Analytical instruments or scientific instruments are indispensable for scientific research and industry. The analytical instruments require a detector that converts physical quantities to be measured (measurands) to electric signals. This Tutorial describes the basics of quantum and thermal detectors, the operation principles of superconductor detectors, and the ultimate performance of state-of-art analytical instruments with superconductivity. We still face fundamental issues, such as the classical Fano factor, the relation between energy gap and mean carrier creation energy, quasiparticle dynamics, and the intermediate state in the middle of superconducting transition; and engineering issues, such as the small sensitive area and the spatially nonuniform response. Nevertheless, enormous efforts have matured superconductor detectors, which enables us to solve the inherent problems of conventional analytical instruments. As an example of the analytical results, we describe x-ray spectroscopy and mass spectrometry at our institute by using three detector types: superconductor tunnel junction, transition edge sensor, and superconductor strip. Microwave kinetic inductance and metallic magnetic calorimetric types are also described. The analytical results may contribute to a wide range of fields, such as dentistry, molecular biology, energy-saving society, planetary science, and prebiotic organic molecules in space.
Contents
ABSTRACT | 1 |
I. INTRODUCTION | 2 |
A. Analytical instruments and detectors | 2 |
B. Superconductor materials for detectors | 3 |
II. BASICS OF PARTICLE DETECTION | 5 |
A. General | 5 |
B. Energy resolution of quantum detectors | 5 |
1. The statistical model and carrier number fluctuation | 5 |
2. The Fano factor | 6 |
3. Guidance from the statistical model for quantum detection | 6 |
C. Energy resolution of thermal detectors | 7 |
1. The statistical model with canonical ensemble | 7 |
2. Guidance from the statistical model for thermal detection | 8 |
III. SUPERCONDUCTOR PARTICLE DETETORS | 8 |
A. General | 8 |
B. Superconductor quantum detectors | 8 |
1. A historical overview | 8 |
2. Quasiparticle excitation | 9 |
3. Superconductor tunnel junction (STJ) type | 11 |
4. Microwave kinetic inductance (MKI) type | 24 |
C. Superconductor thermal detectors | 25 |
1. A historical overview | 25 |
2. Transition edge sensor (TES) type | 27 |
3. Metallic magnetic calorimetric (MMC) type | 30 |
4. Superconductor strip (SS) type | 31 |
D. Performance comparison | 37 |
1. Performance figures | 37 |
2. Energy resolution | 37 |
3. Fall time and count rate | 38 |
4. Energy coverage | 38 |
5. Spatial resolution and arraying | 38 |
6. Cryogenics and other aspects | 39 |
IV. ADVANCES FOR PRACTICAL USE | 39 |
A. General | 39 |
B. Superconductor tunnel junction detector (STJD) | 39 |
1. Fabrication of STJ array | 39 |
2. Cryogenics and readout electronics | 41 |
C. Superconductor strip particle detector (SSPD) | 42 |
1. Fabrication of SS array | 42 |
2. Operation in parallel configuration | 43 |
V. ANALYTICAL APPLICATIONS | 45 |
A. General | 45 |
B. X-ray spectroscopy (XRS) | 46 |
1. Scanning electron microscopy (SEM) | 46 |
2. X-ray absorption spectroscopy (XAS) | 49 |
C. Mass spectrometry (MS) | 55 |
1. MS instrumentation | 55 |
2. True MS eliminating m/z overlap | 56 |
3. Neutral particle MS eliminating neutral loss | 59 |
VI. SUMMARY | 62 |
ACKNOWLEDGEMENTS | 63 |
Data Availability Statement | 64 |
APPENDIX A. Definition of energy resolution and probability distribution functions | 64 |
APPENDIX B. IUPAC notation of characteristic X-ray lines and interaction with detector materials | 64 |
APPENDIX C. Terminology | 65 |
1. International standards | 65 |
2. Superconductor detector, superconducting detector, or superconductive detector | 66 |
3. Nanostrip or nanowire | 66 |
References | 66 |
ABSTRACT | 1 |
I. INTRODUCTION | 2 |
A. Analytical instruments and detectors | 2 |
B. Superconductor materials for detectors | 3 |
II. BASICS OF PARTICLE DETECTION | 5 |
A. General | 5 |
B. Energy resolution of quantum detectors | 5 |
1. The statistical model and carrier number fluctuation | 5 |
2. The Fano factor | 6 |
3. Guidance from the statistical model for quantum detection | 6 |
C. Energy resolution of thermal detectors | 7 |
1. The statistical model with canonical ensemble | 7 |
2. Guidance from the statistical model for thermal detection | 8 |
III. SUPERCONDUCTOR PARTICLE DETETORS | 8 |
A. General | 8 |
B. Superconductor quantum detectors | 8 |
1. A historical overview | 8 |
2. Quasiparticle excitation | 9 |
3. Superconductor tunnel junction (STJ) type | 11 |
4. Microwave kinetic inductance (MKI) type | 24 |
C. Superconductor thermal detectors | 25 |
1. A historical overview | 25 |
2. Transition edge sensor (TES) type | 27 |
3. Metallic magnetic calorimetric (MMC) type | 30 |
4. Superconductor strip (SS) type | 31 |
D. Performance comparison | 37 |
1. Performance figures | 37 |
2. Energy resolution | 37 |
3. Fall time and count rate | 38 |
4. Energy coverage | 38 |
5. Spatial resolution and arraying | 38 |
6. Cryogenics and other aspects | 39 |
IV. ADVANCES FOR PRACTICAL USE | 39 |
A. General | 39 |
B. Superconductor tunnel junction detector (STJD) | 39 |
1. Fabrication of STJ array | 39 |
2. Cryogenics and readout electronics | 41 |
C. Superconductor strip particle detector (SSPD) | 42 |
1. Fabrication of SS array | 42 |
2. Operation in parallel configuration | 43 |
V. ANALYTICAL APPLICATIONS | 45 |
A. General | 45 |
B. X-ray spectroscopy (XRS) | 46 |
1. Scanning electron microscopy (SEM) | 46 |
2. X-ray absorption spectroscopy (XAS) | 49 |
C. Mass spectrometry (MS) | 55 |
1. MS instrumentation | 55 |
2. True MS eliminating m/z overlap | 56 |
3. Neutral particle MS eliminating neutral loss | 59 |
VI. SUMMARY | 62 |
ACKNOWLEDGEMENTS | 63 |
Data Availability Statement | 64 |
APPENDIX A. Definition of energy resolution and probability distribution functions | 64 |
APPENDIX B. IUPAC notation of characteristic X-ray lines and interaction with detector materials | 64 |
APPENDIX C. Terminology | 65 |
1. International standards | 65 |
2. Superconductor detector, superconducting detector, or superconductive detector | 66 |
3. Nanostrip or nanowire | 66 |
References | 66 |
I. INTRODUCTION
The superconductor detector technology utilizes superconductivity of at least one of three detector components: an absorber that interacts with particles, a device for electric signal generation, and a readout circuit. In conventional analytical instruments, the signal generation device typically consists of a semiconductor device. Replacement of the semiconductor detector with the superconductor detector leads to ultimate instrumental performance, overcoming the inherent limits of conventional analytical instruments. Although subsequent digital signal processing is conventionally performed by semiconductor digital circuits, the latest superconductor digital technology can also replace semiconductor circuits in future. The state-of-art superconductor electronics for both analog and digital circuits appears in Ref. 1. The superconductor sensors and detectors accounted for a large part of the superconductivity centennial conference—EUCAS-ISEC-ICMC in 2011.2
This article begins with the general aspect of analytical instrumentation and the role of detectors. In the same manner as the conventional particle detectors, the superconductor detectors are principally categorized into two detection schemes: “quantum detection,” in which electric signals are directly produced by electronic excitation, and “thermal detection,” in which temperature rise due to radiation is measured with a thermometer. It should be noted that this categorization is based on how the output signal is created. Both superconductor quantum and thermal detection schemes utilize superconductivity; therefore, all superconductor detector types can be also categorized as quantum detection.
After mentioning the components of analytical instruments, the general operation principles of quantum and thermal detection are explained before deepening our understanding of the superconductor detector operation. Next, we describe the theoretical and experimental efforts to achieve expected high performance of superconductor detectors, followed by engineering advances for equipping analytical instruments with superconductor detectors. Finally, unconventional analytical data obtained mostly by our group at the National Institute of Advanced Industrial Science and Technology (AIST), part of which was formerly Electrotechnical Laboratory (ETL), are highlighted: x-ray spectroscopy (XRS) and mass spectrometry (MS). We have developed superconductor tunnel junction detectors (STJDs), transition edge sensor detectors (TESDs), and superconductor strip particle detectors (SSPDs) for x-ray photons and particles like ions and neutral molecules. They provide novel analytical data concerning dentistry, power semiconductors, atmosphere escape from planets, and prebiotic organic molecules in interstellar space as an origin of life. This article also covers other major detector types: microwave kinetic inductance detectors (MKIDs) and magnetic metallic calorimetric detectors (MMCDs).
Superconductor detectors were initially developed for detecting neutrinos and then applied to other particles. Topical reviews summarized the status in 1996.3,4 A comprehensive review for thermal detectors appears in Ref. 5. This article is arranged such that one can trace a difference between quantum and thermal detectors, merits and demerits, and similarity and dissimilarity between semiconductor and superconductor detectors. For nonexperts and students who newly join in this research field, Appendix A describes the definition of energy resolution and related probability distribution functions, Appendix B explains characteristic x-ray emission lines and excitation inside x-ray detector absorber materials, and Appendix C deals with terminology and nomenclature for superconductor detectors. We use the standard abbreviations for the superconductor detector types, considering the recommendation of the international electrotechnical commission (IEC) standards6 except the term of “superconducting detector” that will be changed to “superconductor detector” (see Appendix C 2). The experts can focus on a specific chapter or sections of interest by skipping general sections.
A. Analytical instruments and detectors
Analytical instruments are typically divided into three components: a probe beam to excite analytes, a spectrometer to separate particles, such as x-ray photons emitted from the analytes, and a detector to produce spectral data. The spectra often consist of a number of particles for the y axis as a function of particle energies for the x axis. We emphasize that detectors often determine the quality of spectra and, thus, the performance of analytical instruments. In contrast to the precise detection of rare events such as dark matter,7 a higher count rate is vastly preferable for high throughput analysis or imaging analysis.
Figure 1(a) shows one example of the typical analytical instruments: electron probe microanalyzer (EPMA), which is equipped with an electron gun with an optics device focusing electrons onto a sample surface and x-ray spectrometers for analyzing characteristic x rays from atoms. Elemental mapping is realized by scanning the electron beam while measuring characteristic x rays. The x-ray spectrometer for EPMA consists of a combination of an analyzing diffraction crystal or grating and an x-ray photon counter, which is called wavelength dispersive spectroscopy (WDS). WDS spectrometers have a typical energy resolution (ΔE) of 10 eV, which is sufficient to separate different characteristic x-ray lines. However, the WDS spectrometers require collimators or slits to fix the diffraction angle, leading to a small solid angle. Therefore, to compensate the small solid angle, a high electron beam current is required, which tends to create sample damage. Another disadvantage of EPMA is that the upper limit of the number of WDS spectrometers is about five because of a geometrical limit. It results in the restriction on the number of simultaneously measurable elements.
Schematic drawing for typical analytical instruments: (a) electron probe micro analyzer (EPMA) and (b) scanning electron microscope (SEM). EPMA consists of an electron probe that excites characteristic x-ray lines, a spectrometer for elemental analysis, and an x-ray photon counter. This is called wave dispersive spectroscopy (WDS). The WDS spectrometer can be replaced by an energy-sensitive detector in (b), which is called energy dispersive spectroscopy (EDS). The semiconductor EDS detectors normally have a ∼10 times worse energy resolution than WDS. On the other hand, superconductor EDS detectors have a high-energy resolution comparable to WDS.
Schematic drawing for typical analytical instruments: (a) electron probe micro analyzer (EPMA) and (b) scanning electron microscope (SEM). EPMA consists of an electron probe that excites characteristic x-ray lines, a spectrometer for elemental analysis, and an x-ray photon counter. This is called wave dispersive spectroscopy (WDS). The WDS spectrometer can be replaced by an energy-sensitive detector in (b), which is called energy dispersive spectroscopy (EDS). The semiconductor EDS detectors normally have a ∼10 times worse energy resolution than WDS. On the other hand, superconductor EDS detectors have a high-energy resolution comparable to WDS.
Another example of similar analytical instruments is the popular scanning electron microscopy (SEM) in Fig. 1(b). SEM tolerates a low electron beam current compared with EPMA, which is effective to avoid radiation damage. X-ray spectra are acquired by another spectroscopic type called energy dispersive spectroscopy (EDS). X-ray detectors for EDS enable energy measurement of individual x-ray photons without mechanical or electrical scanning. The EDS detector can be placed near the sample surface and accept a large portion of x rays emitted at 4π steradian from the sample, realizing a significantly larger solid angle than WDS. Furthermore, simultaneous analysis of multiple elements more than five can be performed. However, the latest semiconductor EDS detectors have an energy resolution of ∼100 eV, which is insufficient for separating important characteristic x-ray lines. Natural linewidths of characteristic x rays from elements embedded in a matrix are typically 10–20 eV, broadened by chemical bonds or electronic band structure. Superconductor detectors enable to combine a WDS-grade high energy-resolution and an EDS-grade high detection efficiency. This is one of the motivations to employ superconductivity for innovative analytical instruments.
Before describing the details, let us glance at a performance difference between semiconductor and superconductor detectors in Fig. 2. We analyzed a hen eggshell and a porcine dentin, which mainly consist of calcite (CaCO3) and hydroxyapatite [Ca5(PO4)3(OH)] reinforced by collagen fibers, respectively. These biological composites have similar compositions for major elements, such as Ca, P, Mg, Na, O, N, and C. There is a small difference in trace mineral elements. Figure 2 compares the x-ray spectra acquired by a silicon drift detector (SDD)8 and an STJD. The x-ray spectra were acquired for 10 min with the SDD and STJD simultaneously. The STJD spectra for the eggshell and the dentin show fine features due to Ca, Cr, and F, drawing a distinction between the two samples. On the other hand, the SDD spectrum exhibits only broad C, N, and O peaks with a subtle sign of F. Figure 2 clearly demonstrates the advantage of the STJD over the SDD in this x-ray energy range.
Advantage of superconductor detectors over semiconductor detectors: (a) secondary electron image of the outer surface of the eggshell, (b) secondary electron image of the cross section of the dentin, and (c) comparison of the x-ray spectra measured with an electron beam energy of 5 keV at 3.5 nA for 10 min. The scanning electron microscope (SEM: JEOL JSM-7200F) is equipped with the STJD and the SDD. The pores appear on the eggshell surface in (a). The one dentinal tubule runs vertically in (b). The SDD has a sensitive area of 25 mm2, whereas the STJD has the 100 pixels of 100-μm square STJs assisted by a polycapillary x-ray lens. Assignment to the characteristic x-ray lines is based on the IUPAC notation with the conventional Siegbahn notation in parentheses (see Appendix B). The details of the dentin are described in Sect. V B 1.
Advantage of superconductor detectors over semiconductor detectors: (a) secondary electron image of the outer surface of the eggshell, (b) secondary electron image of the cross section of the dentin, and (c) comparison of the x-ray spectra measured with an electron beam energy of 5 keV at 3.5 nA for 10 min. The scanning electron microscope (SEM: JEOL JSM-7200F) is equipped with the STJD and the SDD. The pores appear on the eggshell surface in (a). The one dentinal tubule runs vertically in (b). The SDD has a sensitive area of 25 mm2, whereas the STJD has the 100 pixels of 100-μm square STJs assisted by a polycapillary x-ray lens. Assignment to the characteristic x-ray lines is based on the IUPAC notation with the conventional Siegbahn notation in parentheses (see Appendix B). The details of the dentin are described in Sect. V B 1.
B. Superconductor materials for detectors
A research trend of superconductivity is normally searching higher critical temperatures even near or above room temperature. For superconductor particle detectors, on the other hand, low temperature superconductors (LTSs) are normally preferable to achieve a better energy resolution; for example, Nb and Al. In addition, superconductor/normal metal bilayers, such as Nb/Al, Mo/Cu, Ti/Au, are utilized for controlling Tc with the proximity effect.9 Most physical aspects of detector operation using LTSs is understandable based on the standard Bardeen–Cooper–Schrieffer (BCS) theory,10 which postulates a single isotropic superconducting energy gap or s-wave symmetry.
In quantum detection, a superconducting energy gap is utilized as the scale for measuring a particle kinetic energy through the number of quasiparticles created by breaking Cooper pairs. Operating temperatures shall be well below Tc to suppress thermally excited quasiparticles. In thermal detection, superconductor films are held in the middle of superconducting transition or biased just below Ic. Superconductivity requires refrigerators, such as adiabatic demagnetization and 3He–4He dilution for a temperature range below 0.1 K and 3He sorption refrigerators at ∼0.3 K, which are significantly larger than liquid nitrogen cryostats or Peltier coolers for semiconductor detectors.
A question may arise for the practicability of high temperature superconductors (HTSs), which are advantageous to the cryogenic requirement. In 1986, immediately after the first workshop on low temperature detectors (LTD-1), an HTS system of LaBaCuO was discovered.11 The possibility of using HTSs for detectors was already discussed in 1987.12 Since then, many HTSs have been discovered. SQUIDs with HTS Josephson junctions (JJs) were fabricated soon after the HTS discovery.13,14 However, the detection scheme relying on quasiparticle excitation is inapplicable to HTSs, because most HTSs have d-wave symmetry that has a zero energy gap along some crystallographic directions. HTS devices face another difficulty of thin-film growth and microfabrication. Tc values easily change with oxygen content in the cuprate HTSs, such as YBa2Cu3Ox (YBCO), depending on thin-film preparation conditions.15 Moreover, its multi-elemental oxide nature prohibits a high-degree of uniformity. In addition to the material difficulties, HTS tunnel junctions or superconductor–insulator–superconductor (SIS) junctions with a similar quality to LTSs have never been realized because of an extremely short coherence length of less than 1 nm. In most cases, HTS JJs operate in the mode explained by the resistively shunted junction (RSJ) model.16
BiSrCaCuO (BSCCO) crystals have intrinsic JJs along the c axis, forming a stuck of SIS alternation.17 The intrinsic JJs were utilized to generate and detect THz radiation.18,19 THz emission was also observed from YBCO.20 In addition to the THz range, we had an idea for utilizing the intrinsic Josephson effect in BSCCO as a tunnel junction x-ray detector. However, it turned out that BSCCO was unrealistic for detectors because of the d-wave symmetry, unavoidable staking faults of different phases of Bi2Sr2CuOx (2201), Bi2Sr2CaCu2Ox (2212), Bi2Sr2Ca2Cu3Ox (2223), and Bi2Sr2Ca3Cu4Ox (2234),21 and long-range modulation superstructures.22 As a result, HTS-STJDs for single particles have never been realized up to now.
A successful usage of HTSs is seen in the SS type, which is essentially a particle counting device without measuring particle energies. The thermal noise is tolerable in SSPDs and, thus, an operating temperature of Tc/2 or higher is possible. One successful HTS-SSPD consists of MgB2, in which the occurrence of superconductivity is explained within the s-wave framework with multiple energy gaps.23,24 A cuprate superconductor LaSrCuO with the d-wave symmetry is close to single photon detection.25 Several groups have been trying to employ YBCO for single-particle detection,26–29 which may be operatable at 77 K. A preliminary result on single photon detection with YBCO is reported on dark count experiment.30 However, single photon detection with YBCO remains to be realized in future. Most recently, an SSPD exfoliated from a bulk BSCCO single crystal succeeded in single photon detection at temperatures up to 25 K.31
The properties of the above-mentioned typical superconductors are summarized in Table I. Important parameters are listed for detector development. Among them, the Pearl length 32 that is important for SSPD operation may be unfamiliar to some readers. It is the penetration depth for a magnetic field perpendicular to film surfaces.
Important parameters of typical superconductors: Tc is the critical temperature, Δg is the single particle energy gap, ξ is the coherence length, λ is the magnetic penetration depth, τe–p is the electron–phonon scattering time, and Λ is the Pearl length. The τe–p values are estimated at ∼0.5Tc except when it is specified exactly: τe–p has t−n dependence in the dirty-limit, where n is an integer of 2 or 3. Other tabulated parameters are evaluated at 0 K. The Δg, ξ, and λ values within the a–b plane (left) and along the c axis (right) are listed for the anisotropy of MgB2 and YBa2Cu3Ox (YBCO). Since they tend to grow preferentially as the c axis is perpendicular to a substrate surface, λc is applicable to SSPD. The Pearl length is the penetration depth of a thin film with a thickness of tf against a magnetic field perpendicular to film surfaces. The Λ values were calculated for films with a thickness of ξ. For MgB2 and YBCO, λa–b and ξc values are adopted for calculating Λ.
. | Nb . | Al . | NbN . | MgB2 . | YBCO . |
---|---|---|---|---|---|
Tc (K) | 9.23a | 1.2 | 16.15b | 39.2c | 89d |
Δg(0)/e (meV) | 1.52a | 0.17 | 2.46b | 2, 10c | 20e |
ξ(0) (nm) | 23f | 147f | 5b | 39, 35g | 2.5, 0.8h |
λ(0) (nm) | 47i | 52j | 194k | 107, 120g | 92, 635d |
τe-p (ps) | 330l | 395 000m | 88n | 2m | 1.1o |
Λ(0) (μm) | 0.19 | 0.037 | 15 | 0.65 | 21 |
. | Nb . | Al . | NbN . | MgB2 . | YBCO . |
---|---|---|---|---|---|
Tc (K) | 9.23a | 1.2 | 16.15b | 39.2c | 89d |
Δg(0)/e (meV) | 1.52a | 0.17 | 2.46b | 2, 10c | 20e |
ξ(0) (nm) | 23f | 147f | 5b | 39, 35g | 2.5, 0.8h |
λ(0) (nm) | 47i | 52j | 194k | 107, 120g | 92, 635d |
τe-p (ps) | 330l | 395 000m | 88n | 2m | 1.1o |
Λ(0) (μm) | 0.19 | 0.037 | 15 | 0.65 | 21 |
Our data: Tc is the temperature in the middle of superconducting transition of a sputtered polycrystalline Nb film with a residual resistance ratio (RRR) of 4.4, and Δg was estimated by extrapolating Al thickness dependence of 2Δg values on I–U curves of Nb/Al tunnel junctions to zero Al thickness.
Ref. 38.
Ref. 41.
YBa2Cu3O6.995 single crystal: λab(0) = (λa + λb)/2 = 92 nm and λc(0) = 635 nm.43 For epitaxially grown thin films with defects, λab(0) is typically in a range between 180 and 200 nm.44,45
Ref. 46.
Polycrystalline film.33
Ref. 42.
Ref. 47.
Polycrystalline bulk.34
Ref. 36.
Open resonator at 82 GHz.39
30-nm-thick polycrystalline Nb film.35
Ref. 37.
15-nm-thick polycrystalline film at 8 K.40
Ref. 48.
II. BASICS OF PARTICLE DETECTION
A. General
This chapter begins with the introduction to the detection principles on conventional quantum and thermal detectors, which helps us to understand why superconductivity plays a crucial role in achieving an ultimate energy resolution. The term energy resolution (ΔE) corresponds to the smallest detectable difference between two adjacent peaks for particles with different kinetic energies.49 The ΔE values are often expressed by full width at half maximum (FWHM) of a peak for mono-energetic particles (E0) (see Appendix A for the details). A total absorption peak for mono-energetic x-ray photons is a good measure of the energy resolution. High resolution corresponds to a small ΔE value, which should not be confused with resolving power that is expressed by E0/ΔE.
B. Energy resolution of quantum detectors
1. The statistical model and carrier number fluctuation
The statistical physics for semiconductors provides the primary basis for superconductor quantum detectors. In semiconductor detectors, particle incidence leads to electron excitation from the valence band to the conduction band and holes are left in the valence band, keeping charge neutrality or conserving the number of electrons. Those charge carries are collected by an electric field applied to the semiconductor, as illustrated in Fig. 3. Almost all charge carriers are collected in the modern Si detectors. The integration of the current pulse flowing RL is equal to the number of the electron–hole (e–h) pairs, as well established in semiconductor detector physics. In other words, the voltage source applies a constant voltage to the detector only; the current injected to the detector is negligible even when an x ray is absorbed. In this condition, the circulating electron current through RL annihilate the holes; therefore, the output charge number is half the total number of the quasiparticles created in the depletion layer. This is also important to consider the output charge number in superconductor detectors, which will be discussed in Secs. III B 2 and III B 3 b.
Schematic illustration of semiconductor quantum detectors and a readout circuit for the introduction to superconductor detectors. The p-n junction or surface barrier device is biased inversely to make a depletion layer without free electrons and holes. It is a case for x-ray photon incidence that generates an energetic photoelectron or an Auger electron, of which collision cascade creates e–h pairs in the depletion layer with a resistance of Rd. Those charge carriers move toward opposite directions along the electric field, producing an output current pulse in a condition of Rd ≫ Rb ≫ RL. A charge-sensitive preamplifier integrates the current pulse, of which the output charge number is equal to the number of the e–h pairs.
Schematic illustration of semiconductor quantum detectors and a readout circuit for the introduction to superconductor detectors. The p-n junction or surface barrier device is biased inversely to make a depletion layer without free electrons and holes. It is a case for x-ray photon incidence that generates an energetic photoelectron or an Auger electron, of which collision cascade creates e–h pairs in the depletion layer with a resistance of Rd. Those charge carriers move toward opposite directions along the electric field, producing an output current pulse in a condition of Rd ≫ Rb ≫ RL. A charge-sensitive preamplifier integrates the current pulse, of which the output charge number is equal to the number of the e–h pairs.
2. The Fano factor
The Fano factor was devised to bridge a gap in variance between an experimental energy resolution of gaseous ionization detectors and Poisson distribution that should govern radiation counting.51 The experimental resolution of the gaseous detectors was significantly better than that of the Poisson distribution. The physical meaning of the Fano factor is still an open question; however, it is obvious that F is the variance ratio of an exact distribution function to the corresponding Poisson distribution having the same . The Poisson statistics corresponds to the limit of n → ∞ and p → 0 in the binomial distribution, keeping a constant value. The Fano factor is certainly a valuable indicator for a deviation from the Poisson distribution; a large deviation means a better energy resolution.
Instead of the analytical description, the Fano factor can be experimentally defined.53 To be more practical, the ɛ value can be obtained by dividing E0 by the number of detected charge carriers, and ΔE is substituted by the FWHM value of the observed total absorption peak. An empirical F value is obtained from Eq. (4). In this empirical definition, any combination of F and ɛ is possible without any constraints. This is a convenient way; however, this has no physical meaning. There must be a constraint condition between F and ɛ according to the definition of F and ɛ. At present, Monte Carlo simulations provide a reasonable measure of the statistical energy resolution limit.
The Fano factor is still an active research area even these days. Many models for ionization statistics have been proposed to explain F values that are significantly smaller than 1.0: statistical partitioning models including “crazy carpentry model”54 in 1965, “phenomenological models of electron cascades”55 in 2006, and “bathtub-whiskey-glass model”56 in 2008. A recent theoretical study for a low energy region, in which only a few charge carriers are created, adopted a method with freedom on a parameter space. The adopted probability function was the Conway–Maxwell–Poisson distribution, which was a member of the weighted Poisson distribution family covering both cases of F < 1 and F > 1.57 More discussion on the Fano factor appears in that paper and references cited therein.
3. Guidance from the statistical model for quantum detection
Superconductor detectors should follow Eqs. (2) and (4) in principle, since they are common to all quantum detectors having an energy gap. Although there are still uncertainties of F and ɛ, they signify that all quantum detectors should follow a dependence of regardless of semiconductors and superconductors. The lower the particle energy is, the better the energy resolution is.
Based on the statistical definition of F, very small F values mean that carrier creation trials along a particle trajectory are strongly dependent. In other words, in F ≪ 1.0, charge carriers are created efficiently, so that ɛ/Eg should approach 1.0. If all the particle energy is consumed to create e–h pairs and all the carriers are collected, there is no chance of statistical fluctuation: ɛ = Eg and F = ΔE = 0. In contrast, when ɛ/Eg is large, statistical fluctuation of the carrier number should be also large, approaching the Poisson statistics. Therefore, the empirical determination without any constraints is fraught of incorrect results due to experimental or simulation shortcomings. Another possible fluctuation or loss processes can be attributed to the binding energies of EK, EL, EM, etc. A keV photoelectron or an Auger electron cannot carry the whole x-ray photon energy because of the biding energies of electron shells, that is, E0-EK, E0-EL, E0-EM, etc. When we subtract the biding energies from E0, the relation of ɛ and F becomes more reasonable.50
To estimate the ɛ–2Δg relation analytically, we can follow the Klein treatment58 for semiconductors from the point of view of the density of the states (DOS). It calculates an average kinetic energy of excited electrons and holes in semiconductors within a free-particle approximation. Although the DOS profile of superconductors is completely different from that of semiconductors, it is worthwhile to apply the Klein treatment to superconductors. It provides a possible physical basis on the ɛ–2Δg relationship and is not inconsistent with experimental and Monte Carlo simulation data.50 Although we may need to formulate a new model for the ɛ–2Δg or ɛ–Eg relation in superconductors as well as semiconductors, Owens and Peacock plotted energy resolutions expected for different superconductors as a function of 2Δg, extrapolating the data of semiconductors in a few eV toward a lower energy 2Δg region in a few meV.59 The assumed Fano factors were 0.14 for semiconductors and 0.22 for superconductors. The expected ΔE values at 5.9 keV range from 9 eV of MgB2 to 0.5 eV of Hf.
