 350  μm. At low loading the white noise corresponds to a detector NEP of 4×1018  WHz1/2. Figure reprinted with permission from the Journal of Low Temperature Physics.

JATIS_7_1_011015_f003.png

The absorber cross-section is limited to 300  μm in diameter to minimize the detector volume, whereas the large capacitor occupies 50% of the pixel area to decrease the capacitor (TLS) noise. A conical feedhorn coupled to a circular waveguide is used to concentrate light on the absorber, maintaining a high filling factor, 2.3-mm pitch hexagonal-packed focal plane. The inductor/absorber is a single meander of 0.4-μm wide aluminum, patterned to provide an optimal impedance match to the horn’s output waveguide. Segments of meander line are placed in close proximity at the corners, creating capacitive shorts at the optical frequencies that cause the meander to behave as a square mesh and thereby couple efficiently to both polarizations. This is achieved with a 0.3-μm gap and a 0.6-μm overlap length for each of the corners. The pixel is fabricated on a silicon on insulator wafer and has a 27-μm thick backshort, created by etching from the backside to a buried oxide layer, then depositing gold. Electromagnetic simulations indicate band-averaged (790 to 900 GHz) coupling efficiencies of 90%.57 The interdigitated capacitor has 1-μm wide fingers with 2-μm gaps. When patterned from a 40-nm film, the total inductor volume is 76  μm3 and the resonant frequency is 250  MHz.

Initial tests were conducted with a 45-pixel prototype array of TIM detectors patterned on a standard silicon wafer, excluding the backshort.48 Figure 3 shows the noise spectrum for a typical KID measured dark and with optical loading, at a 210-mK base temperature. At 30 fW of optical loading the detector NEP is 4×1018  WHz1/2, with a 1/f knee at 0.5  Hz.

4.2.

Ultralow NEP Far-IR LEKIDs

The TIM detector design presented in Sec. 4.1 may be modified to produce an increased responsivity and reduced NEP through a combination of a lower operating temperature, increased τqp and reduced inductor volume.

4.2.1.

Increasing the responsivity

The fractional frequency responsivity of a KID, Rx, is given as

Eq. (1)

Rx=αγS2(ω)4N0Δ0ηoτqpΔ0VL,
where α is the kinetic inductance fraction, γ=1 for thin films, ηo is the pair-breaking efficiency, N0 is the density of states, Δ0 is the gap energy, VL is the inductor volume, and S2(ω) relates changes in the kinetic inductance to changes in the quasiparticle number density.14 For a superconductor in thermal equilibrium, the (nqpτqp) product is a material-specific constant

Eq. (2)

nqpτqp=N0τ0(kBTc)32Δ2,
where τ0 is the characteristic electron-phonon interaction time.58 The quasiparticle lifetime has been observed to saturate at a maximum value τmax in the low temperature limit of some dark resonators,59,60 leading to a commonly applied empirical parameterization

Eq. (3)

τqp=τmax1+nqp/n*,
for some critical n* value. Consistency between Eqs. (2) and (3) in the high nqp limit requires

Eq. (4)

n*τmax=N0τ0(kBTc)32Δ2.

The τqp inferred from the noise measurements of the TIM devices is consistent with Eqs. (3) and (4) with τmax=35  μs and τ0=139  ns. This τmax value is unexpectedly short; τqp500  μs is commonly achieved in aluminum resonators operated below 150 mK in dark conditions.50,59,61,62 While the shorter τqp achieved in the TIM devices may be in part due to the higher operating temperature, disorder in the material is potentially a more important factor;60 this may be addressed by optimizing the fabrication process. Quarter-wavelength CPW devices recently fabricated at JPL yielded quasiparticle lifetimes of 1 to 5 ms for film thicknesses of 20 to 50 nm.63 We are therefore optimistic the fabrication may be optimized to achieve τmax=1  ms. Achieving an aluminum film with τmax=1  ms and τ0=139  ns would lead to τqp=992  μs in a dark device at 150 mK.

The responsivity may be further increased by reducing the inductor/absorber volume. The geometry currently used provides an impedance match to the incident radiation over the full waveguide cross-section. However, a single linear structure cutting across the center of the circular footprint has been shown to couple well to the fundamental waveguide mode64 and has a much smaller volume. We have designed a dual-polarization sensitive absorber that achieves a simulated peak absorption efficiency of 90% and reduces the inductor volume from 76 to 10  μm3 for a 20-nm film thickness (Fig. 4). This absorber uses a 150-nm linewidth, which pushes the limits of UV lithography but may be easily achieved with electron-beam lithography. Due to the narrower linewidth, the inductance of this lower volume absorber is close to that in the current TIM device, so may be implemented while keeping the readout frequency constant at 250  MHz.

Fig. 4

(a) Geometry for a dual-polarization, minimal volume aluminum absorber coupled to a circular waveguide. (b) Zoom-in of the center of the absorber, showing the metal film (green) on the silicon substrate (blue). The linewidth is 150 nm and the total volume is 10  μm3 for a 20-nm film thickness. (c) HFSS-simulated absorption efficiency, indicating a peak of 90% at 360  μm. Credit: P. K. Day.

JATIS_7_1_011015_f004.png

4.2.2.

Noise model

The dark noise measurements of the TIM detectors indicate a limiting noise level of Sxx,0=1.8×1017  Hz1 (Fig. 3). While the source of this noise floor is not yet fully understood, we may adopt it as an upper limit to the TLS noise at 210 mK and 10 Hz audio frequency. These noise measurements were obtained with a drive power of Pg110  dBm, which is 5  dB below bifurcation. Under these conditions, the resonators are characterized by QiQc105.

