Next Article in Journal
Holography Using Curved Metasurfaces
Next Article in Special Issue
Nearly Single-Cycle Terahertz Pulse Generation in Aperiodically Poled Lithium Niobate
Previous Article in Journal
Multi-Spectral Quantum Cascade Lasers on Silicon With Integrated Multiplexers
Previous Article in Special Issue
Bow-Tie Cavity for Terahertz Radiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

THz Mixing with High-TC Hot Electron Bolometers: A Performance Modeling Assessment for Y-Ba-Cu-O Devices

1
CentraleSupélec, CNRS, Univ. Paris-Sud, Université Paris-Saclay, Sorbonne Université, Group of Electrical Engineering, GeePs, 91190 Gif sur Yvette, France
2
Institute for High Frequency and Communication Technology, University of Wuppertal, D-42119 Wuppertal, Germany
*
Author to whom correspondence should be addressed.
Photonics 2019, 6(1), 7; https://doi.org/10.3390/photonics6010007
Submission received: 15 December 2018 / Revised: 17 January 2019 / Accepted: 24 January 2019 / Published: 25 January 2019
(This article belongs to the Special Issue Terahertz Photonics)

Abstract

:
Hot electron bolometers (HEB) made from high-TC superconducting YBa2Cu3O7x (YBCO) oxide nano-constrictions are promising THz mixers, due to their expected wide bandwidth, large mixing gain, and low intrinsic noise. The challenge for YBCO resides, however, in the chemical reactivity of the material and the related aging effects. In this paper, we model and simulate the frequency dependent performance of YBCO HEBs operating as THz mixers. We recall first the main hypotheses of our hot spot model taking into account both the RF frequency effects in the YBCO superconducting transition and the nano-constriction impedance at THz frequencies. The predicted performance up to 4 THz is given in terms of double sideband noise temperature TDSB and conversion gain G. At 2.5 THz for instance, TDSB ≅ 1000 K and G ≅ − 6 dB could be achieved at 12.5 μW local oscillator power. We then consider a standoff target detection scheme and examine the feasibility with YBCO devices. For instance, detection at 3 m through cotton cloth in passive imaging mode could be readily achieved in moderate humidity conditions with 10 K resolution.

1. Introduction

First considered by Gershenzon et al. [1], the superconducting hot electron bolometer (HEB) principle was applied by these authors to low-TC (TC is the superconducting critical temperature) and high-TC HEBs, Nb and YBa2Cu3O7x (with x lower than 0.2, and called YBCO hereafter), respectively. The concept then fruitfully evolved [2], rather in favor of low TC devices to start with, mainly Nb and NbN, exhibiting the diffusion cooled and phonon cooled processes, respectively [3]. In fact, the research on HEBs was mainly driven by immediate applications to THz radio astronomy as mixers for heterodyne reception [4,5]. Besides, YBCO was attractive in several respects, e.g., operating temperature with light cryogenics, or fast response, due to its very short electron-phonon relaxation time τep, in the ps range [6,7,8]. Early YBCO HEB mixers were demonstrated at mm-wave [9,10,11,12] or THz [9,10,13,14,15] frequencies. Difficulties were encountered, however, which mainly arose from YBCO chemical reactivity to atmospheric water and carbon dioxide and to the fabrication process as well, which compromised the durability of good performances for frequency mixing operation [16]. They also arose from the superconducting film to substrate phonon escape time τesc, much longer for YBCO [17] than for Nb, for instance [18], with a detrimental effect on the instantaneous bandwidth [11,12,13,19]. More recently, MgB2 has met a sustained interest as a compromise between NbN and YBCO, in terms of operating temperature, but also in terms of instantaneous bandwidth [20,21].
The just mentioned experimental work was permanently sustained by modeling and simulation efforts to predict and interpret HEB performance. We now introduce those performance simulation aspects, because they indeed are our main concern in the present article.
Several steps in the HEB modeling approach have to be considered. In the “0D” or point bolometer model, only energy transfer between thermal reservoirs is evaluated. In the superconducting material (Figure 1a), the electron reservoir at temperature Te interacts with the phonon reservoir at temperature Tp; this interaction is governed by the above-mentioned electron-phonon relaxation time τep. The phonons release their energy to the device substrate with the above-mentioned escape time τesc [22]. If the electron-electron interaction time τee is much shorter than τep, (which is the case for Nb, NbN or YBCO), hot electron bolometric action occurs, such that Te > Tp, provided that τesc << τep. This latter condition is related to the film-substrate interface quality, which is mainly correlated with the device technological elaboration process. As a further refinement of this model, a three thermal reservoir model can be considered, including the substrate to the cold finger thermal resistance [23]; although 0D, this latter model also discusses the dimensions of the device constriction to achieve good sensitivity HEBs. As shown in Figure 1b, the constriction is defined by its length L, width w and thickness θ.
Further modeling developments took the HEB device geometry into account to describe the electron temperature profile Te(x) with respect to the constriction length coordinate (0 < x < L). First developed for low TC HEBs [24], this so-called “hot spot” modeling was further applied to high TC devices [25]. The YBCO HEB model then evolved to include more realistic descriptions, such as the phonon temperature spatial dependence Tp(x) [26] and the operating radio frequency (RF) influence on the shape of the resistance vs. temperature superconducting transition [27].
The aim of this article is twofold. First, we shall provide a complete and detailed description of the YBCO HEB model and extend the results of [25,26,27] up to 4 THz. We shall also compare our modeling results in terms of the device performance as a THz heterodyne mixer with the few experimental data available for YBCO HEB mixers, which can be found in the literature.
Second, we shall consider the potential of such a device to provide a lightweight system for THz standoff detection for such applications as security and control. We shall therefore evaluate the standoff detection limits using passive imaging with YBCO HEB mixers. In this framework, we shall simulate well-defined and credible scenarios in terms of target distance, temperature resolution, atmospheric relative humidity, and transmission through clothing.