One concern unique to superconductors is the possible formation of normal cores or reduced Δg cores in several keV x-ray absorption60,61 because of numerous quasiparticles in contrast to semiconductors that have a constant Eg value independent of e–h density. When it happens, quasiparticle loss occurs, resulting in a low charge output or a low energy resolution.
C. Energy resolution of thermal detectors
1. The statistical model with canonical ensemble
A schematic diagram of the thermal detection is shown in Fig. 4. The essence of cryogenic thermal detectors is that cryogenic low temperature T enables a small thermal energy fluctuation of an absorber with a small heat capacity C. In principle, there is no lower detection limit for photon energies or wavelengths because of no energy gap.
Schematic illustration of thermal detectors for the introduction to superconductor detectors. The absorber with the heat capacity C coupled to a thermometer with resistance of R denoted by the black bar is connected to the heat bath of Tb through heat conductance G. Two detection schemes are illustrated: single particle (kinetic energy) and particle flux (power). Either single-particle detection or flux measurement is possible, depending on the relation between detector time response and particle event frequency.
Schematic illustration of thermal detectors for the introduction to superconductor detectors. The absorber with the heat capacity C coupled to a thermometer with resistance of R denoted by the black bar is connected to the heat bath of Tb through heat conductance G. Two detection schemes are illustrated: single particle (kinetic energy) and particle flux (power). Either single-particle detection or flux measurement is possible, depending on the relation between detector time response and particle event frequency.
Thermal system consisting of a large heat bath with temperature Tb and energy Eb, and the subsystem with temperature T, energy Es, constant volume V, and particle number N. It is assumed that the total energy of the whole system Et is constant and energy exchange E is possible between the heat bath and the subsystem under thermal equilibrium.
Thermal system consisting of a large heat bath with temperature Tb and energy Eb, and the subsystem with temperature T, energy Es, constant volume V, and particle number N. It is assumed that the total energy of the whole system Et is constant and energy exchange E is possible between the heat bath and the subsystem under thermal equilibrium.
2. Guidance from the statistical model for thermal detection
Superconductor thermal detectors should follow Eq. (6). Comparing Eqs. (4) and (6), it is apparent that ΔE of thermal detectors is independent of incoming particle energies, assuming that a temperature increase of the subsystem or absorber is negligibly small compared with the bath temperature.
Equation (6) is not a practical limit in modern thermal detectors. When the power spectrum of fluctuation is concerned, it spreads over a wide range of frequencies. Output signal pulses have a shape of exponential decay, so that it is possible to narrow a readout bandwidth. This is called optimal filtering. Accordingly, the energy fluctuation observed experimentally can be smaller than Eq. (6). Nevertheless, Eq. (6) is a primary guidance for detector design. The energy resolution becomes better or smaller when a thermal detector has a smaller C and operated at a lower T. In contrast to quantum detectors, of which energy resolution is proportional to and independent of T, thermal detectors have the energy resolution independent of E0 and proportional to T. Thermal detectors can cover a wide particle-energy range at a constant energy resolution in principle. This difference suggests that thermal detectors are advantageous for high-energy particles, whereas quantum detectors perform well for low energy particles.
III. SUPERCONDUCTOR PARTICLE DETECTORS
A. General
We can utilize a superconductor at one part or more for the three components: absorber, electric signal generation device, and readout circuit. In general, quantum detectors have a superconducting absorber, whereas thermal detectors have a superconductor thermometer. Superconductors have a smaller C value than normal metals because of the gap. Therefore, a superconductor can also be used as the absorber for thermal detectors. However, because of long-lived electronic excitations in superconductors, a normal metal absorber like a Bi/Au bilayer is a choice for modern superconductor thermal detectors.63
B. Superconductor quantum detectors
1. A historical overview
Equation (2) indicates that an absorber with a small Eg value is beneficial to obtain a better energy resolution. Are there any solid-state materials that have a uniform small energy gap across a large volume enough for realizing a practical sensitive area or volume? The answer is to use superconductors, of which energy gaps are in a range of meV that is about 1000 times smaller than that of Si. The meV gaps are kept uniform across a macroscale without being influenced by local defects. Therefore, polycrystalline superconductors can be employed in contrast to the requirement for high quality single crystals for semiconductor detectors to avoid carrier trapping.
Provided that superconductor detectors have the same p value or Fano factor as semiconductor detectors, Eq. (2) indicates that superconductor detectors enable an approximately 30-times better energy resolution than that of semiconductor detectors: . It is surprising that the 30-times better energy resolution was predicted 58 years ago in the Wood's master’s thesis under the supervision of White in 1965.64 In the 1960s, they faced practical problems of superconductor materials, fabrication, cryogenics, and readout electronics. In his master’s thesis on Al/AlOx/Pb junctions, he wrote “The most formidable problem was degradation of Al/AlOx/Pb junctions showing junction resistance increased with time after fabrication.” He also concluded “The experimental and theoretical study reported in this thesis has shown the plausibility of successfully developing a superconducting charged particle detector. There is, however, a considerable difference between a device working in theory and a device working in practice.” For the doctor’s thesis, they moved onto an Sn/SnO2/Sn junction. They successfully observed output pulses from the Sn/SnO2/Sn tunnel junction for 5.1-MeV α-particles, although no total absorption peak was observed in 1969.65,66
Fifteen years after the work by Wood and White, an Sn–SnO2–Sn junction was utilized for detecting α-particle by Kurakado and Mazaki in 1980.67 The Sn-based junction successfully acquired pulse height spectra, which was the first spectroscopic measurement with STJDs; however, no distinct total absorption peak corresponding to 5.1-MeV α-particles was observed. A Monte Carlo simulation on ɛ and F values was performed for Sn in 1982.68 The result on Sn was ɛ = 1.68Δg/e (=0.969 meV) and F = 0.195 for stimulation at 5.75 eV (=104 ⋅ Δg/e), which is an important guideline for superconductor quantum detectors. This result is consistent with succeeding simulations for other superconductors.69–71 The parameter values should be corrected to ɛ = 1.68 × 2Δg and F = 0.098 for quasiparticle pairs that are created simultaneously, as discussed in Section III B 2. In addition to α-particles, Twerenbold et al. utilized a Sn-based junction to detect 20-keV lysozyme protein ions (14 305 Da) for mass spectrometry (MS) in 1996.72,73 Hilton improved an energy resolution for biomolecules by using a normal–insulator–superconductor (NIS) junction in 1998, enabling charge-state discrimination for bovine serum albumin (BSA, 66 430 Da).74 Wenzel et al. realized detection of macromolecules, such as immunoglobulin M (IgM, 1 MDa) and von Willebrand factor (1.5 MDa), by using a 16-pixel Nb/Al-STJD in 2005.75
The energy resolution limits expected by using the Monte Carlo simulation71 on Nb and Eq. (4) are 4.3 eV for 5895 eV (55Fe x-ray source), 1.3 eV for 525 eV (O K-L3), and 1.1 eV for 392 eV (N K-L3), assuming that all the quasiparticles are collected once. They are far better than 119 eV for 5895 eV, 36 eV for 525 eV, and 31 eV for 392 eV by using a Monte Carlo simulation with ɛ = 3.70 eV and F = 0.117 for Si.76 Experimentally, we recorded an intrinsic energy resolution of 3.6 eV at 400 eV with a Nb/Al-STJD in this article. A Ta/Al-STJD of Lawrence Livermore National Laboratory (LLNL) achieved 1.5 eV at 250 eV and 3.5 eV at 1 keV.77 The STJ operation adds fluctuation due to imperfect charge collection, multiple quasiparticle tunneling, and spatial nonuniformity as discussed in Sec. III B 3.
Another striking feature of the meV gap is to extend a range of single-photon energy measurement toward the lower energy side. Nb- and Ta-based STJDs marked a remarkable performance of ΔE/E0 ≈ 10% near 1 eV in a range from ultraviolet to near-infrared photons.78–80 This was the first single-photon energy measurement in an infrared–optical range. A similar measurement was also realized by thermal detection with TESDs.81
The other quantum detector type has a superconductor strip resonator structure to sense a surface impedance change in a microwave frequency range, based on Cooper pair breaking. It was recently invented in 2003: microwave kinetic inductance detector (MKID) for a large-scale array format for imaging.82 Quasiparticle excitation due to radiation induces a Q-factor change. Each pixel has its own resonance frequency so that it is possible to read out all pixels through one readout line.
2. Quasiparticle excitation
In this section, we describe the e–h pair excitation in semiconductors first as a guidance on how to understand the quasiparticle excitation in superconductor quantum detectors. This approach also holds great significance as it addresses the crucial need for a proper understanding of the statistical treatment concerning the relationship between the number of quasiparticles in superconductor detectors and the output charge number.
There are two types of charge excitations in metals, semiconductors, insulators, and superconductors: electron-like excitation (quasielectron) and hole-like excitation (quasihole), which are collectively called quasiparticles. Energetic particles, such as photons, electrons, protons, and even atoms, create quasiparticles. Let us consider a case of x-ray photons, since all the particles end up the same excitation process as x rays except for initial interaction. The x-ray absorption can be transformed into an electron energy loss process after the emission of a keV photoelectron or an Auger electron from an atom. The highly excited electrons relax to the order of eV, which is equivalent to Eg for ∼0.1 ps, through inelastic scattering with electrons. They relax to the order of meV equivalent to Δg of superconductors in an elapsed time range of 1 ns–1 μs through electron–electron or electron–phonon scattering.83 These collision cascades excite quasiparticles. Figure 6 covers the excitation energy range below 10 eV and depicts the DOS profiles of semiconductors and a superconductor.
Schematic comparison of the density of the states (DOS) and the excitation processes for quasielectrons and quasiholes in normal-conducting Nb (green), superconducting Nb (blue dotted), Si and GaP (green, black, and gray). The actual DOS profiles have a fine structure. It is presumed that the normal-conducting Nb, Si, and GaP have the same electron DOS curves and the same energy gap (Eg = EC − Ev) between the valence band and the conduction band. The red near-horizontal arrow designates the valence-to-conduction band excitation, creating an e–h pair. The superconducting energy gap 2Δ is magnified by a factor of 200. The Fermi-function-like gray dashed curve at EF(Nb) is the multiplication of DOS and the electron occupation fraction in the BCS ground state at T = 0. The blue curved arrow designates the Cooper pair forming process by adding two electrons to a Fermi sea, leaving two holes below EF; the reverse process occurs when superconductivity is destroyed. The red vertical arrow designates the Cooper pair breaking process in the coherent superconducting state, creating two quasiparticles (qp1 and qp2) that have quantum superposition of an electron of the Cooper pair and a hole left behind when forming the Cooper pair.
Schematic comparison of the density of the states (DOS) and the excitation processes for quasielectrons and quasiholes in normal-conducting Nb (green), superconducting Nb (blue dotted), Si and GaP (green, black, and gray). The actual DOS profiles have a fine structure. It is presumed that the normal-conducting Nb, Si, and GaP have the same electron DOS curves and the same energy gap (Eg = EC − Ev) between the valence band and the conduction band. The red near-horizontal arrow designates the valence-to-conduction band excitation, creating an e–h pair. The superconducting energy gap 2Δ is magnified by a factor of 200. The Fermi-function-like gray dashed curve at EF(Nb) is the multiplication of DOS and the electron occupation fraction in the BCS ground state at T = 0. The blue curved arrow designates the Cooper pair forming process by adding two electrons to a Fermi sea, leaving two holes below EF; the reverse process occurs when superconductivity is destroyed. The red vertical arrow designates the Cooper pair breaking process in the coherent superconducting state, creating two quasiparticles (qp1 and qp2) that have quantum superposition of an electron of the Cooper pair and a hole left behind when forming the Cooper pair.
In Si and GaP at 0 K, all states above Ec are vacant and all states below Ev are occupied by electrons. The Fermi level EF of intrinsic semiconductors is placed at the middle of the energy gap. In a superconducting state, the superconducting energy gap 2Δ emerges at EF according to Eq. (10). The EF of Nb is 5.32 eV. Figure 6 clearly depicts the dissimilarity between the bandgap Eg and the superconducting gap 2Δ. Later, however, they are identically treated in statistical physics for the fluctuation of the number of charge carriers.
An excitation path in semiconductors is designated by the red near-horizontal arrow for Si in Fig. 6. An electron always accompanies a hole, forming the e–h pair according to charge neutrality, as illustrated in Fig. 3. When they meet, they recombine and annihilate. During electron energy relaxation, phonons are emitted. All the phonons cannot contribute to carrier creation in semiconductors because the maximum phonon energy or the Debye energy of Si is 55.6 meV, which is exceedingly smaller than an Eg value of 1.1 eV. In Nb, similar e–h pair creation beyond Eg initially occurs, regardless of whether Nb is in a superconducting state or not. In superconducting Nb, the electrons in a few eV range break Cooper pairs, exciting quasiparticles at levels far higher than Δ. Phonons are emitted during quasiparticle relaxation processes and on the occasion of recombination into Cooper pairs. The striking difference from semiconductors is that the Debye energy of 23.7 meV of Nb is higher than 2Δ/e (3.04 meV): therefore, most phonons can break Cooper pairs again, contributing to quasiparticle generation. High-energy quasiparticles finally relax to the edge of Δ since phonons below 2Δ cannot break Cooper pairs.
The above-mentioned quasiparticle excitation in superconductors naturally postulates that Cooper pairs exist implicitly. The processes of forming and breaking Cooper pairs are complex yet essential for superconductor quantum detectors. A simple and straightforward explanation of these processes for detector operation is rarely found in the available literature. Figure 6 provides a schematic illustration that sheds light on these processes. For forming a Cooper pair, it is necessary to add two electrons to a Fermi sea. When we consider an isolated superconductor, it corresponds to the electron excitation process leaving two holes below EF as designated by the blue curved arrow near EF(Nb), holding charge neutrality or conserving the electron number. This excitation requires an energy, but the energy reduction due to phonon-mediated attractive force, which forms the Cooper pair with the electrons near EF, significantly exceeds the excitation energy. Almost no textbooks describe these holes accompanied. It can be understood by considering the gray dashed line for the electron occupation distribution in the BCS ground state at T = 0; it is different from the Fermi step function in a normal-conducting state at T = 0. The holes left behind when forming the Cooper pair apply a hole-like character to quasiparticles when the two electrons in the Cooper pair breaks up. The excitation of quasiparticles requires the preservation of superconducting coherence, provided that the number of broken Cooper pairs is not enormous. When superconductivity is broken in a region beyond a coherence length, the two electrons simply revert back to the initial two electrons, following the reverse direction of the blue curved arrow, thereby precluding the generation of quasiparticles.
Returning to the coherent state, the red vertical arrow in Fig. 6 designates Cooper pair breaking. A photon or a phonon with an energy above 2Δ creates two quasiparticles by breaking a Cooper pair, obeying the laws of momentum and energy conservation. More precisely, we should consider the excitation on the momentum k space. When a Cooper pair is broken, a pair of excitations appears at and . The character of the two excitations can be electron-like or hole-like, depending on how momentum and energy are transferred to the two excitations or whether k1 and k2 are above or below kF. On ensemble average, the quasiparticles have quantum superposition of 50%-electron and 50%-hole characters. In Fig. 6, the quasiparticles are represented by two ovals surrounding a filled circle and an open circle. When the excitation level is high enough well beyond Δ, the quasiparticles are close to either pure electron or pure hole. This is the case that we assume for the tunneling processes in Fig. 8 in Sec. III B 3.
Typical STJ structure: (a) top-view optical micrograph, (b) cross-sectional structure along the diagonal direction passing though the contact holes for the top and bottom electrodes in (a), and (c) TEM cross-sectional view near the junction center.
Typical STJ structure: (a) top-view optical micrograph, (b) cross-sectional structure along the diagonal direction passing though the contact holes for the top and bottom electrodes in (a), and (c) TEM cross-sectional view near the junction center.
Semiconductor description model for quasiparticle tunneling in the junction at a voltage bias of eUb at a finite temperature. The DC Josephson current is suppressed by a parallel magnetic field to observe the quasiparticle current. The BCS-DOS curves in Fig. 6 are rotated 90° so that tunneling occurs horizontally through the 1-nm thick AlOx insulation barrier in the middle. The Δg/e values of 1.52 and 1.25 meV represent those of the 50-μm square junction that has a layer structure of Nb(300)/Al(5)/AlOx/Al(5)/Nb(100) with the thicknesses in parentheses. Three quasiparticle tunnel paths are denoted by the arrows. All paths of 1, 2, and 3 transfer one electron from left to right. In this diagram, the quasiparticles implicitly have either pure-electron character or pure-quasihole one, ignoring the quantum superposition of electron and hole characters.
Semiconductor description model for quasiparticle tunneling in the junction at a voltage bias of eUb at a finite temperature. The DC Josephson current is suppressed by a parallel magnetic field to observe the quasiparticle current. The BCS-DOS curves in Fig. 6 are rotated 90° so that tunneling occurs horizontally through the 1-nm thick AlOx insulation barrier in the middle. The Δg/e values of 1.52 and 1.25 meV represent those of the 50-μm square junction that has a layer structure of Nb(300)/Al(5)/AlOx/Al(5)/Nb(100) with the thicknesses in parentheses. Three quasiparticle tunnel paths are denoted by the arrows. All paths of 1, 2, and 3 transfer one electron from left to right. In this diagram, the quasiparticles implicitly have either pure-electron character or pure-quasihole one, ignoring the quantum superposition of electron and hole characters.
The excitation in superconductors is called Bogoliubov quasiparticles or Bogoliubons. The original paper was published in 1958,84 but it is still an active research area,85 influencing many fields beyond superconductivity.86 The character of Bogoliubov quasiparticles changes continuously between hole-like excitation and electron-like excitation on the single-particle picture; for example, 70% electron-like or 70% hole-like characters. This article follows a simple view for detectors: the absorption of an energy quantum above 2Δ create two quasiparticles, of which characters are either purely electron-like or hole-like, that is a quasielectron–quasihole pair; the superposition is an ensemble average. In this context, e–h pairs in semiconductors and quasiparticle pairs in superconductors can be treated similarly when we consider output current pulses flowing an external readout circuit in Fig. 3, although the origin of 2Δ in superconductors are totally different from Eg in semiconductors.
Up to here, we have described what happens when energetic particles are absorbed in superconductors. How can we read out an increase of the number of quasiparticle pairs and measure the incident particle energy E0? Unlike semiconductor detectors in Fig. 3, it is impossible to collect carriers by applying a high electric field to superconductors. Fortunately, there are two readout methods: quantum tunneling in STJD biased in a subgap region and RF surface impedance change in MKID.
3. Superconductor tunnel junction (STJ) type
a. Structure of STJ type
Figure 7 shows the top view and cross section of an Nb/Al-based STJD. One of the critical requirements for analytical instruments is a large sensitive area that should be enough to acquire signals such as characteristic x rays emitted from analytes. Detector junction sizes are usually between 100 and 200 μm2 to avoid an unreasonably large junction capacitance and to ensure a pinhole free 1-nm thick tunnel barrier.
Detector junctions normally have the Al layers with a thickness of several 10 nm on both sides of the AlOx tunnel barrier layer in expectation of quasiparticle trapping effect based on energy gap engineering proposed by Booth in 1987 and Kraus in 1989.87,88 Most of the junctions have the polycrystalline structure for both top and bottom electrodes.89 The bottom electrode can be made from a single crystalline superconductor film, which is advantageous to optical photon detection.78,90 Higher Z superconductors such as Nb or Ta located at both ends of the layer structure serves as an x-ray absorber with a meaningful absorption efficiency below a few keV. Furthermore, the proximity bilayers maintaining Tc higher than 1.2 K of bulk Al are favorable for raising an operation temperature up to 0.5 K. The amorphous AlOx tunnel barrier layer formed by oxidizing the Al surface has a thickness of about 1 nm without pinholes. It was reported that less than 10% of the junction area dominantly contributed to tunnel current because of a Gaussian variation of AlOx thicknesses.91 In addition, the AlOx barrier layer had an oxygen deficiency at the interface between the Al layer and the amorphous AlOx layer.92
For connecting STJs and readout circuits, wiring leads are indispensable. The latest microfabrication technology for STJDs appears in Ref. 93, in which the junction and wiring layers are separated vertically by using a planarization process with chemical mechanical polishing (CMP). This fabrication technique has an advantage over large-scale close-packed array detectors (See Fig. 35).
b. Quasiparticle tunneling paths for detectors
When a particle impinges the STJD at <Tc/10, a similar Isub current flows, although quasiparticle excitation distribution is different from the thermal equilibrium distribution in Fig. 8. The readout circuit for STJDs resembles that for semiconductor detectors in Fig. 3. Similarly, a bias source injects a negligibly small current to the junction biased in Ub < (Δ1 + Δ2)/2e when the DC Josephson current is suppressed by applying a parallel magnetic field. Therefore, on the analogy of semiconductor detectors, the experimental charge output value should correspond to the number of quasielectron-quasihole pairs rather than the number of total quasiparticles, holding charge neutrality or conserving the number of electrons. Statistical fluctuation should be also treated by the number of quasiparticle pairs because they are created simultaneously. In this context, Eg in semiconductors and 2Δ in superconductors is equivalent as the minimum energy gaps against electron–hole pair creation. We should pay attention that Eg is insensitive to the density of e–h pairs in semiconductors, whereas Δ is dependent on the quasiparticle density.
c. Response to particle incidence
STJ output pulse and sub-gap structure: For proper detector operation, junctions are cooled below ∼Tc/10, eliminating thermal excitation. Particle incidence induces energy-mode imbalance.96 A possibility of charge-mode imbalance will be discussed separately.50,97 The increase of the quasiparticles generates a current pulse through the tunnel paths when a junction is biased in a subgap region. The typical bias point for particle detection in our experiments is just below Δg/e or just after the beginning of a rapid current increase, as indicated by the vertical dashed line on the I–U curve in Fig. 9(a). Although it exceeds the optimum bias point for maximum pulse height, this bias point is effective in avoiding bias-point jumps due to Fiske steps.98 It is also beneficial for operating many junctions simultaneously and stably.
STJD response to a 6-keV x-ray photon: (a) current–voltage (I–U) curve at 0.47 K and (b) output current pulse at a bias voltage of 0.60 mV. The bias point is indicated by the vertical dashed line in (a). The 200-μm square junction has a layer structure of Nb(200)/Al(30)/AlOx/Al(30)/Nb(200) with thicknesses in nm in the parentheses, an energy gap of Δg(0 K)/e = 0.83 meV, a normal resistance of Rn = 23.1 mΩ, and a critical current density ic of 200 A/cm2. The dashed arrow in (a) denotes the readout circuit load line. The solid line in (b) shows a quasiparticle effective lifetime τeff of 1.42 μs.
STJD response to a 6-keV x-ray photon: (a) current–voltage (I–U) curve at 0.47 K and (b) output current pulse at a bias voltage of 0.60 mV. The bias point is indicated by the vertical dashed line in (a). The 200-μm square junction has a layer structure of Nb(200)/Al(30)/AlOx/Al(30)/Nb(200) with thicknesses in nm in the parentheses, an energy gap of Δg(0 K)/e = 0.83 meV, a normal resistance of Rn = 23.1 mΩ, and a critical current density ic of 200 A/cm2. The dashed arrow in (a) denotes the readout circuit load line. The solid line in (b) shows a quasiparticle effective lifetime τeff of 1.42 μs.
When a particle incident event creates quasiparticles, the bias point shifts from almost zero current to a high current according to the 210-Ω load line (dotted arrow) in Fig. 9(a). A cryostat wiring lead resistance of 160 Ω and a readout circuit input resistance of 50 Ω determine the load line. This voltage shift may affect quasiparticle tunneling processes slightly, but it is negligible in an x-ray energy range below 1 keV. Figure 9(b) shows the current pulse for a 6-keV x-ray photon from a 55Fe source. The rise time is determined by a readout bandwidth, whereas the fall time corresponds to the effective quasiparticle lifetime τeff. The single exponential decay curve (solid line) has a τeff value of 1.42 μs, which is significantly longer than a material-dependent intrinsic lifetime because of the trapping effect of phonons created upon quasiparticle recombination into Cooper pairs.99 In addition, phonon escape to a substrate should be considered.100 The integration of the current pulse equals to the detected charge corresponding to the number of the created quasiparticle pairs or the particle kinetic energy E0. The charge pulse for 6 keV has ∼106 electrons, which is surely higher than ∼1600 electrons in Si. Even though no strong charge collection mechanism exists in superconductors, the extremely high output-charge ensures a high signal/noise (S/N) ratio and a high-energy resolution statistically.
In Fig. 9(a), the subgap current begins to increase considerably at Δg/e, which deviates from the BCS-DOS prediction. The current increase at 2Δg/ne, where n is integral, was attributed to multiple Andreev reflection (MAR) before,101–103 although other mechanisms were also proposed. It is called the subharmonic gap structure. On the MAR model, multiple charges of ne are transferred at the bias voltage 2Δg/ne. It was believed that probable pinholes in an AlOx barrier layer caused MAR.103 However, the gap structure may also exist in pinhole-free junctions in a variety of barrier transparency.104 Furthermore, no pinholes were found in a 1-nm thick AlOx layer.91 The junctions for analytical instruments in this article are classified as a low transparency category with jc = ∼200 A/cm2 to obtain high resistance in a subgap region. Pinholes are unlikely at this low jc, but the anomaly clearly appears at n = 2, although higher-order subharmonic gap structure is obscure in the inset of Fig. 9(a). The subharmonic structure without pinholes can be explained by multiple cycles of quasiparticle sequential tunneling (MQT) proposed by Kozorezov et al.105 The quasiparticles that tunnel via path 1 and path 2 in Fig. 8 acquire an energy of eUb. The forward-and-back multiple tunneling causes energy accumulation. This MQT process leads to current jumps at 2Δg/e n, where n is the number of tunneling times. Although the MQT model is most promising, the origin of the subharmonic gap structure is still a long-standing open question.104
Pulse height spectrum: Figure 10 shows a histogram of output charge values for the x-ray photons of 6.00 keV from a synchrotron radiation source. The x rays irradiated an area with a diameter of 5 μm at the center of the 146-μm square Nb/Al junction with 30-nm-thick Al layers.106 The histogram of photon counts per a small energy bin as a function of detected energies is termed the pulse height spectrum. The energy resolution ΔE value of 92 eV for the bottom electrode exceeds a typical ΔE value of 120 eV for Si detectors. However, Fig. 10 exhibits two disadvantages of STJDs in this x-ray energy range.
Pulse height spectrum or X-ray spectrum for monochromatic 6.00-keV photons from synchrotron radiation microbeam with a diameter of 5 μm.106 The microbeam irradiate the center of the 146-μm square Nb(200)/Al(30)/AlOx/Al(30)/Nb(200) junction. Each channel has a bin width of 1 eV. The energy nonlinearity is recognized from the positions of the peaks for Si K-L2.3 (1.74 keV), the Nb L escape (3.84 keV), and 6.00 keV. In the semi-log plot inset, one can recognize the Si K-L2,3 and the Nb L escape peaks. The pulser peak shows an electrical noise of 20 eV. The 6.00-keV photons were absorbed both in the top and bottom electrodes, producing the double peak.107,108 Reproduced with permission from Pressler et al., IEEE Trans. Appl. Super. 11, 696 (2001). Copyright 2001 IEEE.
Pulse height spectrum or X-ray spectrum for monochromatic 6.00-keV photons from synchrotron radiation microbeam with a diameter of 5 μm.106 The microbeam irradiate the center of the 146-μm square Nb(200)/Al(30)/AlOx/Al(30)/Nb(200) junction. Each channel has a bin width of 1 eV. The energy nonlinearity is recognized from the positions of the peaks for Si K-L2.3 (1.74 keV), the Nb L escape (3.84 keV), and 6.00 keV. In the semi-log plot inset, one can recognize the Si K-L2,3 and the Nb L escape peaks. The pulser peak shows an electrical noise of 20 eV. The 6.00-keV photons were absorbed both in the top and bottom electrodes, producing the double peak.107,108 Reproduced with permission from Pressler et al., IEEE Trans. Appl. Super. 11, 696 (2001). Copyright 2001 IEEE.
The first disadvantage is that 90% of the incoming 6-keV photons pass through the whole junction layer structure, which produces a double peak due to the absorption events in either the top electrode or the bottom one.107,108 The double peak phenomenon is primarily due to a difference in τeff between the top and bottom electrodes. The top Nb electrode is covered by a few nm thick NbOx metallic layer, which acts as a quasiparticle trap, leading to a short τeff.107,109 Because of the low absorption efficiency of the Nb electrodes, the Si substrate absorbs 90% of the incoming photons, which induced phonon excitation. The junction detects these high-energy phonons, appearing as a large number of phonon events (or substrate events) below the channel number 60 in Fig. 10. The second disadvantage is that the peak positions of the Si K-L2,3 (Kα1,2) events at 1.74 keV and the Nb-L escape events at 3.84 keV in Fig. 10 are not proportional to E0. This is termed energy nonlinearity, which disturbs the assignment of peaks appearing on pulse height spectra in material analysis. The section describes the energy nonlinearity or linearity and how to solve the problem.