For a fixed capacitor geometry, the TLS noise has a well-characterized dependence on operating temperature and internal power: Sxx,TLSPint1/2T1.7,65 where Pint=(2Qr2Pg)/(πQc). As we consider changing the drive power and temperature in our proposed low volume KIDs, we therefore assume the TLS noise will be

Eq. (5)

Sxx,TLS=1.8×1017  Hz1(4Qr2105Qc)1/2(Pg1014  W)1/2(T210  mK)1.7.
The bifurcation power will be reduced with the smaller volume.66 Additionally specifying that we maintain the drive power at least 5 dB below bifurcation leads to the constraint

Eq. (6)

Pg1014  W(fr250  MHz)(VL76  μm3)(Qc105)(1052Qr)3.
The fractional frequency noise contributed by an LNA with noise temperature Tn and zero detuning may be written as14

Eq. (7)

Sxx,LNA=4χc2Qi2kBTnPg,
with

Eq. (8)

χc=4QcQi(Qc+Qi)2.
We assume Tn=5  K.67

Generation-recombination noise is given as14

Eq. (9)

NEPGR2=2hνPo(1+n0)+4ηaχqpΔ0ηo2Pa+4Δ02η02(Γth+Γr),
where η00.57 is the pair-breaking efficiency, Pa is the absorbed readout power, ηa is the efficiency with which absorbed readout power is converted to quasiparticles, χqp=Qi/Qqp, and Γth and Γr are the thermal generation and total quasiparticle recombination rates, respectively. The four terms in Eq. (9) represent photon generation noise, readout power generation noise, thermal generation noise, and recombination noise, respectively.

4.2.3.

NEP optimization

Given the noise terms described in Sec. 4.2.2, our low volume KID may be optimized to achieve the minimum NEP for a given optical loading. Here, we consider this optimization assuming a Po=1018  W load at λ=350  μm, as appropriate for OSS. We assume VL=10  μm3 and vary T and Pg. We assume our devices maintain Qi=Qc=105.

A considerable uncertainty in this optimization is the pair-breaking efficiency of the readout power, ηa. The absorption of microwave photons will drive the quasiparticle distribution out of thermal equilibrium and increase the number density nqp.50,68,69 These excess quasiparticles will reduce the responsivity, and this absorption will be the source of an additional shot noise term. Reference 50 measures ηa103 for a fr=5.3  GHz resonator with a similar readout power density (between 1016 and 1015  Wμm3) as envisioned for our low volume devices, whereas Ref. 69 calculates ηa0.10.5 in simulations of a fr=4  GHz resonator with absorbed power densities between 1019 and 1014  Wμm3. This efficiency may be expected to depend on the resonator frequency and will likely be different given the lower frequency (fr250  MHz) considered here. We choose to adopt ηa=0.1 in our modeling, near the upper end of the above estimates.

In Figs. 5 and 6, we show the contributions of the various noise terms as a function of T and VL. For the TLS noise value adopted here, an operating temperature of 150 mK represents a good balance between rising TLS noise at lower temperatures and rising thermal GR noise at higher temperatures (Fig. 5). For VL=10  μm3, the total NEP is minimized by maximizing the readout power, subject to the limit described in Eq. (6). The detector noise is NEPdet=4×1020  WHz1/2 and is dominated by TLS noise. This model is summarized in Table 2. We note that further reductions to the inductor volume will reduce the NEP (Fig. 6). Such reductions are possible with the absorber shown in Fig. 4 by reducing the linewidth below 150 nm.

Fig. 5

Sensitivity model predictions for the low volume LEKID proposed here, with ηa=0.1, VL=10  μm3, and maximized Pg (see text for details). Detector NEP (black dashed) includes all contributions to the NEP aside from the photon generation noise. The optimal operating temperature is T150  mK.

JATIS_7_1_011015_f005.png

Fig. 6

Same as Fig. 5 with T=150  mK. With an optical load of Po=1018  W, the photon generation noise is NEPphoton=3.4×1020  WHz1/2, and the detector noise with VL=10  μm3 is NEPdet=4.4×1020  WHz1/2. Further reductions in the detector NEP are possible by further reducing VL.

JATIS_7_1_011015_f006.png

Table 2

Proposed low NEP designs.

Existing 350  μmProposed low-NEP lumped element
ParameterSPACEKIDsTIM lumped element350  μm25  μm
Temp (mK)120210150150
τmax (μs)14803510001000
Linewidth (nm)2000400150200
VL (μm3)130761054
fres (MHz)47002502501700
Pitch (mm)1.62.32.30.4
α (kin. ind. fraction)0.090.760.760.76
Rx (W1)4×1091.2×1093.1×10115.4×1010
Sxx,TLS (Hz1)1.4×10181.8×10179.0×10173.4×1016
NEPdet (WHz1/2)3×10194×10184.4×10203.5×1019
Notes1234
Note: 350  μm designs include (1) the SPACEKIDs device as shown in Fig. 2,2 (2) the fabricated/characterized TIM detector, (3) the same TIM device with an increased τmax, reduced temperature, and reduced inductor volume under a 1-aW optical load. (4) 25  μm design adapts the low volume 350  μm device to optically couple at shorter wavelengths, evaluated at 0.5-aW load.

4.3.