Energy linearity: We measured the energy linearity of the 200-μm square junction with the 70-nm thick Al layers in detail. A synchrotron radiation microbeam with a diameter of 10 μm irradiated the center of the junction at 6, 7, 8, 9, and 10 keV. Figure 11 shows pulse heights as a faction of photon energies with the data of the Si K-L2,3 and Nb L escape peaks. The solid line of the junction with the 70-nm-thick Al layers has an extremely better linearity up to 8 keV than the dashed line of the junction with the 30-nm Al layers in Fig. 10. Furthermore, the 70-nm STJD covers up to ∼30 keV with a tolerable energy nonlinearity as displayed in the lower right inset, which plots the data for photons and multiply charged lysozyme ions (14 305 Da) accelerated at 7 kV.110 Beyond ∼30 keV, a strong nonlinearity may prevent proper particle energy measurement. The upper-left inset shows the full illumination peak data at energies less than 1 keV. The full illumination energy resolution at 400 eV was 28 eV with an electrical noise of 25 eV, which corresponds to an intrinsic energy resolution of 13 eV.
Energy linearity of the 200-μm square STJD with a layer structure of Nb(200)/Al(70)/AlOx/Al(70)/Nb(200)/Si in nm in the parenthesis. For comparison, the dashed curve with the open squares depicts the non-linear data of the STJD with the 30-nm thick Al layers in Fig. 10. A synchrotron radiation microbeam with a diameter of 10-μm irradiated the center of the junction at 6, 7, 8, 9, and 10 keV. The lower right inset shows the data of the photons and the multiply charged ions of lysozyme (14 305 Da), accelerated at 7 kV. The upper-left inset shows the full illumination data for a range less than 1 keV. The full illumination energy resolution at 400 eV was 28 eV with an electrical noise of 25 eV, which corresponds to an intrinsic energy resolution of 13 eV.
Energy linearity of the 200-μm square STJD with a layer structure of Nb(200)/Al(70)/AlOx/Al(70)/Nb(200)/Si in nm in the parenthesis. For comparison, the dashed curve with the open squares depicts the non-linear data of the STJD with the 30-nm thick Al layers in Fig. 10. A synchrotron radiation microbeam with a diameter of 10-μm irradiated the center of the junction at 6, 7, 8, 9, and 10 keV. The lower right inset shows the data of the photons and the multiply charged ions of lysozyme (14 305 Da), accelerated at 7 kV. The upper-left inset shows the full illumination data for a range less than 1 keV. The full illumination energy resolution at 400 eV was 28 eV with an electrical noise of 25 eV, which corresponds to an intrinsic energy resolution of 13 eV.
Energy resolution: To increase the absorption efficiency and expand the covering energy range, the thickness of the top Nb layer was increased to 300 nm. Figure 12 shows a pulse height spectrum of 400-eV full illumination for a 100-μm square asymmetric junction having a layer structure of Nb(300)/Al(70)/AlOx/Al(70)/Nb(100)/Si in nm. The absorption coefficients for the 300-nm-thick Nb top film are 80% at 200 eV, 99.7% at 400 eV, and 92.2% at 700 eV;111 the effect of the double peak is negligible below ∼1 keV. The FWHM value at 400 eV is 4.1 eV, which is 10 times better than 45 eV of the latest 10 mm2 silicon drift detector (SDD).112 After subtracting an electrical noise of 2.0 eV, the full illumination intrinsic energy resolution is 3.6 eV. A better energy resolution in this soft x-ray range was obtained with a Ta/Al-based junction developed by Friedrich et al. primarily because α-Ta has a lower Δg/e of 0.7 meV than 1.52 meV of Nb. The Ta/Al junction marked 2.5 eV at 250 eV and 4 eV at 1 keV including an electrical noise of 2 eV.77
Full illumination pulse height spectrum of the 100-μm square Nb(300)/Al(70)/AlOx/Al(70)/Nb(100)/Si junction in nm for the 400-eV photons from synchrotron radiation. The 4.1-eV energy resolution includes an electrical noise of 2.0 eV, which corresponds to an intrinsic resolution of 3.6 eV. The positions of the K lines for the light elements are indicated for reference. The dotted line shows the peak with an FWHM of 45 eV of the latest 10-mm2 Si drift detector (SDD).112 The top Nb layers absorbs 99.7% of the incoming 400-eV photons, leading to no double peak. The low-energy tail stems from a spatial non-uniformity, which will be eliminated by a shadow mask in future. Material analysis results are shown in Figs. 2 and 44.
Full illumination pulse height spectrum of the 100-μm square Nb(300)/Al(70)/AlOx/Al(70)/Nb(100)/Si junction in nm for the 400-eV photons from synchrotron radiation. The 4.1-eV energy resolution includes an electrical noise of 2.0 eV, which corresponds to an intrinsic resolution of 3.6 eV. The positions of the K lines for the light elements are indicated for reference. The dotted line shows the peak with an FWHM of 45 eV of the latest 10-mm2 Si drift detector (SDD).112 The top Nb layers absorbs 99.7% of the incoming 400-eV photons, leading to no double peak. The low-energy tail stems from a spatial non-uniformity, which will be eliminated by a shadow mask in future. Material analysis results are shown in Figs. 2 and 44.
The energy resolution of the Nb/Al-STJD exceeds the typical natural linewidths of elements embedded in a matrix. For example, the carbon atoms in CaCO3 have a K-line width of ∼14 eV (Sec. I A, Fig. 2). In addition, the Zn L3-M4,5 (Zn L) line and other lines in a dentin sample has a linewidth of ∼15 eV (Section V B 1 d, Fig. 44). Broadening of characteristic lines exceeds the level widths of electronic shells because of chemical bonds to other elements or electronic band structure. Therefore, no further energy resolution improvement is necessary for elemental mapping analysis in a range below ∼2 keV.
d. Effect of Abrikosov vortex trapping
Magnetic environments in the course of the cooling process considerably affect output charge values. We observed that the output charge is significantly reduced after cooling in the geomagnetic field. This reduction stems from Abrikosov vortices (AVs) vertically piercing the junction layer structure. Since polycrystalline Nb films normally behave as a type II superconductor, AVs can be trapped while crossing Tc.113,114 The effects of AV trapping were studied by low temperature scanning electron microscope (LTSEM) at Tübingen.83,115 The reduction of output signal amplitude and quasiparticle effective lifetime as a function of perpendicular magnetic field strengths were observed at 1.6 K.113 Moreover, LTSEM signals indirectly visualized each AV trapped in junctions with electron-beam stimulated signals.116
We directly visualized AVs by measuring magnetic field distribution with a SQUID microscope.117 To investigate AV trapping patterns, the 200- or 400-μm square junctions were cooled in a perpendicular magnetic field between 0 and 7.3 μT. Figure 13 shows the dependence of AV trapping patterns on the perpendicular magnetic field strengths during cooling. The black dots in Fig. 13(a) denote the AVs carrying one magnetic flux quantum Φ0 of 2.067 × 10–15 Wb. The number of AVs increases with increasing the magnetic fields from 1.3 to 7.3 μT. Above ∼10 μT, the SQUID microscope was unable to distinguish individual AVs because the AV density exceeds a spatial resolution of ∼20 μm. We need to shield a geomagnetic field of 35.4–35.9 μT at AIST-Tsukuba118 for proper STJ operation.
Abrikosov vortex trapping in the Nb(200)/Al(30)/AlOx/Al(30)/Nb(200)/MgO(20)/Si junction cooled in a different magnetic field B between 0 and 7.3 μT:117 (a) SQUID microscope images of the 400-μm square junction and (b) AV density as a function of applied perpendicular magnetic fields for the 400- and 200- μm square junctions. The MgO layer, which is occasionally inserted as a stopping layer for reactive ion etching, is magnetically negligible. It was difficult to keep the same temperature for different magnetic fields unfortunately. The magnetic field strengths are toned from white to black. The black dots correspond to AVs of which direction is downward. In (b), the solid lines, which are linearly fitted to the data, cross the abscissa at 0.68 μT for 200 μm and 0.95 μT for 400 μm. The dotted line is a relation calculated by BL2/Φ0, where L is the junction side length. Reproduced with permission from Ohkubo et al., AIP Conf. Proc. 605, 55 (2002). Copyright 2002 AIP Publishing LLC.
Abrikosov vortex trapping in the Nb(200)/Al(30)/AlOx/Al(30)/Nb(200)/MgO(20)/Si junction cooled in a different magnetic field B between 0 and 7.3 μT:117 (a) SQUID microscope images of the 400-μm square junction and (b) AV density as a function of applied perpendicular magnetic fields for the 400- and 200- μm square junctions. The MgO layer, which is occasionally inserted as a stopping layer for reactive ion etching, is magnetically negligible. It was difficult to keep the same temperature for different magnetic fields unfortunately. The magnetic field strengths are toned from white to black. The black dots correspond to AVs of which direction is downward. In (b), the solid lines, which are linearly fitted to the data, cross the abscissa at 0.68 μT for 200 μm and 0.95 μT for 400 μm. The dotted line is a relation calculated by BL2/Φ0, where L is the junction side length. Reproduced with permission from Ohkubo et al., AIP Conf. Proc. 605, 55 (2002). Copyright 2002 AIP Publishing LLC.
One of the striking features in Fig. 13(a) is that no AVs are trapped near the junction edges. The width of the AV-free edge region decreases with increasing magnetic field. This phenomenon is explained by vortex nucleation under a surface barrier of superconductor films, based on the classical Bean and Livingstone surface barrier model.117 The AV densities are plotted against magnetic fields in Fig. 13(b). The lower critical field Hc1 values are 0.68 μT for the 200-μm junction and 0.95 μT for the 400-μm junction: the large junction has the higher Hc1. In addition, the straight line fitted to the data of the 400-μm junction has a lower gradient than that of the 200-μm junction. It signifies that the large junction size is advantageous to exclude the magnetic flux.
AVs prevent quasiparticle generation, turning Cooper pairs into normal electrons along the reverse direction of the blue curved arrow Fig. 6. Moreover, they cause quasiparticle lifetime shortening and thus output charge reduction.113,114 The quasiparticle lifetime depends on AV density and excited quasiparticle density. Figure 14 exhibits these two loss processes. The absorption events for 5.9-keV x rays are plotted according to output charge values and the pulse rise times of the charge-sensitive preamplifier. The rise times correspond to the τeff values. The 50-μm square junction with the 40-nm thick Al layers was cooled with and without the magnetic shield having an attenuation factor of well over 1000. The x-ray events form the total absorption group at 8.2 × 106 electrons at ∼0 μT and 3.6 × 106 electrons at ∼35 μT. The events due to photoelectron escape, Si K-L2,3, and Nb L escape continuously connect the total absorption event group and the substrate event group.
Scatterplot of effective quasiparticle lifetime τeff against output charge values for the 50-μm square Nb(200)/Al(40)/AlOx/Al(40)/Nb(200)/Si junction cooled at either 0 μT inside a magnetic shield or 35 μT in a geomagnetic field.114 A 55Fe x-ray source fully illuminated the junction. The total absorption events form the groups at 8.2 × 106 electrons after cooling at 0 μT and at 3.6 × 106 electrons after cooling at 35 μT. The curvature at 0 μT indicates strong self-recombination above ∼5 × 106 electrons: the excited quasiparticles recombine into Cooper pairs with themselves. Reproduced with permission from Ohkubo et al., Nucl. Instrum. Methods in Phys. Res. A 444, 237 (2000). Copyright 2000 Elsevier.
Scatterplot of effective quasiparticle lifetime τeff against output charge values for the 50-μm square Nb(200)/Al(40)/AlOx/Al(40)/Nb(200)/Si junction cooled at either 0 μT inside a magnetic shield or 35 μT in a geomagnetic field.114 A 55Fe x-ray source fully illuminated the junction. The total absorption events form the groups at 8.2 × 106 electrons after cooling at 0 μT and at 3.6 × 106 electrons after cooling at 35 μT. The curvature at 0 μT indicates strong self-recombination above ∼5 × 106 electrons: the excited quasiparticles recombine into Cooper pairs with themselves. Reproduced with permission from Ohkubo et al., Nucl. Instrum. Methods in Phys. Res. A 444, 237 (2000). Copyright 2000 Elsevier.
The quasiparticle loss rate is given by 1/τeff = 1/τr + 1/τAV, where τr is the quasiparticle lifetime against recombination into Cooper pairs without AV trapping and τAV is the effective quasiparticle loss time due to AVs. The τr values at ∼0 μT were 2.2 μs for the low-energy substrate events and 1.0 μs for the total absorption events of the 6-keV x rays. The continuous τr reduction with increasing output charge at 0 μT represents self-recombination into Cooper pairs.90 The higher the density of the excited quasiparticles is, the shorter the τr is in a condition that thermal excitation is negligible at 0.3 K. τr was independent of photon energies below ∼2 keV, in which STJDs excellently operate in a linear regime.
The trapped AVs lead to τeff reduction with a factor of 4 and output charge reduction with a factor of 2.3. The AV quasiparticle loss rate 1/τAV is given by 1.82nvπReff2/τ0, where nv is the AV density, Reff is the effective vortex radius, and τ0 is the material dependent characteristic time.113,114,119 τeff for the low-energy events is shortened by the AVs from 2.2 to 0.7 μs, which gives τAV = 1.0 μs. The τeff for the total absorption events are shortened from 1.0 to 0.5 μs, which also gives τAV = 1.0 μs. Therefore, the τAV estimation is consistent. Considering nv = 1.7 × 1010 m−2, τ0 = 0.149 ns,99 and ξ = 23 nm33 or 10 nm108 for the polycrystalline Nb films, the τAV value leads to Reff ≈ 1.5ξ or 3ξ.114 This Reff range is consistent with the values in the literature.113,119
The proper detector operation requires no AVs. Complete exclusion of AVs requires a magnetic shield with an attenuation factor of over 40–50. We are fortunate in not installing a massive magnetic shield for SQUID-MEG but using a simple shield. Nevertheless, electric parts near the detector chip often include ferromagnetic materials, which degrades the detector performance as shown in Fig. 14. Therefore, careful selection of the electric parts is necessary.
e. Superconducting gap engineering and optimum thickness of Al layers
In this section, we selectively use Δ for the order parameter and Δg for the bottom edge of a DOS profile. In other words, the order parameter Δ describes the Cooper pair potential, whereas the gap Δg is the minimum excitation energy appearing on the quasiparticle DOS profiles. Gap engineering involves two superconductors, of which Δg values are different. A combination of Al and either Nb or Ta is common for STJDs. In the “trapping model” of gap engineering, quasiparticles excited inside an Nb layer with an intact Δg,Nb/e value of 1.52 meV are trapped inside an Al layer with a lower Δg/e value, which varies between 1.52 meV and the intact Δg,Al/e value of 0.17 meV depending on the Al thicknesses. The trapping model is certainly valid and effective to collect quasiparticles when a tunnel junction is attached to a very thick bulk superconductor with a higher gap or when two junctions are attached at both ends of a-few-10-μm long superconductor film for imaging detection as proposed by N. Booth87 and H. Kraus.88,120,121 However, for direct detection with a thin-film tunnel junction geometry, this trapping model should be reconsidered and modified according to the theoretical studies on the proximity effect.33,122–124 The “gap-reduction model” is proposed in this section below.
Table II lists the parameters based on the trapping and gap-reduction models for the STJs with 200-nm thick Nb layers and different Al thicknesses. We obtained the gap Δg, effective quasiparticle lifetime τeff, and output-charge number Qc/e from experimental data. The excited quasiparticle number Q0/e(=E0/ɛ) in the Nb absorber is calculated based on the Monte Carlo result;71 it should be noted that the used relation is not ɛ = 1.72Δg for creating one quasiparticle but ɛ = 3.44Δ or ɛ = 1.72 × 2Δg for one quasielectron–quasihole pair, as discussed in Sec. III B 3 b. The ratio of the observed charge number Qc/e to Q0/e gives the average tunneling times . The tunneling probability Ptun values are valid for electrodes with a uniform Δg value: they are listed only for the gap-reduction model to discuss the multiple quasiparticle tunneling model in Sec. III B 3 f.
Comparison of basic parameters between the trapping model and the gap-reduction model for 5.9-keV x rays (E0). The columns for the gap Δg/e, the effective quasiparticle lifetime τeff, and the output charge number Qc/e list the experimental data for junctions with Al thicknesses of 20, 30, 40, and 70 nm and a Nb thickness of 200 nm. The column for the excited quasiparticle-pair number Q0/e lists the values calculated by E0/ɛ. For estimating the ɛ values, we used the intact Δg,Nb/e value of 1.52 meV for the trapping model and the Δg values from I–U curves for the gap-reduction model through ɛ = 1.72 × (2Δg) for creating a quasielectron–quasihole pair.69,71 The column for the average number of quasiparticle tunneling times lists the values calculated by dividing Qc by Q0. The tunneling probability Ptun values are used for quasiparticle multiple tunneling in Sec. III B 3 f.
Al thickness (nm) . | Gap engineering model . | Δg/e (meV) . | τeff (μs) . | Q0/e = E0/ɛ (106 e) . | Qc/e (106 e) . | . | Ptun (=1/G) . |
---|---|---|---|---|---|---|---|
5 | Trapping | 1.52 | 0.12 | 1.13 | 0.29 | 0.26 | |
Gap-reduction | 1.39 | 1.23 | 0.24 | 0.19 | |||
20 | Trapping | 1.52 | 0.55 | 1.13 | 1.7 | 1.50 | |
Gap-reduction | 1.08 | 1.59 | 1.07 | 0.52 | |||
30 | Trapping | 1.52 | 1.4 | 1.13 | 4.0 | 3.5 | |
Gap-reduction | 0.83 | 2.07 | 1.9 | 0.66 | |||
40 | Trapping | 1.52 | 1.5 | 1.13 | 8.2 | 7.3 | |
Gap-reduction | 0.65 | 2.64 | 3.1 | 0.76 | |||
70 | Trapping | 1.52 | 1.9 | 1.13 | 8.5 | 7.5 | |
Gap-reduction | 0.44 | 3.90 | 2.2 | 0.69 |
Al thickness (nm) . | Gap engineering model . | Δg/e (meV) . | τeff (μs) . | Q0/e = E0/ɛ (106 e) . | Qc/e (106 e) . | . | Ptun (=1/G) . |
---|---|---|---|---|---|---|---|
5 | Trapping | 1.52 | 0.12 | 1.13 | 0.29 | 0.26 | |
Gap-reduction | 1.39 | 1.23 | 0.24 | 0.19 | |||
20 | Trapping | 1.52 | 0.55 | 1.13 | 1.7 | 1.50 | |
Gap-reduction | 1.08 | 1.59 | 1.07 | 0.52 | |||
30 | Trapping | 1.52 | 1.4 | 1.13 | 4.0 | 3.5 | |
Gap-reduction | 0.83 | 2.07 | 1.9 | 0.66 | |||
40 | Trapping | 1.52 | 1.5 | 1.13 | 8.2 | 7.3 | |
Gap-reduction | 0.65 | 2.64 | 3.1 | 0.76 | |||
70 | Trapping | 1.52 | 1.9 | 1.13 | 8.5 | 7.5 | |
Gap-reduction | 0.44 | 3.90 | 2.2 | 0.69 |
On the trapping model, the Q0/e values for quasiparticle-pairs are calculated by E0/(3.44 × 1.52 meV) for the intact Nb absorber. On the other hand, the gap-reduction model predicts that Nb/Al electrodes have a constant Δg value throughout the bilayer. The Δg/e values of 0.44–1.39 meV in Table II are evaluated for the different Al thicknesses by extrapolating I-U curve slopes at 2Δg/e to zero current. The Q0/e values on the gap-reduction model are calculated by E0/(3.44×Δg) with reduced Δg values inside the Nb layer, provided that the DOS has sufficient room for the 5.9-keV photon absorption.
Figure 15(a) illustrates the trapping model for the proximity tunnel junction with a symmetric layer structure of Nb(200)/Al(30)/AlOx/Al(30)/Nb(200), of which I–U curve and output pulse are shown in Fig. 9. The order parameter Δ and the gap Δg are plotted as a function of the distances from the Al/AlOx interface. Δ and Δg are equal or different depending on the circumstances. In an ideal BCS-DOS profile, which has a singularity at Δg on the trapping model, we can treat it with Δ = Δg. Because of the Nb/Al proximity effect, the Δg (0 K)/e value of 0.832 meV at the Al/AlOx interface is considerably higher than the intact Al value of Δg,Al/e = 0.17 meV. On the trapping model, Δg inside the Nb layers is equal to the Nb intact value (Δg,Nb/e = 1.52 meV). This surely occurs when Nb/Al interfaces are dirty and have a low transparency. Figure 15(a) illustrates a case that an x-ray photon or a particle is absorbed inside the Nb layer on the left side. After the initial photoionization and electron energy loss, the quasiparticles are excited to high levels. They relax close to the edge of Δg,Nb afterward. Once quasiparticles are trapped in the left Al layer, they cannot enter the Nb layers again because of the higher Δg wall at the Nb/Al interfaces. The quasiparticles confined within the two Al layers can tunnel forth and back before they recombine into Cooper pairs. Since the Al layers are usually several times thinner than Nb absorber layers, the tunneling probability that is inversely proportional to the film thicknesses becomes high. Consequently, the quasiparticles created inside the Nb layer efficiently contribute to the output charge. The tunneling time enhancement is significant for the junctions with the Al layers of 30-nm or thicker in Table II, leading to gains in the ratio of Qc/e to Q0/e or , as depicted in Fig. 16 later. One can explain the experimental data with the trapping model; however, tends to be unreasonably high in some cases. A more plausible interpretation is possible by the gap-reduction model described below.
Schematic DOS profiles and quasiparticle dynamics on the trapping model (a) and the gap-reduction model (b) for the proximity junction in Fig. 9. The blue solid lines depict the order parameter or Cooper pair potential Δ as a function of the distances from the tunnel barrier. The experimental Δg values were obtained from the I–U curves. The schematic DOS profiles are displayed at the Nb free surface, the middle of the Nb layer, and the Al/AlOx interface. The trapping model in (a) assumes the BCS-like DOS profiles with the intact Δg/e value of 1.52 meV inside the Nb layer and 0.832 meV inside the Al layers. For the gap-reduction model in (b), the DOS profiles are inferred from the theory on the proximity effect.33,122–124 Both Nb and Al layers have the constant Δg value throughout the bilayers. The holes created by the Andreev reflection at the Nb/Al interface or inside the Nb layers are not at the correct energy level.
Schematic DOS profiles and quasiparticle dynamics on the trapping model (a) and the gap-reduction model (b) for the proximity junction in Fig. 9. The blue solid lines depict the order parameter or Cooper pair potential Δ as a function of the distances from the tunnel barrier. The experimental Δg values were obtained from the I–U curves. The schematic DOS profiles are displayed at the Nb free surface, the middle of the Nb layer, and the Al/AlOx interface. The trapping model in (a) assumes the BCS-like DOS profiles with the intact Δg/e value of 1.52 meV inside the Nb layer and 0.832 meV inside the Al layers. For the gap-reduction model in (b), the DOS profiles are inferred from the theory on the proximity effect.33,122–124 Both Nb and Al layers have the constant Δg value throughout the bilayers. The holes created by the Andreev reflection at the Nb/Al interface or inside the Nb layers are not at the correct energy level.
Al thickness dependence of the average quasiparticle tunneling times based on the trapping model and the gap-reduction model, and the tunneling probability Ptun for 6-keV x-ray absorption. The blue Ptun curve is plotted here for the discussion in Sec. III B 3 f. The red dashed curve on the trapping model and the red solid curve on the gap-reduction model were calculated by using the data in Table II.
Al thickness dependence of the average quasiparticle tunneling times based on the trapping model and the gap-reduction model, and the tunneling probability Ptun for 6-keV x-ray absorption. The blue Ptun curve is plotted here for the discussion in Sec. III B 3 f. The red dashed curve on the trapping model and the red solid curve on the gap-reduction model were calculated by using the data in Table II.
Figure 15(b) illustrates the gap-reduction model, in which non-negligible DOS exists below the intact Δg,Nb/e of 1.52 meV in the Nb layer with a thickness scale of a few 100 nm in accordance with the Usadal equations.125 The schematic DOS profiles at different locations are based on the calculation results.33,122–124 The Δg/e value of 0.832 meV obtained from the I–U curve holds constant at any location throughout the Nb/Al bilayer. The high-energy quasiparticles relax to the edge of Δg/e = 0.832 meV inside the Nb layer through a possible short trap at the DOS peak energy corresponding to the intact Δg,Nb/e of 1.52 meV.122 The relaxed quasiparticles tunnel to the counter electrode and spread inside the counter Nb/Al bilayer because of the constant Δg: there is no trapping effect. As an example, an Nb(100)/Al(90) bilayer keeping a constant Δg/e of 0.5 meV has a DOS value of 1.43 × 106 states/meV/μm3 at a kinetic energy of 0.3 × Δg,Nb and at the Nb free surface.33 This DOS value is large enough for accepting the quasiparticles excited by a 6-keV x-ray photon; an excited quasiparticle cloud of ∼106 can be accommodated within an energy bin of 0.1 meV inside a presumed Nb volume of 20 × 20 × 0.2 μm3. Since the quasiparticles spread throughout the whole electrode volume, the quasiparticle density is considerably reduced by a factor of ∼100 and thus far below the allowable DOS. For x rays below ∼1 keV, the Nb layer has more DOS room at 0.832 meV. Consequently, we can treat the Nb/Al bilayers as uniform superconductors. This enables a simple treatment for multiple quasiparticle tunneling in Sec. III B 3 f. If the DOS at the Nb free surface is insufficient, quasiparticle densities vary along the horizontal direction in Fig. 15(b), and nonlinear effects such as DOS vanishing can occur below the intact Δg,Nb.
The exact quasiparticle distribution in the proximity bilayer depends on many factors such as particle energies, relaxation processes from a-few-eV electrons to the edge of Δg, Andreev reflection, quasiparticle reflection at the Nb free surface, and dynamic processes concerning the depth dependence of the DOS profiles. At this moment, there is no exact analytical solution of this quasiparticle dynamics. An intermediate situation between (a) and (b) can arise depending on film thicknesses, Nb/Al interface transparency, and particle energies.
Let us compare the two models with respect to a junction with very thick Al layers: Nb(150)/Al(200)/AlOx/Al(200)/Nb(240) in nm fabricated by Mears at al.126 They reported that the 2Δg/e of the Al layers was 0.34 meV at 0.95 K. The Δg/e value of 0.17 meV is equal to that of the intact Al. This is understandable from the extremely thick Al layers: the proximity effect due to the Nb layer is negligible. The trapping model was used to explain the experimental data: the value becomes 10 by dividing Qc/e (=2.3 × 107 electrons) by E0/ɛ (=2.3 × 106 electrons), assuming that x rays are absorbed in the top Nb layer with the intact Δg/e value of 1.52 meV. Since the E0/ɛ value should be the number of created quasiparticle pairs, the corrected value is 20. One should not anticipate this significantly high value for the 200-nm thick Al layers because of the inverse relationship between the quasiparticle tunneling probability and film thicknesses. Alternatively, based on the gap-reduction model, the Δg/e values are constant at 0.17 meV throughout the bilayers with enough DOS. The initial quasiparticle-pair number inside the 150-nm thick top Nb layer is evaluated at E0/ɛ = 1.0 × 107 electrons, which gives . This value on the gap-reduction model is more plausible than 20.
Consequently, the gap-reduction model predicts that the most striking feature of gap engineering is not the strong quasiparticle trapping, but primarily the Δg reduction in the high Z absorber like Nb. The Δg reduction inside the Nb layer leads to the increase of the initially excited quasiparticles. However, most papers up to now have followed the trapping model for thin-film junction detectors. In addition, the transparency at Nb/Al interfaces and the strength of the proximity effect affect the exact DOS profiles inside Nb or Ta layers significantly,127 so that experimental results should vary depending on thin-film deposition conditions. Hopefully, MKIDs with a bilayer geometry may play a crucial role to verify the gap-reduction model better than STJDs. MKIDs simply utilize surface impedance change due to quasiparticle excitation in superconductor strips as described in Sec. III B 4. In addition, when we fabricate a RF filter with a bilayer geometry, the gap-reduction model should be applied for a low power range with a small number of quasiparticles. The gap-reduction model in Fig. 15(b) is proposed in this article for the first time; therefore, further experimental and theoretical studies are required.
We need to optimize the Al layer thickness for analytical instruments within the framework of the gap-reduction model. Table II indicates that the τeff values become longer monotonically with increasing Al thickness, which is beneficial to collecting quasiparticles. This is consistent because Al has an intrinsic quasiparticle lifetime extremely longer than that of Nb: τ0(Nb) = 0.149 ns and τ0(Al) = 438 ns.99 Figure 16 shows the curves for 6-keV x-ray absorption as a function of the Al thicknesses. The blue solid curve for Ptun on the gap-reduction model is discussed in Sec. III B 3 f. The red dashed curve on the trapping model is plotted for comparison. The red solid curve for the gap-reduction model exhibits the maximum value of 3.1 at 40 nm. At 40 nm, the output charge is considerably enhanced by the two mechanisms: lowering Δg in the Nb layers and keeping a reasonably high of 3.1. At 70 nm, slightly decreases to 2.2. Nevertheless, the energy linearity is the best at 70 nm in Fig. 11. The Al thickness of 70 nm is also beneficial to a spatial uniformity in Sec. III B 3 g. Therefore, we adopt 70 nm to realize high-energy resolution XRS.
f. Multiple quasiparticle tunneling model
The junction with the 70-nm thick Al layers in Table II has the parameters of Ptun = 0.69 (G = 1.4). Prediction of the theoretical ΔE limit requires a F value. The Fano factor is the variance ratio of a Monte Carlo simulation peak to the Poisson distribution with the same average number , where Ni is the carrier number and n is the simulation times. It gives . When we take the Monte Carlo simulation result of F = 0.214 for all quasiparticles in Nb,71 it should be modified for quasiparticle pairs. Two quasiparticles are formed simultaneously when a Cooper pair is broken; therefore, the statistical fluctuation should be treated by the number of quasiparticle pairs. The and Ni values become half for the quasiparticle pairs, which gives F = 0.107. Therefore, Eq. (16) suggests that the most fluctuation originates from the tunneling noise rather than the Fano fluctuation. In contrast, the charge collection in semiconductor detectors with a strong electric field is equivalent to , Pt = 1, and Pb = 0 and then G = 0. The high G value of STJDs is one of the disadvantages. Nonetheless, the effect of the extremely small Δg exceeds the extra tunneling noise: STJDs outperform semiconductor detectors by a factor of ∼10 experimentally.
Let us compare the experimental data on the energy resolution with the theoretical calculation based on the gap-reduction model in Table II. Figure 17 shows experimental intrinsic ΔE values as a function of photon energies below 1400 eV for the 100-μm square junctions with the Al layer thicknesses of 30 and 70 nm.130 The monochromatic photons from synchrotron radiation uniformly irradiated the whole junction area. The electrical noises of 6.5 eV for 30 nm and 7.0 eV for 70 nm were subtracted from the measured ΔE values. The red and blue solid lines depict the theoretical curves for 30 and 70 nm, respectively. For the theoretical energy resolutions, we used the Δg and Ptun(=1/G) values in Table II and ɛ = 1.72 × 2Δg and F = 0.107 for quasiparticle pairs, G = 1.5 at 30 nm, and G = 1.4 at 70 nm.
Comparison of photon-energy dependence of the intrinsic energy resolution ΔE after subtracting electrical noise between the experimental data and the multiple tunneling model calculations with Eq. (16).130 The100-μm square symmetric Nb/Al junctions have 200-nm thick Nb layers and Al layers of either 30 or 70 nm. The whole junction area was illuminated by monochromatized x rays from synchrotron radiation. The electrical noises of 6.5 eV for 30 nm and 7.1 eV for 70 nm were subtracted from the measured ΔE values. The red and blue dots plot the experimental data, while the red and blue solid lines depict the model calculations. For the multiple tunneling model curves, it was assumed that F = 0.107 and ɛ = 1.72 × 2Δg for Nb.71 The G or 1/Ptun and Δg values for the gap-reduction model were obtained from Table II. The dotted line denotes a model curve multiplied by a factor of 7 for 70 nm. The experimental intrinsic ΔE of 3.6 eV (4.1 eV with the electrical noise) for the asymmetric STJD in Fig. 12 is also plotted by the solid black square at 400 eV. Reproduced with permission from Ohkubo et al., Nucl. Instrum. Methods in Phys. Res. A 520, 231 (2004). Copyright 2004 Elsevier.
Comparison of photon-energy dependence of the intrinsic energy resolution ΔE after subtracting electrical noise between the experimental data and the multiple tunneling model calculations with Eq. (16).130 The100-μm square symmetric Nb/Al junctions have 200-nm thick Nb layers and Al layers of either 30 or 70 nm. The whole junction area was illuminated by monochromatized x rays from synchrotron radiation. The electrical noises of 6.5 eV for 30 nm and 7.1 eV for 70 nm were subtracted from the measured ΔE values. The red and blue dots plot the experimental data, while the red and blue solid lines depict the model calculations. For the multiple tunneling model curves, it was assumed that F = 0.107 and ɛ = 1.72 × 2Δg for Nb.71 The G or 1/Ptun and Δg values for the gap-reduction model were obtained from Table II. The dotted line denotes a model curve multiplied by a factor of 7 for 70 nm. The experimental intrinsic ΔE of 3.6 eV (4.1 eV with the electrical noise) for the asymmetric STJD in Fig. 12 is also plotted by the solid black square at 400 eV. Reproduced with permission from Ohkubo et al., Nucl. Instrum. Methods in Phys. Res. A 520, 231 (2004). Copyright 2004 Elsevier.
The ΔE values of the junction with 30 nm deviate from the dependence above 500 eV abruptly, exceeding the ΔE of SDDs above ∼800 eV. This abrupt ΔE degradation is due to spatial nonuniformity as is reported in Ref. 90, which is the absorption-position dependence of output charge values. This is the topic of Sec. III B 3 g. The ΔE degradation in the high-energy range is typical for junctions with thin Al layers. Nevertheless, it is intriguing that below ∼400 eV the junction with 30 nm has a similarly small ΔE to that of the junction with 70 nm. This implies that the spatial nonuniformity is small even at 30 nm in that energy range. Direct measurement results on the Al thickness dependence on spatial nonuniformity will be shown in Fig. 21 in Sec. III B 3 g.
The junction with the 70-nm thick Al layers follows the dashed curve, which is the theoretical curve multiplied by 7. The best experimental intrinsic ΔE value for 70 nm is 9.7 eV at 500 eV, which is worse than the theoretical limit of 2.5 eV by . One of the possible reasons for the factor 7 at most energies is explained by the dip structure of the blue dots between 400 and 600 eV. The dip structure stems from the x-ray transmission of the top 200-nm-thick Nb near the Nb M-edge. The green solid line depicts the absorption coefficients of the top Nb layer.111 In the energy range between 400 and 600 eV, almost all x-ray photons are absorbed by the top Nb layer. On the other hand, at the data points below 400 eV and above 800 eV, 30% or more of the photons transmit the top Nb layer and are absorbed in the bottom Nb layer. The absorption events in the bottom Nb layer produce slightly higher output pulses. This causes the wider FWHM values on the occasion of Gaussian fitting to the experimental peaks. Another possibility is a spatial nonuniformity that slightly remains event at 70 nm as shown in Fig. 21 in Sec. III B 3 g.
Additionally, we should consider the fact that the Monte Carlo simulations were performed by assuming imparted energies of 5.75–56 eV.68,71 These energies are extremely smaller than those of keV x rays by a factor of 10–1000. In the x-ray energy rage, we should consider a cascade of radiative or nonradiative transition processes related to the K, L, and M shells as mentioned in Appendix B. The number of the e–h pairs fluctuates before considering quasiparticle pair generation. This causes additional broadening of the experimental peaks. Consequently, the possible origins for the factor 7 are the double peak formation, the spatial nonuniformity, and the e–h pair fluctuation.
g. Spatial nonuniformity of STJ response
The spatial nonuniformity occasionally exceeds other noise sources; for example, the junction with the 30-nm thick Al layers in Fig. 17. The linear ΔE degradation results in the significant deviation from the statistical fluctuation proportional to . To investigate spatial nonuniformity directly, we first employed the low temperature scanning electron microscope (LTSEM) at the University of Tübingen.83,115 The base temperature of LTSEM was ∼1.5 K at the lowest by pumping 4He with an open hole for an electron beam. Despite the high base temperature, LTSEM has a high spatial resolution of well better than 0.1 μm that enables fine imaging analysis of STJDs. The LTSEM measurement revealed that the bottom electrode of a Nb/Al junction with the 10-nm thick Al layer normally produces a higher signal than the top electrode; it is explained by a difference in τeff and oxidation of the top Nb surface.107 Furthermore, for a geometrical design guideline on wiring leads and junction shapes, we analyzed the junction shown in Fig. 18: wide wiring leads, which is necessary for measuring full I–U curves beyond Ic, and the round junction with a diameter of 113 μm, of which sensitive area is equivalent to 100-μm square. The layer structure was Nb(200)/Al(20)/AlOx/Al(20)/Nb(200)/Si in nm for the operation at ∼1.5 K. The electron-pulse beam enabled to measure 2D profiles of signal amplitudes and fall-time constants.
LTSEM imaging diagnosis of the round junction with a layer structure of Nb(200)/Al(20)/AlOx/Al(20)/Nb(200)/Si in nm: (a) color scale image of the response amplitudes to electron pulses, and (b) cross-sectional structure along the light green line in (a). The junction has a diameter of 113 μm, the 40-μm wide wiring leads, and the very large 30-μm square contact hole for the top electrode. The 10-keV electron pulses of ∼100 ns scanned the junction surface at 2.1 K. The output signal amplitudes are scaled by the rainbow color order of blue, green, yellow, orange, red, and white in (a). The white solid line depicts the signal amplitudes along the light green arrow. The pulse fall time depends on position: 210 ns on the bottom wiring lead, 150 ns at the junction center, and 190 ns on the top wiring lead at the contact hole center.
LTSEM imaging diagnosis of the round junction with a layer structure of Nb(200)/Al(20)/AlOx/Al(20)/Nb(200)/Si in nm: (a) color scale image of the response amplitudes to electron pulses, and (b) cross-sectional structure along the light green line in (a). The junction has a diameter of 113 μm, the 40-μm wide wiring leads, and the very large 30-μm square contact hole for the top electrode. The 10-keV electron pulses of ∼100 ns scanned the junction surface at 2.1 K. The output signal amplitudes are scaled by the rainbow color order of blue, green, yellow, orange, red, and white in (a). The white solid line depicts the signal amplitudes along the light green arrow. The pulse fall time depends on position: 210 ns on the bottom wiring lead, 150 ns at the junction center, and 190 ns on the top wiring lead at the contact hole center.
The 2D image in Fig. 18(a) visualizes that the amplitudes near the junction rim are lower than the junction center, which is commonly observed for any junction shape. This leads to the low energy tail of the x-ray total absorption peak as observed in Fig. 12. To eliminate the low energy tail, one solution is to fabricate a shadow mask in front of junctions so that the junction edges are covered. The white solid line exhibits the position dependence of signal amplitudes along the green arrow. It visualizes the response of the contact hole and the wiring leads. The pulse fall-time constants in Fig. 18(a) are 150 ns in the top electrode and 210 ns in the bottom electrode or bottom lead just near the junction, which causes the double peak phenomenon.107,108 In the contact hole, the fall time of 190 ns is longer than 150 ns on the top electrode. This can be attributed to quasiparticle diffusion time to the junction area in the thick Nb as shown in Fig. 18(b). The LTSEM observation suggests that contact holes must be small and wiring leads must be narrow as much as possible to achieve a high-energy resolution.
It is necessary for analyzing STJDs at a normal operation temperature of ∼0.3 K. We developed the imaging instrument using an x-ray microbeam with a diameter of 5–10 μm from synchrotron radiation: low temperature scanning synchrotron microscope (LTSSM) proposed by Pressler.106,131,132 The LTSSM employs 3He sorption cryostat and x-ray windows that block infrared radiation, which ensures ∼0.3 K. Before and after the LTSSM studies, location-selective x-ray irradiation experiments were performed in Refs. 133–135. However, precise full 2D x-ray imaging analysis was realized only by LTSSM.
Junction structure for the quasiparticle diffusion model.136 The model deals with quasiparticle diffusion D, tunneling probability from top to bottom γt-b, tunneling probability from bottom to top γb-t, reflection at the junction edges R, quasiparticle loss via phonon escape and other processes γloss = 1/τeff. For symmetric junction, γ = γt-b = γb-t. The red regions denote quasiparticle clouds due to diffusion.
Junction structure for the quasiparticle diffusion model.136 The model deals with quasiparticle diffusion D, tunneling probability from top to bottom γt-b, tunneling probability from bottom to top γb-t, reflection at the junction edges R, quasiparticle loss via phonon escape and other processes γloss = 1/τeff. For symmetric junction, γ = γt-b = γb-t. The red regions denote quasiparticle clouds due to diffusion.
The diffusion model was applied to the 146-μm square junction with a τeff value of 1.4 μs and a l value of 10 nm in Fig. 10.106 The fitting parameters are R, D, and γ in Eq. (18). Figure 20 shows the comparison between the model calculation and the LTSSM experimental data of the bottom electrode events at bias points of 0.15 and 0.30 mV, both of which are well below Δg/e = 0.841 meV. The 5-keV x-ray microbeam scanned the square junction diagonally. The best fit to the experimental profiles was obtained at D = 1.8 cm2/s, R = 0.999952, γ = 1/0.594 μs−1 for 0.15 mV and D = 4.5 cm2/s, R = 0.999916, γ = 1/0.343 μs−1 for 0.30 mV. The R values are slightly less than 1.0, which implies that the edge loss results in the convex dome shape.
Comparison in the diagonal scan profiles between the quasiparticle diffusion model and the LTSSM diagonal scan data at 5 keV for the 146-μm square junction with a layer structure of Nb(200)/Al(30)/AlOx/Al(30)/Nb(200)/Si in nm at a parallel magnetic field of 31.6 mT.106 The LTSSM data points were extracted from the total absorption peaks for the bottom electrode events. The fixed parameters are γloss = 1/1.4 μs and l = 10 nm. The best fits were obtained with β = 0.353, Λ = 8.65, Γ = 0.703 (D = 1.8 cm2/s, R = 0.999952, γ = 1/0.594 μs−1) at 0.15 mV and β = 0.616, Λ = 11.26, Γ = 0.803 (D = 4.5 cm2/s, R = 0.999916, γ = 1/0.343 μs−1) at 0.30 mV. Reproduced with permission from Pressler et al., IEEE Trans. Appl. Super. 11, 696 (2001). Copyright 2001 IEEE.
Comparison in the diagonal scan profiles between the quasiparticle diffusion model and the LTSSM diagonal scan data at 5 keV for the 146-μm square junction with a layer structure of Nb(200)/Al(30)/AlOx/Al(30)/Nb(200)/Si in nm at a parallel magnetic field of 31.6 mT.106 The LTSSM data points were extracted from the total absorption peaks for the bottom electrode events. The fixed parameters are γloss = 1/1.4 μs and l = 10 nm. The best fits were obtained with β = 0.353, Λ = 8.65, Γ = 0.703 (D = 1.8 cm2/s, R = 0.999952, γ = 1/0.594 μs−1) at 0.15 mV and β = 0.616, Λ = 11.26, Γ = 0.803 (D = 4.5 cm2/s, R = 0.999916, γ = 1/0.343 μs−1) at 0.30 mV. Reproduced with permission from Pressler et al., IEEE Trans. Appl. Super. 11, 696 (2001). Copyright 2001 IEEE.
It is reasonable that the γ values change with the bias points, since the net tunneling probability increases as the bias voltage increases in a range below the maximum tunnel current according to the BCS-DOS profile in Fig. 8. A diffusion constant of 1.0 cm2/s for polycrystalline Nb in the literature109 is not inconsistent with the present value of 1.8 cm2/s at 0.15 mV. However, the D value of 4.5 cm2/s at 0.30 mV is unreasonable. Since the quasiparticle diffusion constant should be essentially unaffected by the bias points. Therefore, we should consider factors other than quasiparticle diffusion.
It is surprising that, in a photon energy range below 300 eV, the junction with the 30-nm thick Al layers has almost the same energy resolution as the junction with the 70-nm Al layers, as shown in Fig. 17. This signifies that the abnormally large spatial nonuniformity emerges in a high photon energy range like 5 keV. The LTSSM scan at 300 eV failed because of the optics problem; however, it is apparent that the spatial uniformity rapidly degrades above ∼500 eV for the 30-nm thick Al layers. The concrete explanation of this behavior is difficult at this moment. Here, it is only pointed out that for practical usage of STJDs, we should pay attention to photon energies and Al thicknesses. Figure 21 shows the LTSSM scans at 6 keV for the junction with different Al thicknesses:138 the spatial uniformity is improved by increasing Al thickness significantly. One possibility of this behavior is that the thick Al layers lead to a large quasiparticle diffusion due to the long τeff, and, thus, the spatial nonuniformity becomes insignificant. It is also probably related to the gap-reduction model in Fig. 15. The thicker the Al layers are, the more the DOS below the intact Δg,Nb exists in the Nb layer. This leads to a linear response even to the 6-keV photons. For the 30-nm thick Al layers, the DOS below Δg,Nb is less, but still enough for the quasiparticle density at photon energies below 300 eV. The possible mechanism of the spatial uniformity degradation at the high-energy range will be discussed considering the MQT process and bias points separately.50
h. Count rate
One of the important performance figures is the count rate (count per second: cps). For analytical instruments with imaging capability, a high count rate is crucial for a tolerable measurement time to scan an area to be analyzed. Figure 22(a) shows the energy resolution of the 100-μm square symmetric junction with the 70-nm thick Al layers as a function of photon count rates at 400 eV. Because of a large electrical noise of 17 eV in this experimental run at a synchrotron radiation beamline, the ΔE value was 20 eV instead of 12 eV in Fig. 17. No energy resolution degradation was observed up to ∼6 kcps that is the maximum photon flux at the bending magnet beamline. A single junction can be operatable up to ∼50 kcps without significant degradation of the energy resolution.139,140 However, this count rate is still 10 times lower than a typical count rate of several 100 kcps of SDDs.112 Therefore, we need to array STJ pixels and readout circuits simultaneously to realize a comparable count rate. It is possible to produce a detector chip with 1024 pixels,141 but a typical complete system with readout circuits has 100-to-200 channels.
Al layer thickness dependence of 6-keV LTSSM diagonal scans for the 200-μm square junction biased near Δg/e.131,138 The scan step was 10 μm. The parallel magnetic fields ranged from 10 to 17.6 mT. The photon absorption event numbers are toned from black to white. The y axis is normalized at the highest output charge. Both top and bottom electrodes absorb the 6 keV photons, leading to the double peak. The inset for Al (30 nm) is the full 2D image of the top electrode events. Reproduced with permission from Pressler et al., Appl. Phys. Lett. 77, 4055 (2000). Copyright 2000 AIP Publishing LLC. Reproduced with permission from Ukibe et al., Nucl. Instrum. Methods in Phys. Res. A 520, 260 (2004). Copyright 2004 Elsevier.
Al layer thickness dependence of 6-keV LTSSM diagonal scans for the 200-μm square junction biased near Δg/e.131,138 The scan step was 10 μm. The parallel magnetic fields ranged from 10 to 17.6 mT. The photon absorption event numbers are toned from black to white. The y axis is normalized at the highest output charge. Both top and bottom electrodes absorb the 6 keV photons, leading to the double peak. The inset for Al (30 nm) is the full 2D image of the top electrode events. Reproduced with permission from Pressler et al., Appl. Phys. Lett. 77, 4055 (2000). Copyright 2000 AIP Publishing LLC. Reproduced with permission from Ukibe et al., Nucl. Instrum. Methods in Phys. Res. A 520, 260 (2004). Copyright 2004 Elsevier.
Count-rate performance of STJDs: (a) FWHM energy resolutions ΔE of the 100-μm square single junction shown in Fig. 17, and (b) FWHM values of the C K-L2,3 line measured with the 100-pixel array detector in Fig. 12. In (a), the 400-eV synchrotron radiation beam with an FWHM of better than ∼0.1 eV illuminated the whole junction area. The energy resolutions around 20 eV are worse than 12 eV in Fig. 17 because of a high electrical noise of 17 eV at the synchrotron radiation beam line. In (b), a carbon nanotube electron gun excited a conductive carbon tape, producing the characteristic x rays of the C K, O K, and Al L-lines. The Bremsstrahlung x rays are also detected. The inset in (b) displays an x-ray spectrum with the photon count rate per a 1-eV bin in kcps, integrating all the 100 pixels at a photon count rate of 100 kcps. By subtracting the STJD energy resolution of 6 eV from the measured 12.5 eV in (b), the natural linewidth of the C K-line is 11.0 eV.
Count-rate performance of STJDs: (a) FWHM energy resolutions ΔE of the 100-μm square single junction shown in Fig. 17, and (b) FWHM values of the C K-L2,3 line measured with the 100-pixel array detector in Fig. 12. In (a), the 400-eV synchrotron radiation beam with an FWHM of better than ∼0.1 eV illuminated the whole junction area. The energy resolutions around 20 eV are worse than 12 eV in Fig. 17 because of a high electrical noise of 17 eV at the synchrotron radiation beam line. In (b), a carbon nanotube electron gun excited a conductive carbon tape, producing the characteristic x rays of the C K, O K, and Al L-lines. The Bremsstrahlung x rays are also detected. The inset in (b) displays an x-ray spectrum with the photon count rate per a 1-eV bin in kcps, integrating all the 100 pixels at a photon count rate of 100 kcps. By subtracting the STJD energy resolution of 6 eV from the measured 12.5 eV in (b), the natural linewidth of the C K-line is 11.0 eV.
Figure 22(b) shows the count-rate performance of a 100-pixel array STJD with a ΔE value of 6 eV in Fig. 12. Briefly, the signal readout circuit for the 100 pixels consists of arrays with charge-sensitive preamplifiers, flash ADCs, and digital signal processing (DSP) circuits that have pulse shaping processors and multi-channel analyzer memories for the 100 pixels. The details of the readout circuits are described in Sec. IV B. For operating the 100 pixels simultaneously and stably, the bias point was just after the rapid subgap-current increase near Δg/e as usual. The 100 pixels had very similar I–U curves, but each pixel had own gain with a typical relative standard deviation of 4% (standard deviation divided by average peak position on pulse height spectra). Therefore, before the count-rate measurement, energy calibration of each pixel was performed by using known characteristic x-ray peaks. A control software automatically processes this gain calibration for the 100 pixels. When precise gain calibration is necessary for measuring ∼1-eV peak shift in chemical analysis, as shown in Fig. 47 later, non-linear off-line calibration should be performed by using a few characteristic peaks. For this count-rate measurement, the linear calibration using one peak was enough, because the natural K-line width of the carbon atoms in the carbon tape was 11.0 eV after subtracting the STJD energy resolution from the measured 12.5 eV.
For the 100-pixel STJ array detector in Fig. 22(b), we employed an x-ray source using a carbon nanotube electron gun142 that irradiated a conductive carbon tape on an Al plate, producing the characteristic x rays of the C K, O K, Al L-lines, and Bremsstrahlung. This count-rate test reproduces an actual material analysis condition like SEM. Real-time acquisition of pulse height spectra was performed without off-line data processing. Figure 22(b) shows the FWHM values for the C K-line as a function of the photon count rates including all characteristic x rays and Bremsstrahlung. The FWHM values including an electrical noise of 2 eV begin to degrade rapidly above 200 kcps, which is the upper limit for x-ray material analysis. Insufficient bias current supply to each junction at a high count rate limits the count rate. The count rate of 200 kcps is the same as that of the 25-mm2 SDD in Fig. 2: this value is the highest record among superconductor detector systems. STJ arrays with 200 or more pixels will achieve a higher count rate and a larger sensitive area in future.93
4. Microwave kinetic inductance (MKI) type
Principle of microwave kinetic inductance detector (MKID): (a) circuit element arrangement, (b) equivalent circuit, and (c) resonance characteristic. Each pixel is divided into the inductive part with Ls and the capacitive part C1,2,3…. The LC circuit is coupled to the center conductor via the coupling capacitor Cc. It is possible to read out the large number of pixels with a single RF input and output line at different frequencies. Single photons or continuous particle flux break Cooper pairs in the inductor segment, which leads to the increase of Ls and Rs. This change induces the resonance frequency shift Δf and the Q-value reduction depicted by the red curve in (c).
Principle of microwave kinetic inductance detector (MKID): (a) circuit element arrangement, (b) equivalent circuit, and (c) resonance characteristic. Each pixel is divided into the inductive part with Ls and the capacitive part C1,2,3…. The LC circuit is coupled to the center conductor via the coupling capacitor Cc. It is possible to read out the large number of pixels with a single RF input and output line at different frequencies. Single photons or continuous particle flux break Cooper pairs in the inductor segment, which leads to the increase of Ls and Rs. This change induces the resonance frequency shift Δf and the Q-value reduction depicted by the red curve in (c).
Considering superconductors in an RF range, kinetic inductance due to Cooper pairs have a non-negligible contribution to the total inductance in most cases. The complex surface impedance is expressed by Zs = Rs + iωLs, where Rs is the surface resistance and ω is the angular frequency. At temperatures lower than ∼Tc/10, a condition of Rs ≪ Ls is fulfilled since almost no quasiparticles are thermally excited. The Ls has two components of energy storage: magnetic field within a penetration depth and Cooper pair inertia. The sum of the magnetic inductance and the kinetic inductance is roughly proportional to μ0λ, where μ0 is the vacuum permeability and λ is the magnetic penetration depth. Particle incidence results in Cooper pair breaking and produces quasiparticles. The decrease of the Cooper pair density leads to the increase in λ and thus in Ls since the London penetration depth is inversely proportional to the square root of the density of Cooper pairs. In addition to the change in Ls, Rs increases because of quasiparticles dissipating an energy. The increase of Ls and Rs induces the resonance curve change shown in Fig. 23(c). The absorbed energy is determined by measuring the resonance amplitude change or the degree of phase change.
An advantage of the MKI type is that a single microwave cable is enough to read out thousands of pixels. The covering energy range is between infrared light and soft x rays for the single-particle detection mode without additional absorber structure. Particle flux measurement is also possible. The MKID type is suitable in a range of near-infrared to ultraviolet for astrophysics,143,144 and frequently used for cosmic microwave background (CMB) radiation.145,146 The performance, physics, and applications appear in the latest LTD workshop proceedings.147
For optical/IR astronomy, the figures of merit are compared for MKID, TESD, and STJD in Ref. 148. They concluded that the MKI type is the most promising because of a tremendous number of pixels and a large sensitive area. This is a reasonable conclusion in the field of optical/IR astronomy. One of the latest MKID applications is the 20 000-pixel camera for the 8-m Subaru telescope at Mauna Kea.149 The MKID camera has a photon detection capability of arrival time, location, and photon energy, which cannot be realized by charge coupled device (CCD) cameras in a near-infrared range of 800–1400 nm.
The MKI type can play a crucial role in investigating the relation between ɛ and Δg in superconductors, since output signals reflect the number of broken Cooper pairs in a simple superconductor strip form unlike the complex signal generation processes in STJDs. In addition, the gap-reduction model for superconducting gap engineering in Fig. 15(b) can probably be verified by using proximity bilayer films.
C. Superconductor thermal detectors
1. A historical overview
Superconductor thermal detectors, such as TESD and SSPD, are also installed into analytical instruments for synchrotron radiation x-ray analysis, mass spectrometry, infrared spectroscopy, and visible-infrared spectroscopy.150–154 In this article, SSPD is categorized as thermal detection, since it utilizes S-N transition accompanied with heating and cooling processes due to ETF.155–157 The history of thermal detector development is to realize a smaller C value and a higher temperature-sensitivity dR/dT value for TESDs or to induce S-N transition with ETF in SSPDs.
For calorimeters, single-particle events must produce a reasonably large temperature rise to be measured with a thermometer.158 Since the heat capacity of solids decreases as temperature decreases, it is effective to cool down an absorber at a cryogenic temperature. In a cryogenic temperature range, however, there were no proper thermometers in the 1930s. After the discovery of superconductivity in mercury in 1911, the possible use of sharp S-N transition as a thermometer was proposed in 1939–1941.159,160 The first successful superconducting phase transition (SPT) bolometer was reported by Andrews in 1942.161 This was also the first practical application of superconductivity in all fields, which was 13 years earlier than a superconductor electromagnet in 1955.162 An SPT thermometer using a Ta wire coil was attached to an Al sheet absorber with a thickness of 80 μm and an area of 8 × 10 mm2. The Al absorber sheet was supported by silk threads for thermal isolation. He observed a detector response that was consistent with the Stefan–Boltzmann law for blackbody radiation between 24 and 55 K. Figure 24 depicts the superconducting transition of the Ta wire coil in (a), the equivalent circuit of the SPT bolometer in (b), and the temperature balance diagram in (c). The thermal circuit was designed such that the heat escape to a heat bath balanced with the Joule self-heating of the Ta coil thermometer that is in the middle of S-N transition.