Ultralow NEP LEKIDs at 25  μm

The TIM detectors described in Sec. 4.1 may be adapted to operate at shorter wavelengths with the use of a different absorber geometry. The wire grid absorber used in the 350  μm detectors is 0.4  μm wide; scaling this design with wavelength would require prohibitively narrow lines at λ<100  μm. An alternative approach is to interrupt the absorber line with a meander that increases the resistance per unit length while simultaneously introducing a distributed capacitance to compensate for the accompanying distributed inductance.70 An impedance match to silicon may then be maintained with much wider lines. This effort is complementary to and also driven by the mid-IR KID development for the Galaxy Evolution Probe, which requires detectors down to 10  μm wavelength.71

This concept is shown in Fig. 7. The absorber couples to vertically polarized radiation and is designed to be backside-illuminated through the silicon substrate. For a 10-μm wavelength configuration, the linewidth is 200 nm and the unit cell is 2.4  μm on a side. Agilent HFSS simulations indicate that this geometry achieves a peak single-polarization absorptivity of just over 70% near 10  μm, and performs well at wavelengths beyond 25  μm, the shortest wavelength required for Origins. This linewidth can be fabricated with a UV stepper, and further optimization targeting 25  μm would allow even larger linewidths. In addition, Ref. 70 presented a resonant dual-polarization design for which simulations yield absorption of 70% in each polarization.

Fig. 7

Photograph of the inductor/absorber portion of a 10-μm LEKID. The inset shows the basic unit cell, optimized for the absorption of vertically polarized radiation, replicated over the full 60-μm diameter absorber area. Simulations indicate a peak single-polarization absorption efficiency of 70% near 10  μm. Figure duplicated unaltered from Ref. 70; license can be found at Ref. 72.

JATIS_7_1_011015_f007.png

Coupling these absorbers to free space radiation is achieved with microlenses. Microlenses are commercially available at near-infrared wavelengths, but their substrates do not have low absorption at mid-infrared wavelengths. Thus, they should be made of silicon for wavelengths >20  μm (and likely germanium for 10 to 20  μm because of the silicon absorption features in the 14 to 16  μm range). Microlens arrays could either be fabricated separately and bonded to backilluminated KID arrays73 or etched into the backside before or after KID fabrication. There are numerous examples of silicon microlens fabrication in the literature that could achieve the μm-level tolerances required for operation down to 25  μm,7480 although those involving micromachining and laser etching could be slow for arrays of 104 detectors. Another promising approach is the deposition of micro-Fresnel plate lenses.81 This approach has the virtue of fabrication simplicity but rejects 50% of the radiation outright.

Reference 70 presented dark measurements of an array of KIDs with an absorber as shown in Fig. 7. The aluminum thickness was 40 nm, resulting in an inductor volume of 54  μm3. The IDCs were smaller than those used in the TIM detectors, reducing the pitch to 0.4 mm and pushing the readout frequencies up to the 1.4 to 2.0 GHz range. The frequency noise was dominated by TLS noise, with typical values at 10 Hz of Sxx=(30.6)×1016  Hz1 at T=100 to 200 mK, consistent with the expected T1.7 scaling. Assuming the same film properties as in Sec. 4.2.1 allows an estimate of the NEP. For a 0.5-aW load (appropriate for OSS at 25  μm), the predicted responsivity is Rx=5.4×1010W1, and the detector noise is strongly limited by TLS with an NEP of 4×1019  WHz1/2. These parameters are included in Table 2.

In summary, lumped element KIDs appear to be viable for Origins at wavelengths as short as 25  μm. Simulations indicate good optical efficiency for microlens-coupled absorbers, and dark noise measurements indicate NEPs matching the FIP instrument goal.

5.

Detector Array Requirements

In addition to the individual detector NEP goals, there are a number of array-level requirements for FIP and OSS summarized in Table 1. Many of these requirements (e.g., tile size and multiplexing density) are met with current laboratory and suborbital demonstrators, whereas others (e.g., the broad spectral range, small pixel pitch, and cosmic ray suppression) represent technical challenges.

5.1.

Array Size and Multiplexing Density

The FIP and OSS instruments will each have of order 105 detectors, in multiple arrays of up to 16,000 pixels each. The arrays will be tiled from 1 to 4 kilopixel subarray dies. Kilopixel arrays have been fielded on ground-based telescopes18,29 and larger arrays are near deployment (e.g., the TolTEC millimeter-wave camera with three arrays, including one that is 3.6 kilopixels34). The SPACEKIDs demonstration array has 961 pixels fabricated on a single wafer, and with stringent criteria on allowed crosstalk achieves a >83% yield.2

The digital readout electronics baselined for Origins adopt a multiplexing density of 2000 pixels in a 4-GHz band, or 500  detectors/GHz. The SPACEKIDs array is designed to place 961 channels in 1.7 GHz of bandwidth, a density of about 565  pixels/GHz. The potential for crosstalk between adjacent channels in readout space is minimized by designing for a high Q=105 and scheduling the readout frequencies to keep nearest neighbors in readout space at least one detector apart while maintaining a smooth variation in readout frequency over the wafer to minimize scatter due to film variations. A small fraction of devices is found to be close enough in readout space to exhibit crosstalk at the 30-dB level.

Operating at lower fr offers a natural multiplexing advantage, as for a fixed Q the bandwidth per resonator is correspondingly reduced.18 Additional improvements in multiplexing density are possible with LEKIDs using post-fabrication resonator trimming methods, which can adjust resonant frequencies after fabrication to eliminate frequency–space collisions arising from lithography errors and material variation.82,83

5.2.

Spectral Coverage and Pitch

The FIP and OSS instruments will cover λ=50  to  250  μm and λ=25  to  588  μm, respectively (Table 1). The SPACEKIDs detectors discussed in Sec. 3 are limited to operation at λ>270  μm due to the use of a NbTiN ground plane in the antenna. However, the use of thick aluminum as a ground plane allows operation at higher frequencies with only 10% additional loss.50 An extension of this design to higher frequencies and larger bandwidths has been demonstrated with the use of a leaky lens antenna, which couples to radiation over an octave of bandwidth from 107 to 215  μm.21 While the resonator design is complex, the KIDs used in this demonstration achieved a limiting NEP2.5×1019  WHz1/2. Extending this design down to 25  μm would require fabricating 300 nm features and maintaining a 1-μm air gap between the lens and antenna. As discussed in Sec. 4.3, the LEKIDs under development for 25  μm operation use a lens-coupled absorber with a 200 nm or larger linewidth. A version with 200 nm features has been successfully patterned with good yield and characterized dark.

Because FIP and OSS span the full far-IR band, the minimum required pixel pitches are smaller than in the existing prototypes designed for λ=350  μm. The pitch for FIP is 500  μm in both bands, whereas for OSS it is 400×700  μm at the shortest wavelength. This is significantly smaller than the 1.6 mm size of the quarter-wave CPW resonator used for the SPACEKIDs demonstrator and requires reducing the per-pixel area by a factor of 10. The 10-μm LEKID array fabricated in Ref. 70 achieves a 400-μm pitch with an area-efficient absorber/inductor design and small interdigitated capacitors. These small IDCs necessarily mean increasing the resonance frequency from 250 MHz to 1.7 GHz. However, this is still less than the 5  GHz frequency typical of CPW resonators and is consistent with the readout scheme of 2000 pixels per 4 GHz band. While we do not believe they are required, parallel plate capacitors may be useful for these small pitch designs by adding further design flexibility.

5.3.

Cosmic Ray Mitigation

Cosmic ray interactions with the detector substrate will produce high energy phonons, and the resulting glitches in the detector time streams can potentially impact a large fraction of the data.84 One way to mitigate this effect with KIDs is to deposit a low-Tc film on the substrate, with the expectation that high energy phonons entering this film will quickly down-convert to phonons with energies near 2Δ0 (where Δ0 is the gap energy). At such low energies, these phonons will not be capable of breaking Cooper pairs in the KID and will not produce a detector response.2,55,85

Reference 55 demonstrated the effectiveness of this technique with laboratory measurements using the same type of array as in the SPACEKIDs program. They found that the dead time due to cosmic ray glitches could be reduced by a factor of 40 by depositing a layer of Ta on the backside (with Tc=0.65  K compared with Tc=1.25  K in the Al). Crucially, this work showed that the dead time per pixel was independent of array size, such that further increases in the array would not increase the data loss rate. When accounting for the high event rate in an L2 orbit, these authors predict a per-pixel dead time of 1.4%. A comparable reduction in the dead time was found for devices in which the KID was placed on a thin membrane, isolating it from the larger substrate. In the SPACEKIDs demonstration array, a TiN mesh was employed to absorb stray light and act as a phonon trap, and the estimated dead time for an event rate appropriate for L2 was 4%.2

The OLIMPO balloon experiment conducted a stratospheric flight with arrays of lumped element KIDs operating at 150 to 460 GHz. As with the TIM prototypes, these KIDs use front-side illuminated, feedhorn-coupled absorbers with a metallized backshort. The backshort was made from 200 nm aluminum and doubles as a phonon trap (the Tc is slightly less than that in the 30-nm aluminum film used to pattern the KIDs). An analysis of the data conducted in flight resulted in <4% of the data contaminated by cosmic rays.44

6.

Summary

The far-IR detectors for the FIP and OSS instruments on the Origins Space Telescope will need to achieve extremely low NEPs over the 25 to 588  μm band and be deployed in kilopixel subarrays that may be suitably scaled up to arrays with 104 detectors. We address the current status for KIDs to meet these requirements, focusing on the antenna-fed CPW devices developed in the SPACEKIDs program, as well as lumped element devices based on an extension of the detectors under development for the TIM balloon experiment. The SPACEKIDs detector array has achieved an NEP of 3×1019  WHz1/2, suitable for imaging with FIP but not yet reaching the NEP goal of 3×1020  WHz1/2 for spectroscopy with OSS. The lumped element TIM detectors have achieved an NEP of 4×1018  WHz1/2, and reducing this to <1×1019  WHz1/2 should be possible with changes to the inductor/absorber designed to increase the responsivity. An array of lumped element KIDs designed to operate at 10  μm has been fabricated and characterized, achieving a dark NEP of 4×1019  WHz1/2.

Kilopixel subarrays with sufficiently high multiplexing density have been demonstrated on ground-based telescopes, and this requirement should not pose a challenge for Origins. The SPACEKIDs and TIM detectors have both been demonstrated at 350  μm, but simulations suggest both architectures can be adapted to work down to 25  μm. The small 500  μm pitch demanded by FIP and OSS represents a challenge for the SPACEKIDs CPW resonators, but a 400  μm pitch has been demonstrated with the lumped element design. Cosmic ray contamination will be an issue for any detector architecture, but hardening KID arrays using low Tc phonon traps promises to reduce the cosmic ray dead time to <5%.