Electrothermal feedback (ETF) concept of the first superconducting phase transition (SPT) bolometer in1942:161 (a) superconducting transition of the Ta wire coil thermometer, (b) equivalent circuit based on the IEC 60 617 database for graphical symbols,163 and (c) temperature balance diagram. In (b), the symbols with two dots and a right arrow denote the S-N boundaries: the side of the right arrow is in a superconducting state. The resistor symbol with the attribute of temperature dependence denotes the partially superconducting Ta coil. Rw is the resistance of wiring leads from room temperature to a cryogenic base temperature of 2.3 K. The bias resistance Rb is exceedingly larger than R + 2Rw. In (c), Tb is the base temperature and G is the heat conductance due to the silk threads. The switch in (b) selects two bias modes with and without Rs: I2 ≈ I1/[1 + (R + 2Rw)/Rs)]. The Rs value is adjusted such that the I22R curve is more gradual than the GΔT curve in (c) to stabilize the heat balance. When the switch is turned on, the temperature of the Ta coil is automatically regulated within the middle of S-N transition. The potentiometer reads the voltage change across the Ta coil. The modern TES type utilizes a SQUID amplifier with zero input resistance instead of the potentiometer. The SQUID amplifier allows the readout of current change under constant UT. This rotates the green I22R line to the orange dotted line UT2/R clockwise, strengthening ETF.
Electrothermal feedback (ETF) concept of the first superconducting phase transition (SPT) bolometer in1942:161 (a) superconducting transition of the Ta wire coil thermometer, (b) equivalent circuit based on the IEC 60 617 database for graphical symbols,163 and (c) temperature balance diagram. In (b), the symbols with two dots and a right arrow denote the S-N boundaries: the side of the right arrow is in a superconducting state. The resistor symbol with the attribute of temperature dependence denotes the partially superconducting Ta coil. Rw is the resistance of wiring leads from room temperature to a cryogenic base temperature of 2.3 K. The bias resistance Rb is exceedingly larger than R + 2Rw. In (c), Tb is the base temperature and G is the heat conductance due to the silk threads. The switch in (b) selects two bias modes with and without Rs: I2 ≈ I1/[1 + (R + 2Rw)/Rs)]. The Rs value is adjusted such that the I22R curve is more gradual than the GΔT curve in (c) to stabilize the heat balance. When the switch is turned on, the temperature of the Ta coil is automatically regulated within the middle of S-N transition. The potentiometer reads the voltage change across the Ta coil. The modern TES type utilizes a SQUID amplifier with zero input resistance instead of the potentiometer. The SQUID amplifier allows the readout of current change under constant UT. This rotates the green I22R line to the orange dotted line UT2/R clockwise, strengthening ETF.
When the switch in Fig. 24(b) is turned off, it operates at the balance point between the blue line and the black line: G • ΔT = I12R(T), where G is the heat conductance between the Al absorber and the heat bath. However, this balance point is unstable because I12R is steeper than G • ΔT. When the switch is turned on, the balance stability is improved: G • ΔT = I22R(T) provides the stable balance. By choosing a suitable Rs value, the temperature of the Ta coils is automatically regulated within the middle of the S-N transition. For radiation input, I2 decreases because of the increase of R, which provides a weak negative ETF. The details of the ETF operation of the first SPT bolometer will be described elsewhere.50
It is surprising that the first superconductor bolometer had the self-regulation circuit to automatically hold the temperature within the S-N transition width. However, the potentiometer–galvanometer combination in 1942 required a sufficient voltage change across the Ta coil, which precludes the strong negative ETF operation with constant voltage biasing. The modern ETF operation was realized by a SQUID amplifier based on a superconductor galvanometer (SLUG-SQUID) conceived in 1966.164 The modern TES type proposed by Irwin in 1995 utilizes a condition of Rs ≪ R and Rw = 0 with SQUID readout circuitry;165,166 therefore, constant UT can be applied to an SPT or TES thermometer and current change can be measured. The self-heating power of the TES thermometer is expressed by UT2/R, which leads to a negative slope. This bias scheme clockwise rotates the green I22R line to the dotted orange UT2/R line in (c). As temperature increases or R increases because of the heat input J, the self-heating power is reduced according to UT2/R, which is called strong negative ETF.165
A Mo/Au-TESD marked an energy resolution of 1.58 eV for 5.9-keV x-ray photons with a fall time of probably ∼1 ms in 2012.167,168 Conventional calorimeters with a semiconductor thermistor thermometer had not reached to this impressive ΔE value. Besides an x-ray range, TESDs perform well for photons in an infrared-to-optical range. In addition to STJDs, a W-TESD achieved an energy resolution of 0.15 eV and a fall time of 60 μs for photons above 0.3 eV by Cabrera et al. in 1998.81 The fall time was improved to 200 ns in 2009 by Fukuda et al.169 Furthermore, an Au/Ti TESD by Hattori et al. marked a ΔE value of 0.067 eV and a fall time of 4.6 μs in 2022.170
Magnetic calorimetry using Er-doped garnets was proposed and demonstrated for 5.5 MeV α-particles by Bühler and Umlauf in 1988.171,172 The metallic magnetic calorimeter using a metallic host, which has a stronger spin-phonon interaction than dielectric garnet hosts, was proposed by Bandler et al. in 1993.173 Enss et al. demonstrated spectroscopy for x rays and γ rays in 2000.174 The MMCD marked an energy resolution of 1.58 eV and a fall time of ∼1 ms for 5.9-keV x-ray photons.175
An unconventional thermal detector type utilized a simple superconductor strip (SS type or SSPD), which appeared 38 years later after the discovery of superconductivity. A superconductor NbN strip of 6-μm thick and 0.4 × 3.5-mm square detected α-particles in 1949, which was reported by Andrews,176 who was also the author of the SPT bolometer.161 This was the first SSPD for single-particle counting, although the term of bolometer was used. The NbN strip was firmly mounted on a copper base for a good thermal contact unlike conventional bolometers or calorimeters. Voltage output pulses were produced by a resistive change due to α-particles under constant current bias. When it was held in the middle of S-N transition just like TES, the best counting performance was obtained.
A different operation mode for superconductor strips maintained at a temperature well below Tc was proposed many years later in 1962.177 The strip was supposed to be biased near a half of critical current (Ic). The presumed detection mechanism for 5-MeV α-particles or 60-MeV fission fragments was an S-N transition of a Sn strip of 100-nm thick and 10-μm wide. No growth of the normal region with Joule heating was expected in this simulation, which was different from modern SSPDs with ETF heating. An experimental paper on this detection scheme was reported in 1965.178 The SS type was also capable of detecting electrons from a beta source with an NbN strip of 0.36-μm thick, 4-μm wide, and 2-mm long in 1990.179 It was also possible to detect 6-keV x rays with a W strip of 0.24-μm thick, 1.8-μm wide, and 0.55-mm long in 1992.180 The W strip by Gabutti et al. operated in either a high bias current region for hot spot formation and subsequent Joule heating or a low bias region for normal region formation bridging the strip width without Joule heating. The recovery times were in a range of 10–50 ns, which were ∼10 times longer than modern SSPDs; nonetheless, the former detection mode is the same as the one proposed for the detection of infrared-to-optical photons.
The infrared photon detection succeeded with a nanomized NbN nanostrip of 5-nm thick, 200-nm wide, and 1-μm long by Gol'tsman et al. in 2001.181 The proposed operation mode was the same as the W strip, that is, Joule-heating-assisted signal creation after initial hot-spot formation or recently proposed vortex motion. The photon detection in a telecommunication wavelength region with SSPDs covers applications such as quantum cryptography, quantum computing, and time-of-flight (TOF) ranging using single photon detection capability.182 Furthermore, possible application of SSPD to ion detection in TOF MS was suggested by Frank et al. in a review paper on superconductor detectors for MS in 1999.183 A possibility of kinetic inductance operation of SSPDs for MS was reported in 2005.184 Experimentally, the first successful acquisition of MS spectra for a biomolecule of bovine serum albumin (BSA) was reported by Suzuki et al. in 2008.185 Neutrons were also detected with a MgB2 strip with 10B by Ishida et al. in 2008.186
It should be noted that the pioneers conceived the ideas using superconductivity for thermal detection in the 1940s. In particular, we have reverence for Andrews, who is the pioneer of both TES type in 1942 and SS type in 1949.161,176 The modern TESDs and SSPDs have marked far better performance and far wider particle-energy coverage compared with the early years.
2. Transition edge sensor (TES) type
a. Structure and operation of TES type
The TES type usually consists of an absorber, a superconductor film thermometer, and a SQUID current–voltage conversion amplifier. Figure 25 depicts the equivalent circuit, detector structure, and I–U characteristic. The remarkable change from the first SPT bolometer161 in Fig. 24 is that the potentiometer–galvanometer combination located at room temperature is replaced by the SQUID amplifier at a cryogenic temperature. It is also noticeable that Rs is repositioned from room temperature to the vicinity of TES to hold a constant bias voltage, providing Rs ≪ RT(T).
Operation principle of electrothermal feedback (ETF) in transition edge sensor detectors (TESDs): (a) notional equivalent circuit, (b) schematic structure of the 200-μm square Ir(100)/Au(25) bilayer TES in nm, and (c) experimental IT–UT curve in green and the heating power curve in red of the Ir/Au TES.187 The graphical symbols with a cross and two dots denote Josephson junctions in (a).163 The ring with two Josephson junctions symbolizes the DC SQUID with the small cross symbol attributed to magnetic field dependence.163 The notional circuit in (a) is different from actual SQUID readout with a feedback loop. The magnetic field due to IT is transmitted to the SQUID ring by the mutual inductance between the coil L and the SQUID ring readout circuit. The top and side views of the Ir/Au TES are depicted in (b) with the thermal model in the lower panel. The heat conductance G connects the Ir/Au TES at 110 mK and the heat bath at 50 mK.
Operation principle of electrothermal feedback (ETF) in transition edge sensor detectors (TESDs): (a) notional equivalent circuit, (b) schematic structure of the 200-μm square Ir(100)/Au(25) bilayer TES in nm, and (c) experimental IT–UT curve in green and the heating power curve in red of the Ir/Au TES.187 The graphical symbols with a cross and two dots denote Josephson junctions in (a).163 The ring with two Josephson junctions symbolizes the DC SQUID with the small cross symbol attributed to magnetic field dependence.163 The notional circuit in (a) is different from actual SQUID readout with a feedback loop. The magnetic field due to IT is transmitted to the SQUID ring by the mutual inductance between the coil L and the SQUID ring readout circuit. The top and side views of the Ir/Au TES are depicted in (b) with the thermal model in the lower panel. The heat conductance G connects the Ir/Au TES at 110 mK and the heat bath at 50 mK.
In Fig. 25(a), the DC SQUID ring is magnetically coupled with the coil L connected in series with the TES–Rs circuit. The TES current is converted to magnetic flux and then voltage US. The circle with two Josephson junction symbols represents a DC SQUID with the small cross attribute for magnetic dependence.163 Although only one DC SQUID is present in (a), a SQUID series array is normally employed to obtain output voltage sufficient for signal processing at room temperature.
In the ETF operation below 1.2 μV, IT changes such that a relation of UTIT = 9.35 pW is satisfied according to Eq. (21). RT(T) decreases rapidly below the onset temperature of the superconducting transition, following UT2/9.35 pW. As a result, the temperature of the Ir/Au TESD is automatically fixed within the S-N transition width. The thermal escape of 9.35 pW is maintained by the SiNx membrane. The membrane structure in Fig. 25(b) is fragile because the backside is fully open. We conceived a novel durable structure by using a silicon-on-insulator (SOI) wafer, reinforcing the backside of the membrane with the Si substrate.188
The ETF improves the energy resolution beyond the limit of conventional calorimeters by a factor of 1/α0.5. A TESD specialized for a soft x-ray range marked an energy resolution of 1.1–1.5 eV for 500 eV x-ray photons.150,151 which is 2–3 times better than STJDs. For 6-keV photons, the best ΔE value is 1.58 eV of a Mo/Au TES with an Au or Bi/Au absorber,167,168 which outperforms the energy resolution of 29 eV of the Nb/Al-STJD with a SQUID amplifier.191 The TES type covers infrared to γ rays, designing the necessary thickness and heat capacity for absorbers. However, the best TESD energy resolution is normally obtained at a slow pulse fall time of ∼1 ms. The details of the TESD-wide applications are referred to Refs. 5 and 150.
b. Spatial nonuniformity of TES response
Spatial nonuniformity is also a problem of the TES type, since it has a dimension larger than λ or λL. Figure 26 shows LTSSM images of a 50-nm thick and 500-μm square Ir-TES on a SiNx membrane.132 The pure Ir film has Tc = 134.5 mK, ΔTc = 1.3 mK, and Rn = 2.15 Ω just above Tc. While scanning the Ir-TESD biased at 3.61 μV with a 3-keV x-ray microbeam having a diameter of 20 μm at the ETL-TERAS synchrotron radiation facility, the current pulse parameters, such as τrise, τfall, and pulse height, were recorded according to the x-ray microbeam coordinates by Pressler et al.132 The Ir film geometry in (a) is superimposed with the white line on the LTSSM signals toned by the rainbow color scale from black to white in (b)–(d). In (e), it is noticeable that the pulse height near the upper edge indicated by the red spot in (b) is three times higher than that near the TES center indicated by the light green spot. The images of τrise and τfall also show the large nonuniformities. This behavior can be explained by phase separation between a superconducting region and a normal-conducting region below.
Low temperature scanning synchrotron microscope (LTSSM) images of the pure-Ir TESD with the 3 keV x-ray microbeam with a diameter of 20-μm at the ETL-TERAS synchrotron:192 (a) optical image, (b) current pulse height, (c) rise time τrise, (d) fall time τfall, and (e) current pulse shapes at the red and light green spots in (b), and (f) IT–UT curve with the open red circle for the 3.61-μV bias point. The pulse height and the time values are toned by the rainbow color scale from black to white. The Ir film rim is superimposed on (b)–(d) with the white lines. The dotted black line in (a) and the gray lines in (b–d) indicate the S-N boundary; the exact S-N boundary is curved as shown by the black solid line in (b). The numbers in (e) indicate the τrise and τfall values. The dashed line in (f) is a normal-conducting slope at 2.15 Ω. The Ir film has Tc = 134.5 mK and ΔTc = 1.3 mK. Imaging analysis was performed at a bath temperature of 55 mK. The white vertical allows in (a) denote the predicted current density distribution. Reproduced with permission from Ohkubo et al., IEEE Trans. Appl. Super. 13, 634 (2003). Copyright 2003 IEEE.
Low temperature scanning synchrotron microscope (LTSSM) images of the pure-Ir TESD with the 3 keV x-ray microbeam with a diameter of 20-μm at the ETL-TERAS synchrotron:192 (a) optical image, (b) current pulse height, (c) rise time τrise, (d) fall time τfall, and (e) current pulse shapes at the red and light green spots in (b), and (f) IT–UT curve with the open red circle for the 3.61-μV bias point. The pulse height and the time values are toned by the rainbow color scale from black to white. The Ir film rim is superimposed on (b)–(d) with the white lines. The dotted black line in (a) and the gray lines in (b–d) indicate the S-N boundary; the exact S-N boundary is curved as shown by the black solid line in (b). The numbers in (e) indicate the τrise and τfall values. The dashed line in (f) is a normal-conducting slope at 2.15 Ω. The Ir film has Tc = 134.5 mK and ΔTc = 1.3 mK. Imaging analysis was performed at a bath temperature of 55 mK. The white vertical allows in (a) denote the predicted current density distribution. Reproduced with permission from Ohkubo et al., IEEE Trans. Appl. Super. 13, 634 (2003). Copyright 2003 IEEE.
Figure 27 shows the LTSSM 1D pulse-height profiles at different bias points.132 The black arrow in the 90°-rotated optical image in (a) indicates the x-ray microbeam scan direction. As the bias voltage increases in (b), the maximum pulse heights decrease and simultaneously the maximum positions move from right to left. The inset in (b) shows the measured Ir-TES resistance as a function of the distances between the film right edge and the open arrow positions. These observations indicate that as the bias voltage increases, the S-N boundary moves from right to left, expanding the normal region maintained by self-heating. In the course of this S-N boundary movement, the maximum pulse height values decrease and τfall values increase, which represents a reduction of α in Eq. (22). The α value monotonically decreases as UT approaches to 8.5 μV, at which the whole film becomes normal-conducting. The predicted temperature variation profiles are shown in Fig. 27(c), providing a thermal healing length of ∼700 μm. The estimated peak-to-peak temperature variation of ±8 mK exceeds ΔTc. This is caused by a low thermal conductivity of the pure Ir film. Intriguingly, the superconducting region and the normal one occasionally swapped each other by a single photon event as exhibited by the dashed line in (b). This indicates that the side of the normal region is independent of current directions. Furthermore, the photon absorption events in the superconducting region produced spike pulses that are the same mechanism as SSPDs.192 This indicates that the operation of the TES and SS types are close, which will be discussed in detail elsewhere.50
Superconducting-normal (S-N) phase separation in the pure-Ir TESD:132 (a) 90°-rotated optical image of the Ir TES with the 3 keV x-ray microbeam scan direction, (b) LTSSM 1D pulse-height profiles at the different bias voltages, and (c) predicted temperature variation profiles in the Ir film with a thickness of 50 nm. The inset in (b) shows the measured TES resistance values as a function of the distances from the film right edge to the open arrow positions, which indicate the S-N boundary locations. The normal and superconducting regions occasionally swaps by one single photon event as indicated by the dashed curve at 3.3 μV in (b). Reproduced with permission from Pressler et al., Appl. Phys. Lett. 81, 331 (2002). Copyright 2002 AIP Publishing LLC.
Superconducting-normal (S-N) phase separation in the pure-Ir TESD:132 (a) 90°-rotated optical image of the Ir TES with the 3 keV x-ray microbeam scan direction, (b) LTSSM 1D pulse-height profiles at the different bias voltages, and (c) predicted temperature variation profiles in the Ir film with a thickness of 50 nm. The inset in (b) shows the measured TES resistance values as a function of the distances from the film right edge to the open arrow positions, which indicate the S-N boundary locations. The normal and superconducting regions occasionally swaps by one single photon event as indicated by the dashed curve at 3.3 μV in (b). Reproduced with permission from Pressler et al., Appl. Phys. Lett. 81, 331 (2002). Copyright 2002 AIP Publishing LLC.
The spatial nonuniformity should be reduced to achieve a superior energy resolution. Proximity bilayer structure is crucial for increasing thermal conductivity as well as adjusting Tc. We designed the bilayer structure such that the temperature variation would be less than ∼1 mK. Figure 28 shows the LTSSM scan data of the 200-μm square Ir(100)/Au(25) bilayer in nm. The photon event number image for 6-keV x-ray microbeam in (a) visualizes the Ir/Au film geometry and the Nb leads. The LTSSM 1D pulse-height profiles in (b) are flat except the Nb lead regions, which signifies that no S-N phase separation occurs. Interestingly, the τrise image in Fig. 28(c) still exhibits a small nonuniformity between 5.0 and 6.4 μs, which conceivably represents a slight nonuniformity of bias current like Fig. 26(a). Nevertheless, the 55Fe full-illumination energy resolution was 9.4 eV. This energy resolution was not the best in 2003 but exceeded ∼120 eV of SDDs by a factor of more than 10.
Improvement of spatial uniformity in the TESD with a 200-μm square Ir/Au proximity bilayer: (a) LTSSM image for the detected event number of 6-keV photons, (b) LTSSM pulse-height profiles along the black arrow in (a) at the different bias voltages, and (c) rise time τrise image between 5 and 6.4 μs. The Ir(100)/Au(25) bilayer in nm has Tc = 110 mK. The τrise image still implies a slight bias-current nonuniformity, which causes peak broadening. A Mo/Au TES with a normal metal feature marked a ΔE value of 1.58 eV at 5.9 keV.167
Improvement of spatial uniformity in the TESD with a 200-μm square Ir/Au proximity bilayer: (a) LTSSM image for the detected event number of 6-keV photons, (b) LTSSM pulse-height profiles along the black arrow in (a) at the different bias voltages, and (c) rise time τrise image between 5 and 6.4 μs. The Ir(100)/Au(25) bilayer in nm has Tc = 110 mK. The τrise image still implies a slight bias-current nonuniformity, which causes peak broadening. A Mo/Au TES with a normal metal feature marked a ΔE value of 1.58 eV at 5.9 keV.167
For the simple square geometry, an unexpected excess noise emerges.193 The excess noise was explained by the RSJ model for Josephson weak links, for example.194,195 Additionally, it is also reasonable that the bias-current nonuniformity, which can occur in films with a dimension larger than λ, causes extra peak broadening due to pulse-height spatial nonuniformity. Experimentally, the excess noise can be reduced by arranging Cu normal metal bars on bilayer films or applying a perpendicular magnetic field.167 In fact, a Mo/Au bilayer film with the normal metal feature and the Au absorber marked a superior energy resolution of 1.58 eV at 5.9 keV.167
3. Metallic magnetic calorimetric (MMC) type
The MMC type is one of the important thermal detector types using a SQUID readout circuit. The details on the MMC operation appear in Ref. 175, for example. The MMC type is based on thermal detection with a paramagnetic temperature sensor. Figure 29 shows a schematic illustration of MMCD, which consists of the paramagnetic material M(T) and the SQUID magnetometer. The paramagnetic temperature sensor is attached to an absorber, which is often Au. A high M(T) coefficient is realized by a diluted alloy, of which temperature dependence is depicted in (b). The block of the absorber and the paramagnetic sensor is thermally connected to a low temperature heat bath via a weak link having a small heat conductance G. An example of the diluted alloys is Au:Er with an Er concentration of ∼300 ppm.174 When a particle impinges the absorber, M(T) decreases as indicated by the arrow in (b). The magnetic flux in the paramagnetic temperature sensor changes under a bias magnetic field H. The flux change is converted to a voltage pulse by the SQUID amplifier.
Operation principle of metallic magnetic calorimetric detector (MMCD): (a) schematic structure and (b) temperature dependent magnetization M(T) of a paramagnetic material, which is normally an Au–Er alloy. The particle incidence causes a temperature rise of the absorber, which is often Au. The temperature rise is transduced to a magnetic flux change in the pickup coil and then the SQUID ring under a bias magnetic field. The single DC SQUID in (a) can be replaced by a multistage design followed by a second stage SQUID array. The best energy resolution was 1.58 eV at 5.9 kV.175
Operation principle of metallic magnetic calorimetric detector (MMCD): (a) schematic structure and (b) temperature dependent magnetization M(T) of a paramagnetic material, which is normally an Au–Er alloy. The particle incidence causes a temperature rise of the absorber, which is often Au. The temperature rise is transduced to a magnetic flux change in the pickup coil and then the SQUID ring under a bias magnetic field. The single DC SQUID in (a) can be replaced by a multistage design followed by a second stage SQUID array. The best energy resolution was 1.58 eV at 5.9 kV.175
The fundamental energy resolution of the MMC type is expressed by Eq. (6) for thermal detection, limited by the energy fluctuation of a canonical ensemble. The time response or τfall is simply given by C/G. More detailed expressions appear in Ref. 175. Experimentally, an impressive energy resolution of 1.58 eV at 6 keV, which is exactly the same as the Mo/Cu TESD,168 was achieved at a base temperature of 30 mK in 2018.175 The best resolving powers are ∼4000 for x rays and ∼6000 for γ rays, which correspond to ∼10 eV in a wide energy range up to 60 keV.196 In addition to the TES type, the MMC type is utilized for XRS, γ-ray spectroscopy, neutrinoless double beta decay for Majorana nature of neutrinos,197 neutrino mass measurement with electron capture in Ho,198 MS for neutral fragments in ion storage rings,199,200 and dark matter search.7 The MMC type has no function for shortening τfall of output pulses in a range of ∼ms. Nevertheless, the MMC type with the strong spin-phonon interaction has the fastest τrise value of 70 ns among calorimeters,201 which is advantageous to the timing measurement in imaging MS.199,200
4. Superconductor strip (SS) type
a. Excitations in superconductor thin-film nanostrips
The SS type usually has a meander geometry of a superconductor nanostrip, which has typically a few nm thickness, a few 100 nm width, and ∼mm length. This nanoscale dimensions enable to count infrared-optical single photons near ∼1 eV that is 106 times smaller than MeV α-particles in the early years of microscale SSPDs. In addition to photons near 1 eV, one can detect electrons,179 x-ray photons,180,202,203 and macromolecules that generate phonons upon impact.185,204,205 Recently, ultrawide strips with also enable photon detection near 1 eV,206,207 which will be separately described in Sec. III C 4 f. SSPDs normally produce pulses with a constant height independent of particle types and energies. Although energy discrimination was possible by adjusting bias current152 or single-photon energy measurement was reported in a range between 0.4 and 2.0 eV,208 SSPDs cannot be generally employed for high resolution spectroscopy or EDS of single particles.
For detector physics, 2D approximation is appropriate in most cases209 because of the nanostrip dimensions: and . Because of the 2D nature of SSPD operation, this article follows the term used by the pioneers in the early years and the IEC terminology:6,210 “strip” or “nanostrip” by contrast to “nanowire” or “microwire.” The term of “wire” represents 1D physics, disregarding lateral extent.
Initial perturbation models for a photon energy range between ∼0.5 and 3 eV, including a telecommunication wavelength (e.g., 1.55 μm or 0.8 eV) remains controversial: bias-current assisted normal-region bridging (hot spot) proposed initially in 2001,181,211 single vortex crossing (SVC),212,213 vortex-antivortex departing (VAP),214,215 and phase slip center (PSC).209,216 The hot spot model is based on normal core formation with a temperature of higher than Tc or hot-electron core formation. Supercurrent detours around the hot spot, leading to bridging the strip width with a normal domain, inducing a voltage drop. In SVC, a single vortex crosses a strip because of Lorenz force. In VAP, Cooper pair breaking causes vortex-antivortex pair creation, depairing, and subsequent movement to the opposite directions. The vortex motion due to Lorentz force causes a voltage drop. Almost all materials for SSPDs are type-II superconductors, which allows vortex nucleation. The initial response to photons well below 1 eV can be either SVC or VAP. The SVC-VAP mixture means that both can occur at similar probabilities. In addition, hot-spot-induced VAP mechanism was also proposed.217 The SVC model is valid only for a case that w is considerably smaller than Pearl length values in Table I: .212,218 The phase slip mechanism is inappropriate for most SSPDs because of the 2D nature except relatively narrow NbN strips with w ≈ 100 nm.219 The detection principles can change depending on superconductor materials as well as the strip dimensions.
According to the physical parameters in Table I, we can assign SVC, VAP, or SVC-VAP mixture to the listed superconductor materials. For tf = ξ and w = 200 nm, the possible response to photons near 1 eV is SVC for NbN and YBCO, VAP for Nb, and SVC-VAP mixture for MgB2. As of now, none of the proposed detection mechanisms can fully explain all the experimental observations. A possibility is that the different mechanisms are engaged depending on particle energies, nanostrip dimensions, elapsed time after particle incidence events, distance from a particle incident point, and operation conditions. Further theoretical and experimental studies are required for detection mechanisms.220–222 In this article, we assume hot spot formation for keV ion detection204 and vortex motion due to either SVC or VAP for photon detection below ∼1 eV. The boundary between hot spot formation and vortex motion is at ∼1 eV for a 4-μm square NbN SSPD with tf = 4.5 nm and w = 100 nm.208
Compared with extremely wide inhomogeneous strips for LTSEM experiments,83 SSPDs consist of high-quality uniform narrow strips. Therefore, the bias current distribution along the lateral direction should be uniform. Even in this condition, the 2D nature was observed by analyzing position dependent local detection efficiency and detection threshold current for ∼1-eV photons at a spatial resolution of 10 nm.223 An ingenious method of quantum detector tomography achieved a surprisingly high spatial resolution. They observed that the local detection threshold bias-currents for an NbN film with tf = 5 nm, w = 150 nm, and l = 100 nm were slightly nonuniform: 27.5 μA at the strip center and 25.3 μA near the edge. This behavior was explained by an SVC model.223 In addition, it conceived that the threshold bias-currents are 27.5 μA for VAP and 25.3 μA for SVC. Therefore, it is concluded that most SSPDs have a 2D nature even in a strip width close to the ISO nanoscale definition of smaller than ∼100 nm.224
b. Practical SSPD structure and output signal
Figure 30(a) shows a laser scanning microscope image of the 200-μm square SSPD with a single-crystalline NbN on a MgO substrate, which was fabricated at NICT by Miki et al.225 The 200-μm square NbN-SSPD in (a) has a meander nanostrip with tf = 10 nm, w = 800 nm, and l = 40 mm, and a space of 200 nm.152, Figure 30(b) shows a typical output pulse from an analogous 50-μm square NbN SSPD with tf = 7 nm, w = 300 nm, l = 4.15 mm, a space of 300 nm, and Tc = 14 K.205 Angiotensin-I biomolecules accelerated at 17.5 kV impinged on the 50-μm square SSPD at a bias current of 170 μA, that is, 94% of Ic and 4.2 K. Angiotensin-I is a peptide biomolecule with a molecular weight (MW) value of 1296. The output pulse parameters are τrise = 360 ps and τfall = 9 ns in Fig. 30(b). Output pulse shapes were independent of MW or molecular species: biomolecules such as bovine serum albumin (BSA) of 66.4 kDa and Immunoglobulin G (IgG) of 146 kDa. Furthermore, the 50-μm square NbN-SSPD generated exactly the same output pulse shape even for 0.8-eV photons.225 The ETF process described in Sec. III B 4 c governs the output pulse shape regardless of photons and molecules at this SSPD dimensions.