Acknowledgments

Funding for this work was provided by the National Aeronautics and Space Administration (Grant No. 80NSSC19K0489). R. M. J. Janssen was supported by an appointment to the NASA Postdoctoral Program at the NASA Jet Propulsion Laboratory, administered by the Universities Space Research Association under contract with NASA. This research was carried out in part at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the National Aeronautics and Space Administration (80NM0018D0004). We thank the anonymous referees for many helpful comments on an earlier draft of this manuscript.

References

1. 

C. M. Bradford et al., “The origins survey spectrometer (OSS): a far-IR discovery machine for the Origins Space Telescope,” Proc. SPIE, 10698 1069818 (2018). https://doi.org/10.1117/12.2314049 PSISDG 0277-786X Google Scholar

2. 

J. Baselmans et al., “A kilo-pixel imaging system for future space based far-infrared observatories using microwave kinetic inductance detectors,” Astron. Astrophys., 601 A89 (2017). https://doi.org/10.1051/0004-6361/201629653 AAEJAF 0004-6361 Google Scholar

4. 

P. Love et al., “1024 x 1024 Si: as IBC detector arrays for JWST MIRI,” Proc. SPIE, 5902 590209 (2005). https://doi.org/10.1117/12.623473 PSISDG 0277-786X Google Scholar

5. 

G. Rieke et al., “The multiband imaging photometer for Spitzer (MIPS),” Astrophys. J. Suppl. Ser., 154 (1), 25 (2004). https://doi.org/10.1086/422717 APJSA2 0067-0049 Google Scholar

6. 

A. Poglitsch et al., “The photodetector array camera and spectrometer (PACS) on the Herschel space observatory,” Astron. Astrophys., 518 L2 (2010). https://doi.org/10.1051/0004-6361/201014535 AAEJAF 0004-6361 Google Scholar

7. 

J. Zmuidzinas and P. Richards, “Superconducting detectors and mixers for millimeter and submillimeter astrophysics,” Proc. IEEE, 92 (10), 1597 –1616 (2004). https://doi.org/10.1109/JPROC.2004.833670 IEEPAD 0018-9219 Google Scholar

8. 

P. Mauskopf, “Transition edge sensors and kinetic inductance detectors in astronomical instruments,” Publ. Astron. Soc. Pac., 130 (990), 082001 (2018). https://doi.org/10.1088/1538-3873/aabaf0 PASPAU 0004-6280 Google Scholar

9. 

D. Farrah et al., “Far-infrared instrumentation and technological development for the next decade,” J. Astron. Telesc. Instrum. Syst., 5 (2), 020901 (2019). https://doi.org/10.1117/1.JATIS.5.2.020901 Google Scholar

10. 

K. Irwin, “An application of electrothermal feedback for high resolution cryogenic particle detection,” Appl. Phys. Lett., 66 (15), 1998 –2000 (1995). https://doi.org/10.1063/1.113674 APPLAB 0003-6951 Google Scholar

11. 

A. Lee et al., “A superconducting bolometer with strong electrothermal feedback,” Appl. Phys. Lett., 69 (12), 1801 –1803 (1996). https://doi.org/10.1063/1.117491 APPLAB 0003-6951 Google Scholar

12. 

K. Irwin, G. Hilton, “Transition-edge sensors,” Cryogenic Particle Detection, 99 63 –150 Springer, Berlin, Heidelberg (2005). Google Scholar

13. 

P. K. Day et al., “A broadband superconducting detector suitable for use in large arrays,” Nature, 425 817 –821 (2003). https://doi.org/10.1038/nature02037 Google Scholar

14. 

J. Zmuidzinas, “Superconducting microresonators: physics and applications,” Annu. Rev. Condens. Matter Phys., 3 (1), 169 –214 (2012). https://doi.org/10.1146/annurev-conmatphys-020911-125022 ARCMCX 1947-5454 Google Scholar

15. 

D. Mattis and J. Bardeen, “Theory of the anomalous skin effect in normal and superconducting metals,” Phys. Rev., 111 (2), 412 (1958). https://doi.org/10.1103/PhysRev.111.412 PHRVAO 0031-899X Google Scholar

16. 

S. Doyle et al., “Lumped element kinetic inductance detectors,” in Eighteenth Int. Symp. Space Terahertz Technol., 170 (2007). Google Scholar

17. 

O. Noroozian, “Superconducting microwave resonator arrays for submillimeter/far-infrared imaging,” California Institute of Technology, (2012). Google Scholar

18. 

L. J. Swenson et al., “MAKO: a pathfinder instrument for on-sky demonstration of low-cost 350 micron imaging arrays,” Proc. SPIE, 8452 84520P (2012). https://doi.org/10.1117/12.926223 PSISDG 0277-786X Google Scholar

19. 

C. McKenney et al., “Design considerations for a background limited 350 micron pixel array using lumped element superconducting microresonators,” Proc. SPIE, 8452 84520S (2012). https://doi.org/10.1117/12.925759 PSISDG 0277-786X Google Scholar

20. 

R. Duan et al., “An open-source readout for MKIDs,” Proc. SPIE, 7741 77411V (2010). https://doi.org/10.1117/12.856832 PSISDG 0277-786X Google Scholar

21. 

J. Bueno et al., “Full characterisation of a background limited antenna coupled KID over an octave of bandwidth for THz radiation,” Appl. Phys. Lett., 110 233503 (2017). https://doi.org/10.1063/1.4985060 APPLAB 0003-6951 Google Scholar

22. 

J. Schlaerth et al., “A millimeter and submillimeter kinetic inductance detector camera,” J. Low Temp. Phys., 151 684 –689 (2008). https://doi.org/10.1007/s10909-008-9728-3 JLTPAC 0022-2291 Google Scholar

23. 

J. Schlaerth et al., “MKID multicolor array status and results from DemoCam,” Proc. SPIE, 7741 774109 (2010). https://doi.org/10.1117/12.857688 PSISDG 0277-786X Google Scholar

24. 

P. R. Maloney et al., “MUSIC for sub/millimeter astrophysics,” Proc. SPIE, 7741 77410F (2010). https://doi.org/10.1117/12.857751 PSISDG 0277-786X Google Scholar

25. 

J. A. Schlaerth et al., “The status of music: a multicolor sub/millimeter MKID instrument,” J. Low Temp. Phys., 167 347 –353 (2012). https://doi.org/10.1007/s10909-012-0541-7 JLTPAC 0022-2291 Google Scholar

26. 

J. Sayers et al., “The status of MUSIC: the multiwavelength sub-millimeter inductance camera,” Proc. SPIE, 9153 915304 (2014). https://doi.org/10.1117/12.2055444 PSISDG 0277-786X Google Scholar

27. 

A. Monfardini et al., “NIKA: a millimeter-wave kinetic inductance camera,” Astron. Astrophys., 521 A29 (2010). https://doi.org/10.1051/0004-6361/201014727 AAEJAF 0004-6361 Google Scholar

28. 

R. Adam et al., “The NIKA2 large-field-of-view millimetre continuum camera for the 30 m IRAM telescope,” Astron. Astrophys., 609 A115 (2018). https://doi.org/10.1051/0004-6361/201731503 AAEJAF 0004-6361 Google Scholar

29. 

A. Catalano et al., “The NIKA2 instrument at 30-m IRAM telescope: performance and results,” J. Low Temp. Phys., 193 916 –922 (2018). https://doi.org/10.1007/s10909-018-1884-5 JLTPAC 0022-2291 Google Scholar

30. 

L. Perotto et al., “Calibration and performance of the NIKA2 camera at the IRAM 30-m telescope,” Astron. Astrophys., 637 A71 (2020). https://doi.org/10.1051/0004-6361/201936220 AAEJAF 0004-6361 Google Scholar

31. 

L. Otal, “The optical system and the astronomical potential of A-MKID, a new camera using microwave kinetic inductance detector technology,” (2015). Google Scholar

32. 

L. Ferrari et al., “MKID large format array testbed,” IEEE Trans. Terahertz Sci. Technol., 8 (6), 572 –580 (2018). https://doi.org/10.1109/TTHZ.2018.2871365 Google Scholar

33. 

G. Wilson et al., “The TolTEC project: a millimeter wavelength imaging polarimeter (Conference Presentation),” Proc. SPIE, 10708 107080I (2018). https://doi.org/10.1117/12.2313347 PSISDG 0277-786X Google Scholar

34. 

J. E. Austermann et al., “Millimeter-wave polarimeters using kinetic inductance detectors for TolTEC and beyond,” J. Low Temp. Phys., 193 120 –127 (2018). https://doi.org/10.1007/s10909-018-1949-5 JLTPAC 0022-2291 Google Scholar

37. 

W. Holland et al., “SCUBA-2: the 10,000 pixel bolometer camera on the James Clerk Maxwell Telescope,” Mon. Not. R. Astron. Soc., 430 (4), 2513 –2533 (2013). https://doi.org/10.1093/mnras/sts612 MNRAA4 0035-8711 Google Scholar

38. 

E. Kovetz et al., “Astrophysics and cosmology with line-intensity mapping,” (2019). Google Scholar

39. 

E. Shirokoff et al., “Design and performance of superspec: an on-chip, kid-based, mm-wavelength spectrometer,” J. Low Temp. Phys., 176 (5-6), 657 –662 (2014). https://doi.org/10.1007/s10909-014-1122-8 JLTPAC 0022-2291 Google Scholar

40. 

T. Takekoshi et al., “DESHIMA on ASTE: on-sky responsivity calibration of the integrated superconducting spectrometer,” J. Low Temp. Phys., 199 231 –239 (2020). https://doi.org/10.1007/s10909-020-02338-0 JLTPAC 0022-2291 Google Scholar

41. 

Concerto Collaborationet al., “A wide field-of-view low-resolution spectrometer at APEX: instrument design and scientific forecast,” Astron. Astrophys., 642 A60 (2020). https://doi.org/10.1051/0004-6361/202038456 AAEJAF 0004-6361 Google Scholar

42. 

J. Vieira et al., “The terahertz intensity mapper (TIM): an imaging spectrometer for galaxy evolution studies at high-redshift,” 208 –215 Gothenburg, Sweden (2019). Google Scholar

43. 

P. Ade et al., “The experiment for cryogenic large-aperture intensity mapping (EXCLAIM),” J. Low Temp. Phys., 199 1027 –1037 (2020). https://doi.org/10.1007/s10909-019-02320-5 JLTPAC 0022-2291 Google Scholar

44. 

S. Masi et al., “Kinetic inductance detectors for the OLIMPO experiment: in-flight operation and performance,” J. Cosmol. Astropart. Phys., 2019 003 (2019). https://doi.org/10.1088/1475-7516/2019/07/003 Google Scholar

45. 

N. Lourie et al., “Preflight characterization of the BLAST-TNG receiver and detector arrays,” Proc. SPIE, 10708 107080L (2018). https://doi.org/10.1117/12.2314396 PSISDG 0277-786X Google Scholar

46. 