Single-crystalline NbN-SSPD and output pulse:152,205 (a) laser scanning microscope image of the 200-μm square meander SSPD with Tc = 14 K and (b) output pulse of an analogous 50-μm square SSPD at a bias current of 0.94Ic and 4.2 K for an angiotensin-I ion, which is a 1296-Da peptide biomolecule. Ions produced by matrix-assisted laser desorption/ionization (MALDI) were accelerated at 17.5 kV. The 200-μm square SSPD in (a) with tf = 10 nm, w = 800 nm, l = 40 mm, and a line space of 200 nm marked τrise = 900 ps and τfall = 27 ns. A typical output pulse of the 50-μm square SSPD with tf = 7 nm, w = 300 nm, l = 4.15 mm, and a line spacing of 300 nm has τrise = 360 ps and τfall = 9 ns in (b). The x axis indicates the time-of-flight (TOF) values for a flight length of 1.37 m: the TOF value of 26.56 μs corresponds to 1296 Da. The 50-μm square SSPD in (b) also produced similar output pulses for photons near 1 eV. Reproduced with permission from Suzuki et al., Rapid Commun. Mass Spectrom. 24, 3290 (2010). Copyright 2010 John Wiley & Sons. Reproduced with permission from Suzuki et al., Appl. Phys. Express 1, 031702 (2008). Copyright 2008 IOP Publishing.
Single-crystalline NbN-SSPD and output pulse:152,205 (a) laser scanning microscope image of the 200-μm square meander SSPD with Tc = 14 K and (b) output pulse of an analogous 50-μm square SSPD at a bias current of 0.94Ic and 4.2 K for an angiotensin-I ion, which is a 1296-Da peptide biomolecule. Ions produced by matrix-assisted laser desorption/ionization (MALDI) were accelerated at 17.5 kV. The 200-μm square SSPD in (a) with tf = 10 nm, w = 800 nm, l = 40 mm, and a line space of 200 nm marked τrise = 900 ps and τfall = 27 ns. A typical output pulse of the 50-μm square SSPD with tf = 7 nm, w = 300 nm, l = 4.15 mm, and a line spacing of 300 nm has τrise = 360 ps and τfall = 9 ns in (b). The x axis indicates the time-of-flight (TOF) values for a flight length of 1.37 m: the TOF value of 26.56 μs corresponds to 1296 Da. The 50-μm square SSPD in (b) also produced similar output pulses for photons near 1 eV. Reproduced with permission from Suzuki et al., Rapid Commun. Mass Spectrom. 24, 3290 (2010). Copyright 2010 John Wiley & Sons. Reproduced with permission from Suzuki et al., Appl. Phys. Express 1, 031702 (2008). Copyright 2008 IOP Publishing.
c. ETF in SSPD
The SSPD operation resembles avalanche ionization under a high voltage in Geiger–Müller counters or avalanche photodiodes (APDs) to improve quantum efficiency. The constant pulse height implies that, on the hot spot model, an initial resistive band formed across the strip width provokes ETF due to Joule self-heating, leading to expansion and contraction of the subsequent resistive domain. In SVC and VAP, the initial resistive band formation can be replaced by local voltage drop induced by photon-assisted vortex motion.215 Figure 31 illustrates the SSPD-ETF operation assuming the uniform bias-current distribution. A photon for direct Cooper pair breaking can replace the IgG molecule ion for phonon-mediated Cooper pair breaking in (a).
Output pulse generation process due to electrothermal feedback (ETF) in the hot spot model: (a) supercurrent detour around the hot spot in the superconducting strip carrying a bias current of at t ≈ 0; (b) initial resistive band formation followed by normal domain expansion due to Joule self-heating until t < 100 ps; (c) recovery to the superconducting state, changing normal current jn to supercurrent js at t ≈ 500 ps; and (d) output voltage pulse of the amplifier with τrise = ∼500 ps and τfall = ∼20 ns. The time scale is based on the experimental observation of the 50-μm square NbN SSPD with tf = 7 nm, w = 200 nm, l = 6.25 mm, and Lk = 1.1 μH for 17.5-keV BSA ion impact.205 This SSPD also has an ability to detect single photons near 1 eV. Note that in (a) the black jc(x) is the critical current density, the red js(x) is the bias-supercurrent density before the event, the blue jb(x) is the current density distribution after the event; in (b) the gray jb(x) is the current density immediately after the normal band is formed at <100 ps, the black jn(x) is the normal current density after the normal domain growth at ∼500 ps; and in (c) the light red js supercurrent returns to the initial red js at a time constant of ∼20 ns. The amplifier load resistor plays the same role in stabilizing fast recovery as the shunt resistor for the first SPT bolometer in Fig. 24 and TESDs in Fig. 25.161,165
Output pulse generation process due to electrothermal feedback (ETF) in the hot spot model: (a) supercurrent detour around the hot spot in the superconducting strip carrying a bias current of at t ≈ 0; (b) initial resistive band formation followed by normal domain expansion due to Joule self-heating until t < 100 ps; (c) recovery to the superconducting state, changing normal current jn to supercurrent js at t ≈ 500 ps; and (d) output voltage pulse of the amplifier with τrise = ∼500 ps and τfall = ∼20 ns. The time scale is based on the experimental observation of the 50-μm square NbN SSPD with tf = 7 nm, w = 200 nm, l = 6.25 mm, and Lk = 1.1 μH for 17.5-keV BSA ion impact.205 This SSPD also has an ability to detect single photons near 1 eV. Note that in (a) the black jc(x) is the critical current density, the red js(x) is the bias-supercurrent density before the event, the blue jb(x) is the current density distribution after the event; in (b) the gray jb(x) is the current density immediately after the normal band is formed at <100 ps, the black jn(x) is the normal current density after the normal domain growth at ∼500 ps; and in (c) the light red js supercurrent returns to the initial red js at a time constant of ∼20 ns. The amplifier load resistor plays the same role in stabilizing fast recovery as the shunt resistor for the first SPT bolometer in Fig. 24 and TESDs in Fig. 25.161,165
As an example, the observed τrise was 640 ps (∼500 ps in Fig. 31) for the 50-μm square NbN SSPD. The corresponding Rk is evaluated at 1.7 kΩ, which is equivalent to a normal-domain length of 1.5 μm.205 Because of IL diverted to RL, the current flowing the strip decreases to jn(x) in (b). The heat sustaining the normal domain rapidly diffuses to the substrate unlike TESDs. When the strip temperature decreases below Tc with a typical time scale of ∼500 ps, the normal current jn is transformed to the light red supercurrent js in Fig. 31(c). The temporal change of resistive domain expansion was measured for parallel NbN nanostrips with tf = 9 nm and w = 100 nm by Ejrnaes et al.228 The expansion lengths ranged from 0.59 to 3.5 μm depending on parallel configuration. These lengths are consistent with 1.5 μm, although we cannot directly compare the parallel and meander SSPDs.
When the duration of output IL pulses is longer than the normal domain cooling time, which is primarily determined by the inelastic electron–phonon scattering time τe-p in Table I, Ik flowing the strip immediately after events is efficiently reduced.155–157 In this negative ETF, the strip returns to a superconducting state without latching in a normal state. The negative ETF strength order is TESD > SSPD > Ta-SPT. No latching in SSPDs occurs when the ratio RL/RK is well less than ∼10. On the other hand, when τfall is shorter than the cooling time, for example, a large RL, the Joule self-heating continues on and the resistive domain expands to the whole strip length, which is the positive ETF leading to latching.155–157 It is necessary for using a small RL; however, this leads to slowing down both τrise and τfall. Therefore, the time constant of the SSPD readout circuit should be designed such that continuous particle counting is as fast as possible without latching.
It is worth considering superconductor materials for fast response. The τe-p values determine the upper limit for high frequency applications for optical or quantum communication. Aluminum in Table I is an essential material for the STJ type, but the extremely long τe-p 395 μs is unsuitable for fast response. Niobium with 330 ps is acceptable, and NbN with 88 ps is favorable. Other LTS materials are also employed. Since the SS type simply utilizes the S-N transition, high temperature superconductors are promising for operation at a higher temperature regardless of multiple superconducting energy gaps or d-wave symmetry. The compounds MgB2 and YBCO with a few ps response are attractive for ultra-high-speed applications. Single photon detection with MgB2-SSPDs is reviewed by Shibata in Ref. 232. Although it is difficult to prepare uniform YBCO films, it is close to single photon detection.30 We fabricated a meander YBCO-SSPD with tf = 200 nm, w = 1.4 μm, Tc = 77 K; however, 17.5-keV molecule detection was unsuccessful because of film inhomogeneity that causes a poor I–U characteristics.26 A BSCCO-SSPD exfoliated from a bulk single crystal has recently succeeded in detecting single photons.31
d. Hot-spot model validation for keV ion incidence
We experimentally validated the above-mentioned hot-spot model for keV ion bombardment.204,233 Threshold bias current Ith values for ion detection were measured as a function of the kinetic energies E0 of Ar ions that irradiated a 200-μm square single-crystalline NbN SSPD with tf = 10 nm, w = 800 nm, and l = 40 mm. The Ar ions were charged singly, doubly, or triply by electron impact (EI) and accelerated at different voltages. Unlike biomolecules, the Ar ions exhibit no fission upon impact; therefore, fluctuation of energies imparted to the NbN film can be minimized. The Ith values were determined at the level of 1% of the saturated detection efficiency in Fig. 32(a). Figure 32(b) is the plot of the Ith values normalized with Ic as a function of E0 values. We assume that the hot spot region has a cylindrical shape, and its diameter is expressed by d = C × E00.5, where C is determined by DOS at the Fermi level, Tc, and the efficiency of quasiparticles creation.226 The dashed line in (b) depicts a calculation curve with C = 3.5 × 10−9 eV−0.5, following Eq. (24) by substituting Ith for Ib. The hot spot model curve well explains the experimental data of the Ar ions in an energy range between 0.6 and 9 keV. The hot spot diameter d is estimated at 110 nm for the 1-keV Ar ions.
Hot-spot model validation with the Ar ions in an energy range between 0.6 and 9 keV:204 (a) detection efficiency values as a function of the normalized bias currents with the dotted line at 1% for the detection threshold Ith and (b) normalized detection threshold current Ith/Ic as a function of the kinetic energies of the Ar ions. The simple meander single-crystalline NbN-SSPD on a MgO substrate has a strip with tf = 10 nm, w = 800 nm, l = 40 mm, Tc = 13 K, and Ic = 590 μA. Reproduced with permission from Suzuki et al., Appl. Phys. Express 4, 083101 (2011). Copyright 2011 IOP Publishing.
Hot-spot model validation with the Ar ions in an energy range between 0.6 and 9 keV:204 (a) detection efficiency values as a function of the normalized bias currents with the dotted line at 1% for the detection threshold Ith and (b) normalized detection threshold current Ith/Ic as a function of the kinetic energies of the Ar ions. The simple meander single-crystalline NbN-SSPD on a MgO substrate has a strip with tf = 10 nm, w = 800 nm, l = 40 mm, Tc = 13 K, and Ic = 590 μA. Reproduced with permission from Suzuki et al., Appl. Phys. Express 4, 083101 (2011). Copyright 2011 IOP Publishing.
e. Simulation with time-dependent GL equation
To track the output pulse generation process in Fig. 31 more precisely, Zen and Mawatari solved a 2D time-dependent Ginzburg–Landau (TDGL) equation coupled with a thermal diffusion equation for a MgB2-SSPD with tf = 10 nm, w = 250 nm, and l = 200 μm,24 following the method in Ref. 215. The calculation results were compared with the experimental result on 20-keV angiotensin-I ions. Figure 33 exhibits the temporal changes of the normal domain in (a), superfluid electron density at the different locations in (b), and output voltage across RL in (c). The molecule ion impact was simulated by the power input Qion depicted by the red dashed-dotted line in (b). Figure 33(b) reveals that the superfluid density becomes zero after a few ps at the incidence location. The normal domain appears and expands to a length of ∼1 μm after 16 ps because of self-heating due to the bias current, and then the MgB2 strip returns to a superconducting state after 400 ps. This time scale is in the same order as the 50-μm square NbN SSPD in Fig. 31. The highest temperature of the resistive domain was 48 K. The superfluid density curves at different positions marked by A, B, C, and D in (a) show an intriguing behavior. Curves A and B near the molecule impact location follow the smooth temporal changes without oscillation, which means that the detection mechanism is the normal-core hot-spot formation. On the other hand, at the place apart from the hot spot region, curve C exhibits a clear oscillation after 10 ps, which signifies the vortex creation and motion. Although the primary behavior is the hot spot formation, VAP depairing can occur depending on the elapsed time and locations.
Simulation with the 2D time-dependent Ginzburg–Landau (GL) equation coupled with the thermal diffusion equation for a MgB2 SSPD with tf = 10 nm, w = 250 nm, and l = 200 μm:24 (a) temporal scenario deduced from the simulation, (b) position dependence of superfluid density at the locations A, B, C, and D indicated in (a), and (c) comparison of output voltage pulse between the simulation and the experimental observation. The MgB2 SSPD with Lk = 30 nH is connected to a load resistor RL = 50 Ω in parallel. The molecular particle impact was simulated by a power input Qion in the middle of the MgB2 strip. We used the simulation parameters such as a bath temperature of 3.4 K, Ib/Ic = 0.98, and λ = 550 nm that was estimated from the fall time of output pulses. It is considerably longer than 40–100 nm of high quality MgB2 bulks probably because of the c axis oriented film form. Reproduced with permission from Zen et al., Appl. Phys. Lett. 106, 222601 (2015). Copyright 2015 AIP Publishing LLC.
Simulation with the 2D time-dependent Ginzburg–Landau (GL) equation coupled with the thermal diffusion equation for a MgB2 SSPD with tf = 10 nm, w = 250 nm, and l = 200 μm:24 (a) temporal scenario deduced from the simulation, (b) position dependence of superfluid density at the locations A, B, C, and D indicated in (a), and (c) comparison of output voltage pulse between the simulation and the experimental observation. The MgB2 SSPD with Lk = 30 nH is connected to a load resistor RL = 50 Ω in parallel. The molecular particle impact was simulated by a power input Qion in the middle of the MgB2 strip. We used the simulation parameters such as a bath temperature of 3.4 K, Ib/Ic = 0.98, and λ = 550 nm that was estimated from the fall time of output pulses. It is considerably longer than 40–100 nm of high quality MgB2 bulks probably because of the c axis oriented film form. Reproduced with permission from Zen et al., Appl. Phys. Lett. 106, 222601 (2015). Copyright 2015 AIP Publishing LLC.
Similar simulations were performed for an Nb SSPD. Although the material parameters of Nb are significantly different from MgB2 in Table I, the calculation result was similar to Fig. 33. This suggests that the keV particle detection is rather insensitive to material parameters. The time evolution for keV ion detection is predominantly governed by normal domain growth and ETF. Contrarily, it is anticipated that the ∼1-eV photon detection is influenced by the material parameters, strip dimensions, bias currents, temperatures, and substrate types.
The co-existence of the different excitation phenomena explains experimental results on particle energy dependent pulse heights. Semenov et al. reported photon energy measurement in a range between 0.4 and ∼2 eV with a 4-μm square NbN SSPD with tf = 4.5 nm and w = 100 nm.208 The best energy resolution was 0.55 eV for 0.8–1.2 eV. Below ∼0.3 eV and above ∼2 eV, no pulse height variation was observed, which corresponds to either model of dark counts due to vortex motion or the hot spot formation with ETF. In the intermediate energy range, the pulse amplitude almost linearly decreased with increasing the photon energy. Although the photon-energy dependence of pulse height is opposite from the common belief, it implies a transition between the vortex motion and the hot spot formation.
f. SSPD with ultra-wide microstrip
Detection of photons in a telecommunication wavelength normally requires superconductor strips with a thickness of several nm and a width of less than a few 100 nm. Even for keV ions, the empirical upper limit of the strip width is ∼1 μm. However, it has recently been proposed that several-μm wide strips still enable to detect photons near 1 eV. The single photon detection was confirmed by using an NbN SSPD with a thickness of 5.8 nm and widths up to 5.15 μm.206,207 The key factors for the photon detection with the ultrawide strips are a ratio of electron-specific heat capacity to phonon one, which should be larger than 1, and a critical current, which should be larger than 70% of depairing current (critical pair-breaking current).206 Recently, new amorphous superconductor Nb0.15Re0.85 with a thickness of 4 nm and a width of up to 2.5 μm enabled 0.8-eV single photon detection at a low bias current of 21% of departing current, which is common to amorphous superconductors.234 The ultrawide microstrips may overcome the inherent disadvantage, that is, a small sensitive area. Nanostrips are difficult to cover a meaningfully large sensitive area. On the other hand, ultrawide microstrips enable a practically large sensitive area more easily; for example, WSi- and MoSi-SSPDs covering a 1-mm2 area with 8 pixels.235 The ultrawide SSPD development is new trend for practical use.
D. Performance comparison
1. Performance figures
Table III lists the important performance figures to clarify the strengths and weaknesses of the different detector types. Energy resolutions are paramount in the LTD community, employing a 55Fe x-ray source at 5.9 keV. However, direct energy resolution comparison among different detector types is often purposeless, since suitable energy coverage ranges are different from detector to detector. Table III lists the ΔE values followed by photon energies after “@.” Analytical requirements add the second paramount figure that is the pulse fall time τfall, which is related to the photon count rate. When a particle arrives in the period of τfall for a preceding particle, it is called tail pileup that results in energy resolution degradation on pulse height spectra. The second type of pileup is called peak pileup that produces sum peaks. The probability of the second type also depends on τfall. Considering the spatial resolutions and the operating temperatures as well as ΔE and τfall, one can conceive suitable application fields for each detector type.
Superconductor detector types and important performance figures for analytical instruments. The detection category column indicates the basic signal creation principles: quantum detection and thermal detection. The performance figures for photons are the energy resolution ΔE for the best FWHM values at the photon energies in eV after the @ mark, the typical fall time τfall of output pulses, the energy coverage, the typical spatial resolutions, and the typical operating temperatures T.
Detection category . | Detector name . | Acronym . | ΔE (eV) . | τfall (μs) . | Coverage . | Spatial resolution (μm) . | T (K) . |
---|---|---|---|---|---|---|---|
Quantum | Superconductor tunnel junction detector | STJD | 0.1 @ 1 | 2–10 | Infrared— X ray | ∼100 | ∼0.3 |
2.5 @ 250 | |||||||
13 @ 6000 | |||||||
Quantum | Microwave kinetic inductance detector | MKID | 0.3 @ 3 | 2–10 | Infrared— X ray | ∼100 | ∼0.3 |
62 @ 6000 | |||||||
Thermal | Transition edge sensor detector | TESD | 0.07 @ 0.8 | 0.2 100–1000 | Infrared— γ ray | ∼100 | ∼0.05 |
1.5 @ 500 | |||||||
1.6 @ 6000 | |||||||
Thermal | Metallic magnetic calorimetric detector | MMCD | 1.6 @ 6000 | ∼1000 | Infrared— γ ray | ∼100 | ∼0.05 |
Thermal | Superconductor strip particle detector | SSPD | N/A | ∼0.01 | Infrared— X ray | <1 | ∼4–13 |
(0.55 @ 0.8–1.2) |
Detection category . | Detector name . | Acronym . | ΔE (eV) . | τfall (μs) . | Coverage . | Spatial resolution (μm) . | T (K) . |
---|---|---|---|---|---|---|---|
Quantum | Superconductor tunnel junction detector | STJD | 0.1 @ 1 | 2–10 | Infrared— X ray | ∼100 | ∼0.3 |
2.5 @ 250 | |||||||
13 @ 6000 | |||||||
Quantum | Microwave kinetic inductance detector | MKID | 0.3 @ 3 | 2–10 | Infrared— X ray | ∼100 | ∼0.3 |
62 @ 6000 | |||||||
Thermal | Transition edge sensor detector | TESD | 0.07 @ 0.8 | 0.2 100–1000 | Infrared— γ ray | ∼100 | ∼0.05 |
1.5 @ 500 | |||||||
1.6 @ 6000 | |||||||
Thermal | Metallic magnetic calorimetric detector | MMCD | 1.6 @ 6000 | ∼1000 | Infrared— γ ray | ∼100 | ∼0.05 |
Thermal | Superconductor strip particle detector | SSPD | N/A | ∼0.01 | Infrared— X ray | <1 | ∼4–13 |
(0.55 @ 0.8–1.2) |
2. Energy resolution
The ΔE values of the superconductor detectors including an electrical noise are typically more than 10 times better than that of semiconductor detectors in a wide photon energy range. EDS of single photons is possible from infrared light to γ rays. The best E0/ΔE values are higher than 3000 by the thermal detection with TESDs and MMCDs.167,175 They achieved a superior energy resolution of better than 2 eV at 6 keV, although the best performance is obtained at a count rate of ∼10 cps/pixel. At this energy resolution, it is possible to measure characteristic x-ray line shapes, which is called XES that is normally performed with WDS. In principle, ΔE in thermal detection is independent of E0 according to Eq. (6): almost the same ΔE value of 1.5 eV was obtained at 500 eV.150,151
The quantum detection with STJDs and MKIDs can be also applied to the range of ∼6 keV; however, ΔE values are 10 times worse than that of the thermal detection: 29 eV of the Nb/Al-STJD developed at LLNL.191 Moreover, because of the thin-film nature, x-ray detection efficiencies are only several %, which precludes practical use for analytical instruments. To solve this problem, Kurakado and Taniguchi developed the x-ray imaging detectors with a series-STJ-array attached to a thick sapphire crystal of 5 × 5 × 0.4 mm3 or a Si substrate of 3 × 3 × 0.4 mm3 as x-ray absorbers.236 The ΔE value of the Si-STJ detector was 63.5 eV at 5.9 keV, which is not as good as that of superconductor thermal detectors, but approximately half of ΔE in semiconductor detectors. Furthermore, Nakamura et al. developed a neutron detector combining STJs and a large Li2B4O7 absorber, which includes the isotopes of 6Li and 10B having a large cross section for neutrons.237 The absorber/STJ combination utilizes phonon-mediated signal generation. The STJD or MKID direct quantum detection performs well in a low energy range below ∼2 keV because of a high spectral sensitivity and ΔE proportional to . The best STJD achieved a ΔE value of 2.5 eV at 250 eV at LLNL.77 The records at AIST were 4.1 eV at 400 eV and ∼200 kcps in this article.
Both STJ and TES types enable photon-number-resolving measurement through pulse height analysis. Recently, photon-number-resolving detection has also been realized by a semiconductor CMOS camera that has 9.4 M pixels with 4.6 μm covering a sensitive area of 18.814 × 10.598 mm2.238 An ultra-low noise readout with 0.27 electrons and a low dark current of 0.006 electrons/pixel/s at −35 °C enabled the photon number discrimination up to ∼200 photons at quantum efficiencies above 30% between 300 and 900 nm. The 9.4-mega pixel imaging camera outperforms small superconductor array detectors at a cryogenic temperature. This may be one of the rare cases that semiconductor detectors currently leave superconductor detectors behind in photon-number resolving imaging. Further improvement in the number of pixels is required to beyond the photon-number-resolving CMOS imager. MKID imagers are advantageous for a range above 900 nm with single photon energy measurement.
In addition to the above-mentioned particles, superconductor detectors can detect electrons, ions, and macromolecules, although a fraction of the energy transfer to an absorber is ∼70%. This low percentage causes a large statistical fluctuation and an E0/ΔE value of ∼10. However, the superconductor detectors can measure the kinetic energies of low-energy particles in a keV range. In general, semiconductor detectors can measure particle energies only in an MeV range.
3. Fall time and count rate
The fall time τfall of output pulses is crucial for analytical instruments that often require a high count rate for high throughput analysis or imaging analysis for a reasonable acquisition time. When one detector has a 100 time shorter τfall, the measurement time can be shortened from 100 to 1h, for example. There are two models on the relation between the incoming count rate n and the recorded count rate m: nonparalyzable model and paralyzable one. For low count rates in , both models can be approximated by . This relation gives a proper estimation for counting devices such as SSPDs. At n = 10 M photons/s, the typical SSPDs enable to count 90% of the incoming photons. This estimation method can be also applied to the superconductor EDS detectors. The incoming photon fluxes at the 90-% counting rate are 10 k photons/s for an STJD with τfall = 10 μs and 100 photons/s for a TESD with τfall = 1 ms. An experimental result on a TESD was ∼20 photons/s at 90%.151 For single-particle spectroscopy, we must rely on a corner of the ΔE degradation against n rather than the simple counting capability. The experimental results were ∼200 kcps for the 100-pixle STJD in Fig. 22, and ∼2 kcps for the 240-pixel TESD.151 The STJDs derive an advantage from a short τfall. On the other hand, for applications to rare events such as dark matter search, a count rate of a few cps with TES or MMC array detectors is enough in most cases. These fields require a high-energy resolution with the thermal detectors rather than a high count rate of quantum detectors.
SSPDs have no energy resolution, but they outperform conventional semiconductor APDs in photon counting with τfall < 1 ns, a low dark count rate, and a small jitter without afterpulses. The jitter can be as short as 4.3 ps compared with 50 ps of an InGaAs detector for near-infrared photons, although it was recorded with a very small NbN SSPD with a width of 80 nm and a length of 5 μm.239 The ultrawide SSPD will solve the problem. SSPDs are beginning to play an active part in the fields of quantum communication, TOF ranging, and quantum computing as well as analytical instrumentation such as MS and correlation spectroscopy.182,240,241
4. Energy coverage
The energy coverage specifies spectroscopic types and samples to be analyzed: for example, a soft x-ray range below 1 keV for light element dopant analysis in compound semiconductors and a γ-ray range for nuclear material analysis. Measurements in an energy range lower than infrared rays include cosmic microwave background (CMB), neutrino mass, and optical astronomy, although they are not in the category of analytical instruments. The superconductor quantum detectors with a high count rate are suitable for a low energy range below a few keV because of the basic relation of . On the other hand, the superconductor thermal detectors perform well even in a high-energy range like γ rays at an almost constant ΔE independent of E0. Absorber designs are adjusted to certain energy ranges.
Both quantum and thermal detectors enable EDS in an infrared to visible range, where conventional photon detectors perform photon counting or photon flux measurement only. The EDS range of semiconductor detectors is above ∼50 eV, whereas superconductor detectors cover down to ∼0.6 eV. It signifies that the boundary between photon and wave detection schemes was lowered by a factor of ∼100 compared with semiconductors. This EDS is achievable only with superconductor detectors. The types of STJ,79 MKI,144 and TES81 typically have a ΔE/E value of 0.1 for infrared-visible photons. This may be promising for imaging multiple fluorescent markers in biological samples242–245 and time-correlated single photon counting (TCSPC).246 For energy-sensitive single-photon 2D imaging, MKID imagers are conceivably most promising.148
For ions or macromolecule detection, the superconductor quantum detectors cover up to ∼30 keV with a reasonable energy linearity, as shown in Fig. 11. Superconductor thermal detectors are expected to perform well up to an MeV range by designing a proper absorber for high-energy particle physics.
5. Spatial resolution and arraying
The spatial resolution of the superconductor array detectors is normally worse than semiconductor imaging sensors such as CCD and CMOS cameras. The semiconductor cameras have more than 1 M pixels with a typical spatial resolution of several μm, which is limited by a photodiode size. On the other hand, a typical pixel size of the superconductor detectors is ∼100 μm to meet a reasonable sensitive area, which is ten times worse than the semiconductor one. Although we can fabricate pixels of a few 10 μm square or smaller, readout circuitry for superconductor pixels is troublesome: the individual parallel readout scheme for STJDs has a channel number limit of ∼1000 or the multiplexing scheme for TESDs has a limit of ∼10 000. MKDI may potentially compete with CCD and CMOS. At this moment, superconductor mega-pixel cameras comparable to CCD and CMOS have not been realized.
Practical use of superconductors for imaging beyond the semiconductor limit is possible with SSPDs. The spatial resolution for x rays is simply determined by a line-and-space pitch of superconductor nanostrips. Several 10-nm resolutions beyond a-few-μm of semiconductors are possible in principle. Practically, considering a sensitive area, we proposed an x-ray 1D imager with a realistic spatial resolution of 500 nm for resonant inelastic x-ray scattering (RIXS) at synchrotron radiation facilities. The proof-of-concept of delay line operation was successful.247 SSPD 2D-imaging is also pursued with a single nanostrip line format,248 a thermally coupled two-layer stack format,249 and row-column multiplexing readout of 1024-SSPD-pixels,250 and 1024 pixels with an orthogonal time-amplitude multiplexing.251 The largest SSPD imager was recently reported: 0.4 M pixels with a sensitive area of 4 × 2.5 mm2 and a spatial resolution of 5 μm,252 which is approaching the photon-number-resolving CMOS camera of 9.4 M pixels with 4.6 μm and 8.814 × 10.598 mm2. The SSPD 1D or 2D imagers are still considerably smaller than CCD or CMOS cameras but may outperform them in future.