S. Rowe et al., “A passive terahertz video camera based on lumped element kinetic inductance detectors,” Rev. Sci. Instrum., 87 (3), 033105 (2016). https://doi.org/10.1063/1.4941661 RSINAK 0034-6748 Google Scholar

47. 

J. Sayers et al., “A millimeter-wave kinetic inductance detector camera for long-range imaging through optical obscurants,” Proc. SPIE, 11411 114110H (2020). https://doi.org/10.1117/12.2557428 PSISDG 0277-786X Google Scholar

48. 

S. Hailey-Dunsheath et al., “Development of aluminum LEKIDs for balloon-borne far-IR spectroscopy,” J. Low Temp. Phys., 193 968 –975 (2018). https://doi.org/10.1007/s10909-018-1927-y JLTPAC 0022-2291 Google Scholar

49. 

J. van Rantwijk et al., “Multiplexed readout for 1000-pixel arrays of microwave kinetic inductance detectors,” IEEE Trans. Microwave Theory Tech., 64 1876 –1883 (2016). https://doi.org/10.1109/TMTT.2016.2544303 Google Scholar

50. 

P. J. de Visser et al., “Fluctuations in the electron system of a superconductor exposed to a photon flux,” Nat. Commun., 5 3130 (2014). https://doi.org/10.1038/ncomms4130 NCAOBW 2041-1723 Google Scholar

51. 

R. M. J. Janssen et al., “High optical efficiency and photon noise limited sensitivity of microwave kinetic inductance detectors using phase readout,” Appl. Phys. Lett., 103 203503 (2013). https://doi.org/10.1063/1.4829657 APPLAB 0003-6951 Google Scholar

52. 

R. Barends et al., “Noise in NbTiN, Al, and Ta superconducting resonators on silicon and sapphire substrates,” IEEE Trans. Appl. Supercond., 19 936 –939 (2009). https://doi.org/10.1109/TASC.2009.2018086 ITASE9 1051-8223 Google Scholar

53. 

J. Gao et al., “A semiempirical model for two-level system noise in superconducting microresonators,” Appl. Phys. Lett., 92 212504 (2008). https://doi.org/10.1063/1.2937855 APPLAB 0003-6951 Google Scholar

54. 

O. Noroozian et al., “Two-level system noise reduction for microwave kinetic inductance detectors,” AIP Conf. Proc., 1185 (1), 148 –151 (2009). https://doi.org/10.1063/1.3292302 APCPCS 0094-243X Google Scholar

55. 

K. Karatsu et al., “Mitigation of cosmic ray effect on microwave kinetic inductance detector arrays,” Appl. Phys. Lett., 114 032601 (2019). https://doi.org/10.1063/1.5052419 APPLAB 0003-6951 Google Scholar

56. 

R. Barends et al., “Reduced frequency noise in superconducting resonators,” Appl. Phys. Lett., 97 033507 (2010). https://doi.org/10.1063/1.3467052 APPLAB 0003-6951 Google Scholar

57. 

R. Nie et al., “Optimization of a quasi-mesh absorber for the terahertz intensity mapper,” IEEE Trans. Terahertz Sci. Technol., 10 704 –712 (2020). https://doi.org/10.1109/TTHZ.2020.3022020 Google Scholar

58. 

S. Kaplan et al., “Quasiparticle and phonon lifetimes in superconductors,” Physical Review B, 14 (11), 4854 (1976). https://doi.org/10.1103/PhysRevB.14.4854 PRBMDO 1098-0121 Google Scholar

59. 

R. Barends et al., “Quasiparticle relaxation in optically excited high-Q superconducting resonators,” Phys. Rev. Lett., 100 257002 (2008). https://doi.org/10.1103/PhysRevLett.100.257002 PRLTAO 0031-9007 Google Scholar

60. 

R. Barends et al., “Enhancement of quasiparticle recombination in Ta and Al superconductors by implantation of magnetic and nonmagnetic atoms,” Phys. Rev. B, 79 020509 (2009). https://doi.org/10.1103/PhysRevB.79.020509 Google Scholar

61. 

P. J. de Visser et al., “Number fluctuations of sparse quasiparticles in a superconductor,” Phys. Rev. Lett., 106 167004 (2011). https://doi.org/10.1103/PhysRevLett.106.167004 PRLTAO 0031-9007 Google Scholar

62. 

H. McCarrick et al., “Horn-coupled, commercially-fabricated aluminum lumped-element kinetic inductance detectors for millimeter wavelengths,” Rev. Sci. Instrum., 85 (12), 123117 (2014). https://doi.org/10.1063/1.4903855 RSINAK 0034-6748 Google Scholar

63. 

A. Fyhrie et al., “Decay times of optical pulses for aluminum cpw kids,” J. Low Temp. Phys., 199 688 –695 (2020). https://doi.org/10.1007/s10909-020-02377-7 JLTPAC 0022-2291 Google Scholar

64. 

J. Hubmayr et al., “Photon-noise limited sensitivity in titanium nitride kinetic inductance detectors,” Appl. Phys. Lett., 106 073505 (2015). https://doi.org/10.1063/1.4913418 APPLAB 0003-6951 Google Scholar

65. 

S. Kumar et al., “Temperature dependence of the frequency and noise of superconducting coplanar waveguide resonators,” Appl. Phys. Lett., 92 123503 (2008). https://doi.org/10.1063/1.2894584 APPLAB 0003-6951 Google Scholar

66. 

L. J. Swenson et al., “Operation of a titanium nitride superconducting microresonator detector in the nonlinear regime,” J. Appl. Phys., 113 104501 (2013). https://doi.org/10.1063/1.4794808 JAPIAU 0021-8979 Google Scholar

68. 

P. J. de Visser et al., “Evidence of a nonequilibrium distribution of quasiparticles in the microwave response of a superconducting aluminum resonator,” Phys. Rev. Lett., 112 047004 (2014). https://doi.org/10.1103/PhysRevLett.112.047004 PRLTAO 0031-9007 Google Scholar

69. 

D. J. Goldie and S. Withington, “Non-equilibrium superconductivity in quantum-sensing superconducting resonators,” Supercond. Sci. Technol., 26 015004 (2013). https://doi.org/10.1088/0953-2048/26/1/015004 SUSTEF 0953-2048 Google Scholar

70. 

J. Perido et al., “Extending kids to the mid-IR for future space and suborbital observatories,” J. Low Temp. Phys., 199 696 –703 (2020). https://doi.org/10.1007/s10909-020-02364-y JLTPAC 0022-2291 Google Scholar

71. 

J. Glenn et al., “Galaxy Evolution Probe concept study report,” (2019). https://smd-prod.s3.amazonaws.com/science-red/s3fs-public/atoms/files/GEP_Study_Rpt.pdf Google Scholar

73. 

F. Defrance et al., “A 1.6:1 bandwidth two-layer antireflection structure for silicon matched to the 190-310 GHz atmospheric window,” Appl. Opt., 57 (18), 5196 –5209 (2018). https://doi.org/10.1364/AO.57.005196 APOPAI 0003-6935 Google Scholar

74. 

P. Savander, “Microlens arrays etched in glass and silicon,” Opt. Lasers Eng., 20 97 –107 (1994). https://doi.org/10.1016/0143-8166(94)90020-5 Google Scholar

75. 

K. P. Larsen, J. T. Ravnkilde and O. Hansen, “Investigations of the isotropic etch of an ICP source for silicon microlens mold fabrication,” J. Micromech. Microeng., 15 873 –882 (2005). https://doi.org/10.1088/0960-1317/15/4/028 JMMIEZ 0960-1317 Google Scholar

76. 

C.-F. Chen et al., “Silicon microlens structures fabricated by scanning-probe gray-scale oxidation,” Opt. Lett., 30 652 –654 (2005). https://doi.org/10.1364/OL.30.000652 OPLEDP 0146-9592 Google Scholar

77. 

P. N. A. Belmonte et al., “Microfabrication and characterization of single-mask silicon microlens arrays for the IR spectra,” Proc. SPIE, 9130 91300D (2014). https://doi.org/10.1117/12.2052527 PSISDG 0277-786X Google Scholar

78. 

Z. Deng et al., “Fabrication of large-area concave microlens array on silicon by femtosecond laser micromachining,” Opt. Lett., 40 1928 –1931 (2015). https://doi.org/10.1364/OL.40.001928 OPLEDP 0146-9592 Google Scholar

79. 

X. Meng et al., “Simple fabrication of closed-packed IR microlens arrays on silicon by femtosecond laser wet etching,” Appl. Phys. A, 121 157 –162 (2015). https://doi.org/10.1007/s00339-015-9402-y Google Scholar

80. 

H. Zuo et al., “CMOS compatible fabrication of micro, nano convex silicon lens arrays by conformal chemical vapor deposition,” Opt. Express, 25 3069 –3076 (2017). https://doi.org/10.1364/OE.25.003069 OPEXFF 1094-4087 Google Scholar

81. 

J. F. Gonzàlez et al., “Infrared antennas coupled to lithographic Fresnel zone plate lenses,” Appl. Opt., 43 6067 –6073 (2004). https://doi.org/10.1364/AO.43.006067 APOPAI 0003-6935 Google Scholar

82. 

X. Liu et al., “Superconducting micro-resonator arrays with ideal frequency spacing,” Appl. Phys. Lett., 111 (25), 252601 (2017). https://doi.org/10.1063/1.5016190 APPLAB 0003-6951 Google Scholar

83. 

S. Shu et al., “Increased multiplexing of superconducting microresonator arrays by post-characterization adaptation of the on-chip capacitors,” Appl. Phys. Lett., 113 082603 (2018). https://doi.org/10.1063/1.5040968 APPLAB 0003-6951 Google Scholar

84. 

Planck Collaboration, “Planck 2013 results. X. HFI energetic particle effects: characterization, removal, and simulation,” Astron. Astrophys., 571 A10 (2014). https://doi.org/10.1051/0004-6361/201321577 AAEJAF 0004-6361 Google Scholar

85. 

A. Monfardini et al., “Lumped element kinetic inductance detectors for space applications,” Proc. SPIE, 9914 99140N (2016). https://doi.org/10.1117/12.2231758 PSISDG 0277-786X Google Scholar

Biographies of the authors are not available.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Steven Hailey-Dunsheath, Reinier M. J. Janssen, Jason Glenn, Charles M. Bradford, Joanna Perido, Joseph Redford, and Jonas Zmuidzinas "Kinetic inductance detectors for the Origins Space Telescope," Journal of Astronomical Telescopes, Instruments, and Systems 7(1), 011015 (18 March 2021). https://doi.org/10.1117/1.JATIS.7.1.011015
Received: 15 June 2020; Accepted: 27 January 2021; Published: 18 March 2021
Lens.org Logo
CITATIONS
Cited by 11 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Sensors

Inductance

Resonators

Space telescopes

Aluminum

Capacitors

Absorption

Back to Top