The simple array configuration of pixels has typically a spatial resolution of 100 μm. To exceed this, another route is to utilize quasiparticle diffusion in a long superconductor absorber strip and readout STJs at both ends, based on the quasiparticle trapping scheme,87,88 as mentioned in Sec. III B 3 e. The achieved spatial resolutions were of 5 μm for a 150 × 450-μm2 1D absorber,87 ∼10 μm for a 100 × 200-μm2 1D absorber,120,253 ∼10 μm for a 31.5 × 125-μm 1D absorber,254 and 35 μm for 3 × 20 2D array of 33.5 × 360 μm2 absorbers.255 However, the largest sensitive area of the STJ imaging detectors was merely ∼7 mm2, which is still extremely smaller than a-few-cm square size of the scientific-application-oriented CCD and CMOS cameras. Therefore, the STJD imaging scheme using quasiparticle diffusion has not been advanced beyond the proof-of-concept.
Unlike astronomical observations, analytical instruments often require no imaging capability. Imaging is realized by scanning a probe beam. For this type of imaging, arraying TES, MMC, or STJ pixels is to achieve a large sensitive area and a high count rate, keeping a high-energy resolution. We built STJD spectrometers with 100-to-200 pixels, 6-eV resolution, and ∼200 kcps at AIST. Concerning other institutes, Friedrich et al. built a 112-pixel Ta/Al-based STJD system with ∼8-eV resolution at 525 eV and a few 100 kcps.256 They also reported the 2.5-eV resolution at 250 eV for the best pixel.77 Lee et al. built a 240-pixel Mo/Cu-TESD system with the 1.5-eV resolution at 500 eV and ∼2 kcps.151
6. Cryogenics and other aspects
The operating temperatures in Table III require different cryogenic systems: 3He sorption type for ∼0.3 K, ADR or 3He–4He dilution type for 0.05 K, and Gifford–McMahon (GM) type or pulse tube type for ∼3 K. In addition, thermal detectors require a high temperature stability in a μK range to keep a peak shift tolerable; base temperature variation directly influences pulse heights. On the other hand, quantum detectors operate without peak shift as long as they are cooled below ∼0.1Tc; no temperature regulation is necessary. With respect to the costs and sizes, we have an interest in palm-sized coolers, although the temperature range of the micromachined coolers is currently 20–150 K in general.257 In future, miniaturized coolers may be utilized like a Peltier cooler for semiconductor detectors.
It is not listed in Table III, but other important factors for analytical instruments are real time signal processing for quick adjustment of measurement conditions, stability for a long run, and operability for the users. Considering all the factors discussed in this section, one can select a superconductor from Table I and a proper superconductor detector type from Table III for each application, and then build a complete innovative spectroscopic system including a cryostat and a readout circuit. A detailed consideration for 3D Atom probe is given elsewhere.50 Successful examples at ETL-AIST will be described in Sec. V after explaining engineering advances in Sec. IV.
IV. ADVANCES FOR PRACTICAL USE
A. General
In foregoing chapters, we reviewed the pioneering ideas ahead of times at early years, the lessons from semiconductor detectors, modern superconductor detector physics, analogies and differences among different superconductor detector types, and detector performance figures. This chapter describes engineering advances mostly at ETL-AIST, especially for STJ and SS types for analytical science. Both STJ and SS types have an advantage in fast response over other types. To realize a practical sensitive area, the STJ type requires multichannel parallel readout. The SS type requires a novel nanostrip configuration design with a single channel readout.
B. Superconductor tunnel junction detector (STJD)
1. Fabrication of STJ array
The first-generation array detectors had both junction pixels and wiring leads on the same Si wafer surface as shown in Fig. 34(a). The wiring leads run along the gaps of adjacent pixels. The close-packed square arrangement was possible up to ∼16 pixels before 2000. The chip mount is shown in (b). Beyond 16 pixels, 100 pixels were arranged in a horizontally long format as shown in (c) such that the wiring leads can reach each pixel without overlap with STJ pixels. When we make a square arrangement for over 100 pixels, the gap between adjacent junctions must be unacceptably large. Alternatively, wiring leads can run on junction pixels after inserting an insulation layer.258 This overlap between junctions and wiring leads considerably reduces a sensitive area. Even in this case, optical photon detection can be performed through a sapphire substrate from the back side.78,79 However, the back illumination is unsuitable for x rays or molecules.
First-generation design (2D configuration) for Nb/Al junction array detectors: (a) 4 × 4 array of the 200-μm square junctions before 2000, (b) 5 × 5 mm2 chip on the carrier, and (c) horizontally arranged 100-pixels in 2005. The red arrow in (b) indicates the location of the16 pixels in (a). In this 2D arrangement, the Nb wiring leads run the gaps between adjacent pixels. The horizontal arrangement in (c) is adapted to an ion beam shape of ESI TOF MS in Fig. 54.
First-generation design (2D configuration) for Nb/Al junction array detectors: (a) 4 × 4 array of the 200-μm square junctions before 2000, (b) 5 × 5 mm2 chip on the carrier, and (c) horizontally arranged 100-pixels in 2005. The red arrow in (b) indicates the location of the16 pixels in (a). In this 2D arrangement, the Nb wiring leads run the gaps between adjacent pixels. The horizontal arrangement in (c) is adapted to an ion beam shape of ESI TOF MS in Fig. 54.
Figure 35 shows the second-generation close-packed square arrangement of 100 pixels fabricated at a clean-room named CRAVITY that was established by Ukibe et al. in 2013.93 The energy resolution of the 200-μm square junction was in a range of 10–20 eV at 400 eV, which is considerably worse than 4.1 eV of the 100-μm square junction primarily because of the spatial nonuniformity. Therefore, we use the junction size of 100 μm for high resolution XRS. The gap between the pixels is 20 μm: the filling factor is 70%, which is considerably higher than 17.5% for the first-generation design in Fig. 34(c). The high filling factor was realized by the 3D configuration using chemical mechanical polishing (CMP). After fabricating the embedded Nb wiring lead layer, tunnel junctions were fabricated on the planarized flat surface of the SiO2 insulation layer as shown in Fig. 35(a). However, the sensitive area is still 25 times smaller than the 25-mm2 SDD. To compensate this, a polycapillary focusing lens compensates the sensitive area difference.259 In addition, the fabrication of 1024 pixels is practically possible.141 The 100-μm version with high-energy resolution is utilized for XRS with synchrotron radiation at KEK PF and SEM at AIST. An application to particle induced x-ray emission (PIXE) analysis at the University of Tsukuba is approaching to a stage of routine use.260
Second-generation design (3D configuration) for the 100-μm square Nb/Al junction array detector with 100 pixels in 2015:93 (a) optical microscope image and cross-sectional structure and (b) photograph of the chip carrier. The 100-pixel STJD was fabricated at the AIST-CRA VITY facility. The close-packed pixel arrangement was realized by using the embedded wiring leads running underneath the junction layer separated by the SiO2 layer. Reproduced with permission from Fujii et al., Supercond. Sci. Technol. 28, 104005 (2015). Copyright 2015 IOP Publishing.
Second-generation design (3D configuration) for the 100-μm square Nb/Al junction array detector with 100 pixels in 2015:93 (a) optical microscope image and cross-sectional structure and (b) photograph of the chip carrier. The 100-pixel STJD was fabricated at the AIST-CRA VITY facility. The close-packed pixel arrangement was realized by using the embedded wiring leads running underneath the junction layer separated by the SiO2 layer. Reproduced with permission from Fujii et al., Supercond. Sci. Technol. 28, 104005 (2015). Copyright 2015 IOP Publishing.
For close-packed arrangement of STJ pixels, junction shapes are important. A wide range of junction shapes such as diamond, sine, quartic, and normal function were investigated to efficiently suppress the DC Josephson effect and the Fiske resonance in a low magnetic field.261–263 A diamond shape is effective in reducing a magnetic field strength.262 The complete sidelobe suppression of a Fraunhofer pattern was realized by the normal function shape: a small magnetic field of 2Φ0 was enough.263 The reduction of a necessary magnetic field strength led to skipping an external large magnet coil and using an integrated coil underneath a junction layer.264 Furthermore, with an annular junction shape, one single vortex trapping suppressed the Josephson effect without an external magnetic coil.265 Nonetheless, considering an area of a 100-pixel STJD, we currently adopted a simple square shape and an external magnet coil as shown in Fig. 37 later. The square shape is favorable to filling a space as much as possible. Applying a magnetic field along the diagonal direction imitates the diamond shape. Integration or omission of a magnet coil is a future challenge.
Cryogenics and electronics for the STJD instrumentation before 2005: (a) 3He sorption cryostat (Infrared Lab. Inc., HDL-8) cooled by liquid N2 and 4He, (b) 4-K shield and 3He pod for sorption-pumping, which is overturned to mount the STJD, (c) STJD chip mounted on the 0.3-K stage inside the 4-K shield, (d) charge-sensitive preamplifier (Canberra 2003T), and (e) analog shaping amplifier (Canberra 2020), analog-to-digital converter (ADC) (Canberra 8706), and timing modules for TOF MS (Ortec 551).
Cryogenics and electronics for the STJD instrumentation before 2005: (a) 3He sorption cryostat (Infrared Lab. Inc., HDL-8) cooled by liquid N2 and 4He, (b) 4-K shield and 3He pod for sorption-pumping, which is overturned to mount the STJD, (c) STJD chip mounted on the 0.3-K stage inside the 4-K shield, (d) charge-sensitive preamplifier (Canberra 2003T), and (e) analog shaping amplifier (Canberra 2020), analog-to-digital converter (ADC) (Canberra 8706), and timing modules for TOF MS (Ortec 551).
Superconductor-tunnel-junction equipped x-ray-absorption-fine-structure instrument (SC-XAFS) with modern cryogenics and digital signal processing (DSP) electronics after 2009:273 (a) fully automated close-cycle 3He sorption refrigerator with the pulse tube cooler (NIKI GLASS, LTS-HE3-LV) at KEK PF synchrotron radiation beam line, (b) cutaway cold finger (VACFIELD, cryofinger), (c) preamplifier array, (d) field-programmable-gate-array based digital signal processing (FPGA-DSP) circuits for deriving timing-energy data from output pulses (Morita E-design LLC., DTEMP), and (e) DSP module stack for 100-channel parallel readout.
Superconductor-tunnel-junction equipped x-ray-absorption-fine-structure instrument (SC-XAFS) with modern cryogenics and digital signal processing (DSP) electronics after 2009:273 (a) fully automated close-cycle 3He sorption refrigerator with the pulse tube cooler (NIKI GLASS, LTS-HE3-LV) at KEK PF synchrotron radiation beam line, (b) cutaway cold finger (VACFIELD, cryofinger), (c) preamplifier array, (d) field-programmable-gate-array based digital signal processing (FPGA-DSP) circuits for deriving timing-energy data from output pulses (Morita E-design LLC., DTEMP), and (e) DSP module stack for 100-channel parallel readout.
2. Cryogenics and readout electronics
Figure 36(a) shows the cryogenic technology before 2005 at AIST. The 3He sorption fringe (Infrared Lab. Inc., HDL-8) in (a) required liquid nitrogen and helium: it took over 6h to liquify 3He by pumping 4He. The cooling process was not automated: we had to watch the 3He liquifying process during pumping. Kraus stated in 2002 that “superconductor detectors were promising for the next generation instrumentation, but liquid cryogens precluded the superconductor detectors from being adopted early by the potential users.”266 In those days, liquid-cryogen-free coolers were under development, and no commercial products were available.
Figure 36(c) shows the STJD chip mounted on the 3He cold stage inside the 4-K shield. The laser beam for alignment illuminates the vertically mounded STJD. Only one junction on the detector chip was connected to the discrete charge-sensitive preamplifier (Canberra 2003T) in (d). The Gaussian shaping amplifier (Canberra 2020) received output pulses and then the Wilkinson-type analog-digital converter (ADC) (Canberra 8706) in (e) digitized pulse heights. A PC-based multichannel analyzer (MCA) recorded x-ray spectra. With this instrumentation, we attempted to measure our first x-ray absorption spectroscopy (XAS) of a Hf-Al oxide film for high-k gate insulation in semiconductors by using a single 100-μm square junction.267 The sample was set on the 4-K stage near the STJD to realize a practical solid angle. We successfully acquired XAS spectra of the oxygen K-edge; however, for sample exchange, we had to repeat the whole sequence of cooling from room temperature to 0.3 K. For MS, the timing single channel analyzer (Ortec 551) and a digital oscilloscope replaced the readout modules to acquire the TOF and kinetic energy values of ions.
In 2005, the cooling procedure was improved by a close-cycle 3He-sorption refrigerator (Janis Research LLC, SVCCF He-3) with a mechanical Gifford-McMahon (GM) cryocooler (SHI Cryogenics Group). This was the trigger for developing a series of practical analytical instruments with a 100-pixel array STJD. Considering crosstalk and bandwidth for timing measurement in TOF MS, we adopted a combination of Cu coaxial cables (Coax Co. Ltd.) inside the cold finger at 0.3 K, CuNi between 0.3 and 3 K, and brass between 3 K and room temperature.268 The base temperature was 0.29 K for a one-shot holding time of 6.5 days after the installation of 100-channel cables. It is consistent with a measured heat input of 5.7 μW through the cables in addition to 100 μW due to blackbody radiation.269
The first fully automated close-cycle 3He refrigerator (NIKI GLASS, LTS-HE3-LV) with a pulse tube cooler (Cryomech Inc.) was installed in 2006. A holding time of 32h at 0.35 K was shorter than the above-mentioned refrigerator; nevertheless, the fully automatic control including heat switches compensated the short one-shot holding time.270 Up to now, we have installed several refrigerators with NbTi twisted-pair leads or NbTi coaxial cables.271,272 Figure 37(a) shows the superconductor-tunnel-junction equipped XAFS (SC-XAFS) instrument installed at a synchrotron radiation beam line in KEK PF.273 In (b), the 100-pixle Nb/Al-STJD is mounted at the end of the cutaway cold finger (VACFIELD, cryofinger) with the split superconductor magnet coil. Aluminized parylene thermal windows were placed on stages at 3 and 50 K to reduce blackbody radiation. The output pulses were amplified by the charge-sensitive preamplifier module in (c). The pulse leading-edge timing and pulse height were determined by the field-programmable-gate-array-based digital signal processing modules (FPGA-DSP) (Morita E-design LLC., Digital Timing Energy Multiprocessor: DTEMP) in (d) and (e). The SC-XAFS instrument has recently been upgraded to 200 pixels. Nowadays, the users of the analytical instruments with superconductor detectors need not pay attention to cryogenic cooling procedure. The details of the data processing will be reported elsewhere.50
C. Superconductor strip particle detector (SSPD)
1. Fabrication of SS array
The advantages of SSPDs over conventional semiconductor APDs include a short dead time and a low dark count rate. Furthermore, the delay line operation of superconductor strips is advantageous to x-ray photon imaging at a sub-μm spatial resolution beyond x-ray CMOS cameras.247 However, the inherent nano-size is an SSPD weakness to realize a practically large sensitive area comparable to that of semiconductor detectors. When we design an SSPD with τfall = 1 ns, a meander NbN strip should have the dimensions of tf = 5 nm, w = 300 nm, and l = 1.5 mm according to Eqs. (25) and (26), which corresponds to a sensitive area of only 0.00045 mm2. For comparison, semiconductor InGaAs PIN photodiodes for telecommunication or near-infrared photometry have a diameter between 0.3 and 3 mm ϕ.274 The latest CMOS imaging camera with the photon number resolving capability has a sensitive area of 19 × 11 mm2.238 Microchannel plate (MCP) ion detectors for MS have a typical diameter between 14.5 and 42 mm ϕ.275 A sensitive area of the simple meander geometry is many orders smaller than that of MCPs. This precludes a practical application to MS and other fields.
The SSPD design goals vary depending on applications such as telecommunication, x-ray imaging, and MS. We focus on the goal for MS in this section. The diameter of molecule ion beams in TOF MS is about 1 cm; therefore, the size of 1 cm2 is our target, keeping a pulse width of ∼1 ns. An increase in nanostrip length to cover a practical sensitive area causes an intolerable increase in kinetic inductance. For example, when a 200-nm wide NbN nanostrip with a thickness of 7 nm and a line-and-space pitch of 400 nm covers a small sensitive area of 1 mm2 at a filling factor of 50%, the nanostrip length becomes 2.5 m. The kinetic inductance of this nanostrip is evaluated at 440 μH, of which τrise and τfall values are 200 ns and 9 μs, respectively, when we assume typical values of Rk = 2 kΩ and RL = 50 Ω for a readout circuit. This extremely slow response is unacceptable for high-mass-resolution TOF MS that requires a few ns response at least. It is necessary to reduce Lk by a factor of hundreds while keeping a large sensitive area. As of 2009, for molecule detection, the largest size of meander-shaped NbN-SSPDs was 200 × 200 μm2: τrise = 900 ps and τfall = 27 ns.276 A larger 500-μm square Nb-SSPD latched in a normal state after one molecule impact event unfortunately.277
At the beginning, it seemed to be impossible that we can cover ∼1 cm2. The 32 × 32 row-column multiplexing readout developed by Wollman et al. marked an average jitter of 250 ps that is enough for TOF MS;250 however, the size of the SSPD array is still as small as 1.6 × 1.6 mm2. Another simple method to realize a large area is to connect nanostrips in parallel, reducing total inductance by Lk/n, where n is the number of parallel strips.278,279 We investigated two operation modes for the parallel configuration: “cascade regime” by Consiglio Nazionale Delle Ricerche (CNR) and ‘single-strip-switch regime ‘ under the collaboration of AIST-CNR.241
The first parallel SSPD configuration was intended to induce cascade S-N switching of all parallel nanostrips such that finite resistance appears. The proof-of-concept cascade SSPD had five parallel lines of 130-nm wide NbN strips covering 3.25 μm2.278 The operation of this parallel configuration is called cascade regime.241 One drawback of the cascade regime is that a series inductance LS, which must be slightly larger than Lk of each nanostrip, is required to induce cascade switching. The series LS limits the response time to τfall = 12.5 ns.278 The design concept in the cascade regime was extended to the series-parallel combination that is a series connection of parallel blocks.280
Another parallel configuration that is effective in increasing the sensitive area further without impairing fast response is the single-strip-switch regime.241 This is the simple parallel connection of nanostrips without a series inductor for cascade switching. At the beginning, we predicted that no measurable output pulses would appear, since the resistance of the superconducting strips in parallel was zero. However, we observed measurable output pulses from a 200-μm square Nb SSPD with a 39 series connection of five parallel strip blocks: τrise = 400 ps, τfall = 500 ps and an FWHM pulse width of 700 ps.281 A circuit simulation has revealed that even when parallel strips are in a superconducting state, small but measurable voltage pulses emerge across a load resistor because of the kinetic inductance of each strip.282 The parallel 200-μm square Nb-SSPD operated in the single-strip-switch regime successfully produced output pulses of ∼1 mV, which is higher than a noise level of 0.1 mV, after amplifying at a gain of 1000 for 17.5-keV biomolecule ions.281 The 200-μm square parallel SSPD recorded τrise = 400 ps,τfall = 500 ps, and a FWHM pulse width of 700 ps. The 700-ps pulse width is the same as the latest MCP (Hamamatsu, F9892-31/-32) with a large sensitive area of 1385 mm2; however, the SSPD sensitive area of 0.04 mm2 is still enormously small. Nevertheless, SSPDs for molecule detection have the advantage of molecular-mass-independent detection efficiency in contrast to MCPs, of which detection efficiency decreases as molecular mass increases and becomes almost zero for large protein molecules over ∼100 kDa in general.283–286 Therefore, when we can fabricate a ∼1-mm2 parallel SSPD, it can overcome the MCP detection efficiency in a mass range over ∼100 kDa.
1-mm square Nb-SSPD with series-parallel configuration for time-of-flight mass spectrometry (TOF MS):282 (a) laser scanning microscope image of 101-series connection of five-parallel blocks by skipping the middle area indicated by the dotted lines, (b) output pulse at a bias current of 0.54Ic for the five-parallel SSPD in (a) at 4.2 K, and (c) output pulse at a bias current of 0.76Ic for 51-series connection of 10-parallel blocks. Each Nb strip in the blocks has tf = 40 nm, w = 1 μm, and l = 1 mm. The five-parallel SSPD in (a) had τrise = 600 ps and τfall = 1.8 ns. The ten-parallel SSPD exhibited the better performance:τrise = 520 ps and τfall = 920 ps. The TOF difference of 75 ns between (b) and (c) around 22.3–22.4 μs arises from a mass distribution of the incoming angiotensin-I molecules. Reproduced with permission from Zen et al., Appl. Phys. Lett. 95, 172508 (2009). Copyright 2009 AIP Publishing LLC.
1-mm square Nb-SSPD with series-parallel configuration for time-of-flight mass spectrometry (TOF MS):282 (a) laser scanning microscope image of 101-series connection of five-parallel blocks by skipping the middle area indicated by the dotted lines, (b) output pulse at a bias current of 0.54Ic for the five-parallel SSPD in (a) at 4.2 K, and (c) output pulse at a bias current of 0.76Ic for 51-series connection of 10-parallel blocks. Each Nb strip in the blocks has tf = 40 nm, w = 1 μm, and l = 1 mm. The five-parallel SSPD in (a) had τrise = 600 ps and τfall = 1.8 ns. The ten-parallel SSPD exhibited the better performance:τrise = 520 ps and τfall = 920 ps. The TOF difference of 75 ns between (b) and (c) around 22.3–22.4 μs arises from a mass distribution of the incoming angiotensin-I molecules. Reproduced with permission from Zen et al., Appl. Phys. Lett. 95, 172508 (2009). Copyright 2009 AIP Publishing LLC.
2. Operation in parallel configuration
The output pulse height of meander SSPDs should be constant according to the hot spot model with ETF Joule self-heating. In fact, the pulse-height distribution of a meander NbN-SSPD was extremely narrow with ΔE/E0 = 0.04 for the particle impact of Ar, sinapinic acid, and BSA.287 On the other hand, in the parallel SSPDs, we observed that pulse heights significantly varied from event to event in a range of 20–80 mV for the 1-mm2 Nb SSPD. This suggests a nonuniform distribution of bias supercurrents among the strips in parallel.282 The operation of the parallel configuration is intrinsically influenced by the nature of superconductivity, based on the London phenomenological theory. In addition, we observed the dynamic change of the bias-supercurrent redistribution in the course of molecule detection.288,289
The details of the operation of the parallel SSPDs will be described elsewhere.50 Here, the dynamic change and the improvement are briefly described. Figure 39 shows a circuit simulation result with LTspice (Linear Technology) for an SSPD with ten parallel connections of 202-mm long strips.288 All the switches in the equivalent circuit in (a) are normally on, shunting the resistors that simulate the effective constant resistance Rk of 160 Ω after particle incidence. The dynamic change of current I1 flowing through one of the strips in parallel is displayed in (b) for two cases of Rb = 0 and 1 Ω. The initial value of I1, I2, I3, etc. was assumed to be 3 mA, and other circuit parameters appear in Ref. 288. The time evolution in red for Rb = 0 represents the experimental condition of the parallel strips. The I1 value abruptly decreases when the switch of the I1 circuit flips off for 3 ns, which simulates a lysozyme incident event in the I1 strip of interest. Events in other parallel strips at random induce the small stepwise bias-current recovery. The output currents flowing RL exhibit the red distribution in Fig. 39(c). The red histogram in Fig. 39(c) well reproduces the experimental data. By setting a proper lower-level discrimination (LLD) for pulse height or IL in (c), it is possible to collect almost all events. However, the broad pulse heights are undesirable.
Circuit simulation of dynamic time evolution for the parallel SSPD in the single-strip-switch regime:288 (a) equivalent circuit, (b) LTspice simulations for the specific strip (I1) among the 10-paralle strips, and (c) histograms of the current IL passing through the 50-Ω load resistor RL. The switches are normally on. Particle impact events are simulated by flipping off one of the switches at random, adding Rk (=160 Ω) to one of the 10 strips for 3 ns. The dynamic changes and the histograms are depicted by red for Rb = 0 Ω and by blue for Rb = 1 Ω in (b) and (c). The red histogram in (c) reproduces the experimental output-charge distribution. Inserting non-zero Rb stabilizes the circuit, generating an almost constant pulse height. Reproduced with permission from Zen et al., Appl. Phys. Lett. 104, 012601 (2014). Copyright 2014 AIP Publishing LLC.
Circuit simulation of dynamic time evolution for the parallel SSPD in the single-strip-switch regime:288 (a) equivalent circuit, (b) LTspice simulations for the specific strip (I1) among the 10-paralle strips, and (c) histograms of the current IL passing through the 50-Ω load resistor RL. The switches are normally on. Particle impact events are simulated by flipping off one of the switches at random, adding Rk (=160 Ω) to one of the 10 strips for 3 ns. The dynamic changes and the histograms are depicted by red for Rb = 0 Ω and by blue for Rb = 1 Ω in (b) and (c). The red histogram in (c) reproduces the experimental output-charge distribution. Inserting non-zero Rb stabilizes the circuit, generating an almost constant pulse height. Reproduced with permission from Zen et al., Appl. Phys. Lett. 104, 012601 (2014). Copyright 2014 AIP Publishing LLC.
When the Rb value is switched to 1 Ω, we obtained the blue time evolution in Fig. 39(b) and the blue histogram in (c), demonstrating that the 1-Ω series resistor for each parallel strip is effective to produce the almost constant pulse height like a simple meander geometry. Therefore, it is possible to realize the narrow output pulse height distribution by inserting a small bias resister in the parallel configuration.
The circuit simulation result of the blue histogram in Fig. 39(c) was experimentally demonstrated by using Al bonding-wires that function as series bias resistors.290 In addition to the discrete Al bonding-wire resistors, non-superconductor resistors were integrated in a detector chip before our simulation. One example was to use discrete Ti:Au resistors inserted between 10-μm long NbN strips and a gold coplanar waveguide, although they did not intend to avoid a broad pulse-height distribution but to suppress afterpulses.279 The two discrete designs require a large space for bonding or an additional area for non-superconductor materials.
Casaburi et al. conceived more ingenious series-resistor design without the need for additional fabrication processes with different materials: it was to utilize a superconductor as a resistor, that is, a “superconductor bias resistor.”291, Figure 40 shows the SSPD design that each 1-μm wide Nb meander strip is connected in series to a narrower Nb resistor strip with a width of 200 nm and a length of 0.8 μm via a large-square thermal buffer.291 The thickness of the Nb film for all components was 40 nm. When a bias current exceeds Ic of the Nb resistor strips but it is still lower than Ic of the Nb detector strips, Joule self-heating sustains a resistive state of 29 Ω at 4.2 K. A confocal microscope experiment demonstrated that the initial bias currents denoted by the blue squares in Fig. 40(c) were almost constant. Vortices can enter into and exit from the loops of the adjacent Nb detector strips through the resistive Nb strips; therefore, the bias current distribution is flat even after detecting particles. In contrast, the red simulation curve in Fig. 40(c) for a parallel SSPD without series resistors shows a parabolic-shape bias-current distribution. With this superconductor resistor concept, a 1-cm square SSPD with a proper series-parallel design and a superconductor resistor integration is promising for MS covering a ultrawide MW range at a quantum efficiency of 100% and a pulse width of ∼1 ns.
Parallel Nb-SSPD with the series superconductor resistor integration:291 (a) optical image of a part of the 11-parallel and 820-μm long Nb detector strips and (b) SEM image of a 0.8-μm long resistive Nb strip, and (c) measured bias current distribution (blue squares) for the detector strips in parallel and simulated current distribution without series resistors (red closed circles connected by solid line). The meander Nb detector strips and Nb resistor strips have a thickness of 40 nm and widths of 1 μm and 200 nm, respectively. A Nb resistor strip in (b) functions as a 29-Ω resistor at 4.2 K under a bias current above Ic. Each meander strip is connected to a Nb resistor strip via a large-square Nb thermal-buffer area of 20 × 20 μm2. In (c), the measured currents in blue are almost constant, whereas the simulated data in red without series resistors exhibit the parabolic curve. Reproduced with permission from Casaburi et al., Supercond. Sci. Technol. 31, 06LT01 (2018). Copyright 2018 IOP Publishing.
Parallel Nb-SSPD with the series superconductor resistor integration:291 (a) optical image of a part of the 11-parallel and 820-μm long Nb detector strips and (b) SEM image of a 0.8-μm long resistive Nb strip, and (c) measured bias current distribution (blue squares) for the detector strips in parallel and simulated current distribution without series resistors (red closed circles connected by solid line). The meander Nb detector strips and Nb resistor strips have a thickness of 40 nm and widths of 1 μm and 200 nm, respectively. A Nb resistor strip in (b) functions as a 29-Ω resistor at 4.2 K under a bias current above Ic. Each meander strip is connected to a Nb resistor strip via a large-square Nb thermal-buffer area of 20 × 20 μm2. In (c), the measured currents in blue are almost constant, whereas the simulated data in red without series resistors exhibit the parabolic curve. Reproduced with permission from Casaburi et al., Supercond. Sci. Technol. 31, 06LT01 (2018). Copyright 2018 IOP Publishing.
V. ANALYTICAL APPLICATIONS
A. General
The superconductor detectors are able to acquire novel data that no conventional detectors can do. The applications of STJDs to MS were reviewed in 1999183 and to synchrotron radiation in 2005.139 Material analysis for all the superconductor detector types was briefly reviewed in 2008.292 The applications of TESDs to x-ray or γ-ray spectroscopy also brought numerous excellent results.5,150 This chapter describes a limited number of material analysis results obtained at ETL-AIST in collaboration with other groups, focusing on XRS and MS with STJDs and SSPDs. Here, before describing our results of analytical science, it is worth listing the most recent topics on intriguing fundamental physical results: neutrino mass determination with STJDs using 7Be electron capture293,294 and quantum electrodynamics test in highly charged ions and muonic atoms with TESDs.295–297 Recent applications to x-ray astronomy, CMB, atomic physics, nuclear physics, and dark matter search appear in the LTD proceedings and papers cited therein.147
B. X-ray spectroscopy (XRS)
1. Scanning electron microscopy (SEM)
a. Electron-induced characteristic x-rays
Frank et al. reported the high performance of STJDs in a soft x-ray range below ∼1 keV in 1998.139 A low acceleration voltage of a few kV or less for the L-line emission brings a high spatial resolution of a few 10–100 nm as reported in 2001 by Wollman et al., who employed a TESD.298 If there is an analyzing condition that the L-line analysis has a better or comparable efficiency compared with the K-line analysis, the superconductor detectors play a crucial role at both sub-μm spatial resolution and a high detection efficiency, covering almost all elements below ∼2 keV.
Figure 41 shows the radiative and nonradiative processes after electron beam excitation. The electron impact on an atom creates a vacancy or hole on one of the electron shells, such as K, L1, L2, L3, M1, M2, M3, etc. The ionization cross section σ values were reported in Ref. 299. The K shell vacancy is mainly filled with an electron on one of the L-shells, emitting major characteristic x rays labeled as K-L2 (Kα1) and K-L3 (Kα2) lines. The two lines are normally indistinctive in EDS experiments so that we use the notation K-L2,3 (Kα1,2) collectively. For L-shell ionization, the L2 and L3 shells have a considerably higherσ than L1. The major lines are transitions from M4 and M5 to L3, which are labeled as L3-M4,5 (Lα1,2). The radiative x-ray emission processes are competitive with a nonradiative Auger electron emission from one of the L, M, or N shells. The radiative x-ray emission yield (ω) values are listed in Refs. 300 and 301. The total efficiencies of the x-ray emission are determined by multiplying σ and ω. In this article, we designate x-ray analysis with a hole on the K shell as the “K-line analysis,” and one with a hole on the L shell as the “L-line analysis.” Note that these processes also occur inside x-ray detector materials.
Radiative x-ray emission (red dotted arrows) and nonradiative Auger electron emission (black dotted arrows) after creating a vacancy or a hole on the K shell with electron impact (blue open circle). The vacancy leads to either radiative x-ray emission of K-L2(Kα2) or competitive nonradiative Auger electron emission from the L3 shell (KLL). Another transition is K-L3 (Kα1) emission or an Auger electron from the M shell (KLM). When the K-L2 and K-L3 lines are indistinguishable on measurement, they are collectively notated as K-L2,3 (Kα1,2). A transition from the M shell to the K shell is also possible: for example, K-M2,3 (Kβ1,3) that is not depicted. When the vacancy is created on the L3 shell, similar processes occur; for example, L3-L4,5 (Lα1,2). The notation for the characteristic x-ray lines is based on the IUPAC recommendation with the conventional Siegbahn notation in the parentheses for reference.
Radiative x-ray emission (red dotted arrows) and nonradiative Auger electron emission (black dotted arrows) after creating a vacancy or a hole on the K shell with electron impact (blue open circle). The vacancy leads to either radiative x-ray emission of K-L2(Kα2) or competitive nonradiative Auger electron emission from the L3 shell (KLL). Another transition is K-L3 (Kα1) emission or an Auger electron from the M shell (KLM). When the K-L2 and K-L3 lines are indistinguishable on measurement, they are collectively notated as K-L2,3 (Kα1,2). A transition from the M shell to the K shell is also possible: for example, K-M2,3 (Kβ1,3) that is not depicted. When the vacancy is created on the L3 shell, similar processes occur; for example, L3-L4,5 (Lα1,2). The notation for the characteristic x-ray lines is based on the IUPAC recommendation with the conventional Siegbahn notation in the parentheses for reference.
b. Comparison between K-line analysis and L-line analysis
We notice that numerous characteristic x rays can be emitted in addition to those in Fig. 41. Which characteristic x-ray line is the most efficient for material analysis? We should examine maximum ionization cross sections σ and radiative fluorescence yields ω. Figure 42(a) depicts the atomic number Z dependence of σ (blue curves) and ω (red curves). The ionization cross sections depend on both Z and electron energy. The dashed blue curve of σK for the K hole and the solid blue curve of σL3 for the L3 hole in (a) trace the maximum values of the electron energy dependent curves calculated by distorted-wave Born approximation (DWBA) method.299 The ωK dashed red curve and the ωL3 solid red curve plot the radiative fluorescence yields in the tables based on experimental and theoretical data.300,301 Figure 42(b) depicts the maximum effective fluorescence cross sections (σ×ω) for the K-L2,3(K) and L3-M4,5(L3) lines. The electron energies required for the maximum ionization cross sections are different from element to element; therefore, the SEM acceleration voltage should be adjusted for each element.
Atomic number Z dependence of maximum electron-impact ionization cross section σ and radiative fluorescence yield ω: (a) σK for the K shell (blue dashed line) and σL3 for the L3 shell (blue solid line), and ωK for the K-L2,3 (Kα1,2) line (red dashed line) and ωL3 for the L3-M4,5 (Lα1,2) line (red solid line); and (b) maximum effective fluorescence cross section (σ×ω) for the K-line analysis (K dashed line) and the L-line analysis (L3 solid line). The vertical thin dashed lines in (a) indicate Na and Zn that are present at a small amount inside dentin in Sec. V B 1 d. The maximum σ values are obtained from the electron-energy dependent curves calculated with a distorted-wave Born approximation (DWBA).299 The ω curves are based on experimental and theoretical data.300,301
Atomic number Z dependence of maximum electron-impact ionization cross section σ and radiative fluorescence yield ω: (a) σK for the K shell (blue dashed line) and σL3 for the L3 shell (blue solid line), and ωK for the K-L2,3 (Kα1,2) line (red dashed line) and ωL3 for the L3-M4,5 (Lα1,2) line (red solid line); and (b) maximum effective fluorescence cross section (σ×ω) for the K-line analysis (K dashed line) and the L-line analysis (L3 solid line). The vertical thin dashed lines in (a) indicate Na and Zn that are present at a small amount inside dentin in Sec. V B 1 d. The maximum σ values are obtained from the electron-energy dependent curves calculated with a distorted-wave Born approximation (DWBA).299 The ω curves are based on experimental and theoretical data.300,301
By comparing the K and L3 curves in Fig. 42(b), it discloses the advantage of the L-line analysis over the K-line analysis for Z > ∼30: the maximum effective cross sections for L3 are always higher than those of K. As an example, the effective cross section for the Zn L-line analysis at 5 keV is 5.8 times higher than that for the Zn K-line analysis at 20 keV, even though the L3 vacancy leads to the Auger electron emission at a probability of 98.8%. Zn is one of the important trace elements in dentin, as described later. Although a precise electron energy adjustment is necessary, the L-line analysis in an electron energy range of 3–10 kV is convincingly more efficient than the K-line analysis in 20–40 kV, neglecting electron penetration ranges and self-absorption. More importantly, electron cascade cloud sizes are normally of orders of 0.1 μm at 5 kV for the L-line analysis and ∼2 μm at 20 kV for the K-line analysis for dentin.302 Therefore, when we choose a correct electron beam energy, the L-line analysis should outperform the K-line analysis with respect to the spatial resolution for mapping as well as the detection sensitivity of elements.
c. SEM equipped with STJD (SC-SEM)
We validate the above-mentioned discussion on the advantage of the L-line analysis by using the superconductor-tunnel-junction equipped SEM (SC-SEM) developed by Fujii et al.259 Figure 43 shows the appearance of the SC-SEM system that employs a low-voltage SEM (JOEL JSM-7200F), a 100-pixel array of 100-μm square STJs, a polycarpellary x-ray lens, a 3He cryostat, and an FPGA-DSP readout-module stack. The silver cylindrical chamber behind the SEM lens barrel is the automatic 3He sorption cryostat (NIKI GLASS, LTS-HE3-LV). The SEM has a wide acceleration voltage range of 0.01–30 kV and spatial resolutions of 1.6 nm at 1 kV and 1.0 nm at 20 kV for secondary electron imaging. While scanning the electron beam, the 1-mm2 STJD and a 25-mm2 SDD simultaneously collect x rays emitted from a sample. To compensate a small STJ sensitive area, the polycapillary lens optics (XOS) is placed between the SEM sample stage and the STJD. The peak-height ratio of STJD/SDD for the K lines of light elements is approximately 0.17 as shown in Fig. 2. This ratio will be improved to ∼1.0 by improving the STJD performance and optimizing the optics in future. The STJD covers a range below ∼2 keV at an energy resolution of ∼6 eV, while the SDD covers a full energy range from ∼100 eV to ∼10 keV at an energy resolution of ∼130 eV.
Superconductor-tunnel-junction equipped SEM (SC-SEM).259 The low-voltage scanning electron microscope (JOEL JSM-7200F) has the 1-mm2 STJD and the 25-mm2 SDD. The acceleration voltage for the electron beam ranges from 10 V to 30 kV. The silver cylindrical metallic chamber behind the white SEM lens barrel is the automatic liquid-cryogen-free 3He cryostat (NIKI GLASS, LTS-HE3-LV). The STJD fabricated at the CRA VITY facility is mounted at the head of a cold finger (VACFIELD, cryofinger) at 0.3 K. A polycapillary lens optics (XOS) is placed between the SEM sample stage and the STJD. A stack of field-programmable-gate-array based digital signal processing (FPGA-DSP) modules (Morita E-design LLC., DTEMP) appears at the far left.
Superconductor-tunnel-junction equipped SEM (SC-SEM).259 The low-voltage scanning electron microscope (JOEL JSM-7200F) has the 1-mm2 STJD and the 25-mm2 SDD. The acceleration voltage for the electron beam ranges from 10 V to 30 kV. The silver cylindrical metallic chamber behind the white SEM lens barrel is the automatic liquid-cryogen-free 3He cryostat (NIKI GLASS, LTS-HE3-LV). The STJD fabricated at the CRA VITY facility is mounted at the head of a cold finger (VACFIELD, cryofinger) at 0.3 K. A polycapillary lens optics (XOS) is placed between the SEM sample stage and the STJD. A stack of field-programmable-gate-array based digital signal processing (FPGA-DSP) modules (Morita E-design LLC., DTEMP) appears at the far left.
Comparison of the SEM X-ray spectra that were acquired simultaneously with SDD and STJD at either 20 kV (red solid lines) or 5 kV (blue solid lines): (a) SDD spectra below 10 keV and (b) and (c) STJD spectra in an energy range indicated by the gray bar in (a). The insets show the semilogarithmic plots. The electron beam scanned a 13 × 17 μm2 area of a porcine dentin cross section for 10 min at 60 pA for 20 kV and 6 nA for 5 kV. The x-ray intensities (count/s) were normalized by the electron current (nA) and the energy bins (1 eV). The measurement location at 5 kV was different from that at 20 kV; the peak intensity patterns are different. For comparison, the SDD spectra divided by 6 are also plotted (gray solid lines) in (b, c). The peaks identified by the STJD are F K-line (676.8 eV), the Cu L-line (929.7 eV), the Zn L-line (1.012 keV), and the Na K-line (1.041 keV).
Comparison of the SEM X-ray spectra that were acquired simultaneously with SDD and STJD at either 20 kV (red solid lines) or 5 kV (blue solid lines): (a) SDD spectra below 10 keV and (b) and (c) STJD spectra in an energy range indicated by the gray bar in (a). The insets show the semilogarithmic plots. The electron beam scanned a 13 × 17 μm2 area of a porcine dentin cross section for 10 min at 60 pA for 20 kV and 6 nA for 5 kV. The x-ray intensities (count/s) were normalized by the electron current (nA) and the energy bins (1 eV). The measurement location at 5 kV was different from that at 20 kV; the peak intensity patterns are different. For comparison, the SDD spectra divided by 6 are also plotted (gray solid lines) in (b, c). The peaks identified by the STJD are F K-line (676.8 eV), the Cu L-line (929.7 eV), the Zn L-line (1.012 keV), and the Na K-line (1.041 keV).
d. Element mapping of porcine dentin
The sample for the L-line analysis was porcine dentin provided by Zaslansky et al.303 The dentin is a tooth part located between enamel and dental pulp. The dentin consists of minerals (70%), organic substances (18%), and water (12%). Fine hydroxyapatite mineral crystals surround the organic matter of collagen fibers and non-collagenous proteins. The major elemental components are C, N, O, P, and Ca with minor trace elements of F, Na, Mg, Al, Si, S, Cl, Cr, Cu, and Zn, which were identified by the SC-SEM. The concentrations of the trace elements are below 0.01%. Sodium is almost everywhere in tissues and other parts of animal bodies; the Na K-line normally masks Zn L-line on EDS spectra. The trace elements play an important role in oral environment against dental caries; the function and distribution of the trace elements are not fully understood.
The x-ray spectra in Fig. 44 were acquired for 10 min simultaneously with the STJD and the SDD, while scanning 13 × 17-μm2 area of a dentin cross section with an electron beam accelerated at either 5 or 20 kV. The electron beam currents were set at 6 nA for 5 kV and 60 pA for 20 kV, considering avoiding a serious radiation damage. The higher statistical fluctuation at 20 kV than that at 5 kV stems from the lower beam current. The SDD spectra in Fig. 44(a) contain the characteristic x-ray lines on the background continuum events due to Bremsstrahlung, of which maximum energies are equal to the electron energies. On the SDD spectrum at 20 kV (red solid line), the weak Zn K-L2,3 (8.63 keV) line is recognizable; however, it is too weak to measure Zn 2D distribution. On the other hand, one can recognize the strong peak at 1.03 keV, which conceivably includes the Zn L3-M4,5 line (1.012 keV) in addition to the Na K-L2,3 line (1.041 keV). However, the SDD spectra cannot clarify whether the 1.03 keV peak includes the Zn L3-M4,5 line or not. Similarly, the SDD spectrum at 5 kV (blue solid line) in Fig. 44(a) also has the peak at 1.03 keV.
In contrast to the SDD spectra, the STJD spectrum at 5 kV in Fig. 44(c) exhibits the clear separation of the Zn-L line from the Na-K line. The peak widths almost represent the natural linewidths of the elements in the matrix. For comparison, the gray solid line depicts the SDD spectrum divided by a factor of 6 in (b) and (c). A similar peak separation should be possible at 20 kV in (b); however, it is unclear because of the low characteristic x-ray/Bremsstrahlung ratio and the low electron-beam current of 60 pA. With the STJD L-line analysis at 5 kV, one can expect 2D imaging of Zn distribution, eliminating the Na K line. Figure 45 shows the elemental mapping images simultaneously acquired with the STJD and SDD. The dentin cross-sectional area of the upper-left SEM image was repeatedly scanned with an electron beam at 5 kV and 6 nA for 1h. The white lines in the SEM image indicate the tubules, which are surrounded by a strongly mineralized-tissue sheath called peritubular dentin (PTD) containing phospholipids. The space among the tubules is filled with intertubular dentin (ITD), which is a collagen matrix containing hydroxyapatite crystallites with sizes of sub-μm to nm. The structure and function of the tubules and PTD are not known completely.304
Element mapping images near the porcine dentin-cement interface. The field of view on the upper left was scanned with an electron beam at 5 kV and 6 nA for 1h. The x-ray intensity values are toned from black to white. The upper-left SEM image and the lower-left SDD F image exhibit that most parts of the peritubular dentin (PTD) contain F, whereas almost no F atoms exist in the intertubular dentin (ITD). The local area indicated by the white square boxes in the SEM image and the SDD F image is enlarged in the two STJD images for the F K-L2,3 and Zn L3-M4,5 lines. No recognizable accumulation of Zn in PTD is observed at this location regardless of the clear F accumulation in PTD.
Element mapping images near the porcine dentin-cement interface. The field of view on the upper left was scanned with an electron beam at 5 kV and 6 nA for 1h. The x-ray intensity values are toned from black to white. The upper-left SEM image and the lower-left SDD F image exhibit that most parts of the peritubular dentin (PTD) contain F, whereas almost no F atoms exist in the intertubular dentin (ITD). The local area indicated by the white square boxes in the SEM image and the SDD F image is enlarged in the two STJD images for the F K-L2,3 and Zn L3-M4,5 lines. No recognizable accumulation of Zn in PTD is observed at this location regardless of the clear F accumulation in PTD.
The lower-left SDD image exhibits that the F distribution coincides with the locations of the tubules: the broad F peak on the SDD spectrum in (c) can be used for imaging because it is isolated. The rectangular area in the SEM image is enlarged in the lower STJD images for the F K-L2,3 line and the Zn L3-M4,5 line. The STJD can retrieve the Zn L-line signal by properly setting a region-of-interest (ROI) energy window to the pulse height spectra: 997–1027 eV for the Zn L3-M4,5 line. At this field of view, most parts of PTD exhibit the enrichment of F, whereas Zn atoms distribute rather uniformly throughout both PTD and ITD. At a high spatial resolution with the 5-keV electron beam, we can see the F distribution inside the PTD walls, of which thickness is ∼1 μm. This is impossible with the K-line analysis at 20 kV.302
Figure 46 shows another location of the dentin cross section: the upper half panels show the SDD and SEM images. The lower half panels show the STJD images. One tubule runs along the diagonal direction. The SDD image for Na and Zn exhibits a slight elemental enrichment in the lower side of the PTD wall, but it is obscure which element is accumulated. On the other hand, the STJD Zn image exhibits that the Zn atoms are accumulated in the lower PTD side. The upper PTD wall contains almost no F and Zn. Interestingly, the Zn-enriched region coincides with the F distribution; therefore, the formation of ZnF2 inside the PTD wall is most probable. It seems that there are two types of element accumulations inside PTD: only F in Fig. 45 and both F and Zn possibly forming ZnF2.
Enlarged SEM element mapping images of the porcine dentin cross section: the upper half panels for the SDD images and the SEM image; and the lower half panels for the STJD images. The STJD and SDD images were simultaneously acquired. The field of view of the SEM image was scanned with an electron beam at 5 kV and 6 nA for 1h. The element concentration is toned from black to white. One tubule is running diagonally in this field of view. The high elemental separation performance of STJD is demonstrated for the Na K line and the Zn L line.
Enlarged SEM element mapping images of the porcine dentin cross section: the upper half panels for the SDD images and the SEM image; and the lower half panels for the STJD images. The STJD and SDD images were simultaneously acquired. The field of view of the SEM image was scanned with an electron beam at 5 kV and 6 nA for 1h. The element concentration is toned from black to white. One tubule is running diagonally in this field of view. The high elemental separation performance of STJD is demonstrated for the Na K line and the Zn L line.
Nonresonant x-ray emission spectroscopy (XES) with SC-SEM for α-Fe (black dotted line), Fe2O3 (red solid line), and Fe3O4 (blue dashed line). The x ray yields measured with the STJD were normalized at 705 eV. The surface of α-Fe is oxidized, so that the weak peak of the O K-L2,3 (Kα1,2) line appears. The shapes of the Fe L3-M1 (Lι) lines are almost the same, whereas the line shapes combining Fe L3-M4,5 (Lα1,2) and Fe L2-M4 (Lβ1) are influenced by the chemical states of Fe atoms.
Nonresonant x-ray emission spectroscopy (XES) with SC-SEM for α-Fe (black dotted line), Fe2O3 (red solid line), and Fe3O4 (blue dashed line). The x ray yields measured with the STJD were normalized at 705 eV. The surface of α-Fe is oxidized, so that the weak peak of the O K-L2,3 (Kα1,2) line appears. The shapes of the Fe L3-M1 (Lι) lines are almost the same, whereas the line shapes combining Fe L3-M4,5 (Lα1,2) and Fe L2-M4 (Lβ1) are influenced by the chemical states of Fe atoms.
Both F and Zn are known as the root caries inhibitors according to the World Health Organization (WHO); however, it is still unknown whether F and Zn are essential trace elements or not. Based on our observation, the accumulation patterns of F and Zn vary depending on the locations. The F atoms are enriched inside PTD at most locations, whereas the accumulation of Zn occurs at the limited locations of PTD walls. We infer that both F and Zn are supplied to the dentin through the odontoblastic processes and diffuse into the PTD and ITD regions. During this element flow, the Zn and F atoms are accumulated together inside the PTD wall at some places. The ZnF2 formation that fixes both elements inside PTD can be a sign of excessive ingestion. STJD element mapping with the L-line analysis paves the way to understanding the roles of F and Zn in teeth. The details of the dental research will be reported separately.
e. X-ray emission spectroscopy with SC-SEM
The natural linewidths of light elements in matrices are 15–25 eV, which are considerably broader than those of isolated atoms because of chemical bonding to other atoms. The STJD energy resolution is significantly narrower than those linewidths; therefore, it enables measuring line shapes reflecting chemical states or electronic structure of valence bands. This is called x-ray emission spectroscopy (XES).
Figure 47 shows an example of the XES spectra. We measured the materials containing Fe: metallic iron (α-Fe) and two iron oxides of Fe2O3 and Fe3O4. One can recognize that the higher-energy shoulder shapes between 710 and 730 eV are different depending on the chemical states of the Fe atoms. Almost no Fe L2-M4(Fe Lβ1) line appears in the α-Fe sample. From the intensities of the Fe L2-M4 lines, it is possible to distinguish Fe2O3 and Fe3O4. Charge states and the surrounding environment of the Fe atoms influence the x-ray line shapes for the transitions from M4 or M5 to L2 or L3. By using a suitable ROI window on XES spectra, imaging of the different chemical states is possible in principle. Chemical nano-imaging at the adhesion interface of carbon fiber reinforced plastic (CFRP) parts is one of the challenging tasks to visualize chemical bonds.305 However, there was no distinctive difference of the XES spectra among the adhesion samples prepared in different conditions. XES nano-imaging is harder than element mapping because of an insignificant change in line shapes.
Besides STJDs, a TESD on SEM has an energy resolution of 2.0 eV that is better than 6 eV of the STJD.306 An SEM-XES instrument with TESD is also under development at NIST since 1997.307,308 The development of SEM with a TES array detector is in progress at Los Alamos National Laboratory (LANL)309 and Hitachi High-Tech.310 The TES type is also utilized for scanning transmission electron microscopy (STEM) at Kobe Steel and JAXA.311,312 In addition to electron probes, an x-ray tube is utilized for XES at NIST.313 Although the STJD energy resolution is a few times worse than that of the best TESDs in a soft x-ray range, the STJD with a high count rate up to 200 kcps may have the advantages of quicker data acquisition and imaging within a tolerable measurement time. No peak shift is expected in STJDs even when the base temperature fluctuates for a long-time measurement because of 2Δg.
2. X-ray absorption spectroscopy (XAS)
a. X-ray absorption fundamentals
X-ray absorbance as a function of photon energies exhibits a fine structure near absorption edges, creating a hole on an electron shell in Fig. 41. The fine structure is called XAFS. It is divided into two regimes: x-ray absorption near-edge structure (XANES) within ∼50 eV from absorption edges, and extended x-ray absorption fine structure (EXAFS) up to 1000 eV above edges.314 Normally, a monochromatized synchrotron radiation beam is employed to measure XAS spectra. Figure 48 illustrates the relationship among XANES, EXAFS, and XES. In contrast to XANES representing an unoccupied conduction band structure, XES is a method to deduce the electronic structure of the outer shells that are influenced by the surrounding chemical environment. Therefore, XANES and XES are complementarily employed for material analysis. EXAFS represents a symmetry of local atomic structure or interatomic distances.
Relationship among XANES, EXAFS, and XES. An atom absorbs an x-ray photon, exciting an inner-shell electron to a conduction band in XANES, and to a vacuum state in EXAFS that corresponds to ionization. XANES and XES provide complementary data on electronic structure of conduction band and valence band. EXAFS represents symmetry of local atomic structure. Pre-edge peaks for transition to unoccupied anti-bond orbitals such as π* in compounds or molecules occasionally emerge below absorption edges.
Relationship among XANES, EXAFS, and XES. An atom absorbs an x-ray photon, exciting an inner-shell electron to a conduction band in XANES, and to a vacuum state in EXAFS that corresponds to ionization. XANES and XES provide complementary data on electronic structure of conduction band and valence band. EXAFS represents symmetry of local atomic structure. Pre-edge peaks for transition to unoccupied anti-bond orbitals such as π* in compounds or molecules occasionally emerge below absorption edges.
A straightforward method to acquire XAFS spectra is measuring the intensity of x rays transmitting a thin-film sample while increasing the probe x-ray energy across absorption edges. However, the sample should have a thin-film form, and more seriously, trace elements in matrices cannot be measured. Therefore, other two methods are often selected for analyzing bulk samples or trace impurity elements: electron yield (EY) measuring photoelectrons and Auger electrons; and fluorescence yield (FY) measuring characteristic x rays emitted after x-ray absorption as depicted in Fig. 41 for electron excitation. Both EY and FY methods are divided into two regimes: “total” and “partial.” The total yield method accepts the whole energy range of electrons and x-rays emitted from a sample, whereas the partial method accepts a selected energy range. The partial electron yield (PEY) method can select a depth range within a few nm for photoelectron or Auger electron emission; therefore, the surface treatment before measurement is crucial. On the other hand, not only can the partial fluorescence yield (PFY) selectively identify elements of interest, but it can also analyze a thicker surface layer: for example, a surface layer of a few 100 nm for the nitrogen dopant in ZnO. Therefore, PFY-XAFS has the advantages of analyzing the bulk structure and trace impurity elements embedded in the matrix lattice structure. However, semiconductor EDS detectors often face insufficient element selection as shown in Fig. 44, which is a major motivation for developing XAFS with a superconductor detector.
The first PFY-XAFS instrument with STJD was developed at LLNL for Advanced Light Source (ALS) by Friedrich et al. in 2002.315 The end-station instrument was equipped with a 9-pixel STJD with 200-μm square Nb/Al junctions with an energy resolution of ∼15 eV below 1 keV. With this LLNL instrument, nitrogen-doped ZnO films epitaxially grown on a ZnO(0001) substrate at AIST were measured in 2004. The data were analyzed by Fons of AIST in 2006.316 Later, the Nb/Al-based STJD was replaced by the Ta/Al-STJD that has a better energy resolution.77 Besides STJDs, TESDs have also acquired remarkable data on XAFS and RIXS in a soft x-ray range of 100–2000eV.151,317,318
We reported our first XAS experiment with a single Nb/Al junction for a high-k gate material of Hf-Al oxide for MOS devices in KEK PF in 2006 as mentioned in Sec. IV B 2 before.267 The single junction was replaced by the 100-pixel Nb/Al-STJD in 2009.273,319 We successfully analyzed nitrogen-doped SiC samples at 800 ppm in 2012 and at 20 ppm in 2020.320,321 The following two sections describe these results on nitrogen dopant in ZnO and SiC.
b. Nitrogen dopant in ZnO
The wide-bandgap semiconductor ZnO is a promising material for optoelectronics such as UV light-emitting diodes. ZnO films were naturally of n-type because of defects such as oxygen vacancies and interstitial zinc atoms; p-type doping is challenging. Nitrogen is a promising p-type dopant. The in situ nitrogen doping during the epitaxial growth of ZnO was performed by molecular beam epitaxy with a plasma radical nitrogen source, which produced yellowish films with a thickness of ∼1 μm. The N concentration was approximately 1% (∼3 × 1020 cm−3). After rapid thermal annealing, the films changed to transparent. The N K-edge at 397 eV is lower than the O K-edge at 533 eV. Therefore, the XANES measurement for the N K-edge was uninfluenced by characteristic x rays from the matrix elements: O K-L2,3 (525 eV) and Zn L3-M4,5 (1.012 keV). However, in those days around 2004, semiconductor detectors such as Si and Ge installed in synchrotron radiation beamlines barely recognized an O K-line peak above an electrical noise. The N K-line was normally buried under the noise level. Therefore, a better S/N ratio was required for the PFY-XAFS measurement. One was the windowless Ge detector at ELETTRA. Another better one was the LLNL-STJD at ALS with a high S/N ratio as well as a better energy resolution.
Figure 49(a) shows the ALS-XAFS spectra before and after annealing. The XANES black spectrum for the as-grown state exhibits a sharp peak at 400.7 eV and a wide fine structure from 395 eV in the pre-edge region to ∼430 eV above the edge. An ab initio multiple-scattering calculation with the FEFF code322 revealed that the N atoms substituted for the O site in the as-grown state.316 The XAFS red spectrum in Fig. 49(a) indicates that after annealing, the fine structure almost disappears except the sharp peak at 400.7 eV. The expanded spectra measured at different polarization directions of the x-ray probe beam are shown in Fig. 49(b). The vibrational structure is typical for the nitrogen molecule N2 gas. Since there is almost no polarization dependence, it is concluded that the N2 molecules are not oriented to the ZnO crystal but stored in the form of randomly oriented N2 bubbles. It signifies that the N substitution for the O site is metastable. This is the reason why reproducible synthesis of p-type ZnO is troublesome.