2. Models and Methods

2.1. Describing the Superconducting Transition at THz Frequencies

To have access to the resistance (and, more generally, the impedance) vs. electron temperature Te of the YBCO constriction at angular frequency ω = 2πf, we describe the superconductor by a two-fluid (2F) model. The concentrations of the superconducting and normal charge carriers are nS and nN, respectively, with nS + nN = n, the total free electron concentration [28]. Defining the two-fluid critical temperature TC2F, the dependence of nN/n = fN(tr) as a function of reduced temperature tr = Te/TC2F was chosen as fN(tr) = tr4. This choice minimizes the electron system free energy, according to the Gorter-Casimir 2F model [28]. Hence the real and imaginary parts of the complex conductivity σ = σ1 − jσ2, assuming Ohm’s law is still valid in an impurity scattering description [29]:
t r < 1 : σ 1 = σ N t r 4 + σ 2 r , σ 2 = 1 ω μ 0 λ 0 2 r 2 1 + r 2 ( 1 t r 4 ) ,
t r 1 : σ 1 = σ N , σ 2 = 0 ,
where σN is the normal state conductivity, r = ωτS is the impurity scattering parameter (for scattering time τS), μ0 is the permeability of a vacuum and λ0 is the London’s penetration depth at tr = 0.
As the two-fluid model depicts a homogeneous superconducting material, we have to describe the experimentally observed broadening of the superconducting transition; such a broadening is related to the film microstructure (e.g., grain boundary weak links). This is performed by introducing a Gaussian distribution of critical temperature values, centered at the mid-transition critical temperature TC0, as discussed in Reference [27], with
g = 1 G N exp ( ( T e T C 0 ) 2 2 ( Δ T G ) 2 ) ,
where GN is a normalization coefficient and ΔTG is the critical temperature standard deviation, to be adjusted to fit the experimental transition.
After having expressed the two-fluid resistivity ρ 2 F ( ω ) = ρ 1 2 F ( ω ) + j ρ 2 2 F ( ω ) = σ 1 , we then fit the experimental DC resistivity transition [30] as a convolution product between ρ12F (ω → 0) and the Gaussian function, so that
ρ 1 exp ( T e , T C 0 , Δ T G ) = ( ρ 1 2 F ( ω 0 ) g ) ( T e , T C 0 , Δ T G ) .
The so-obtained transition is then approximated by a Fermi-Dirac type function, as more appropriate to further numerical treatments (see below Section 2.2, hypothesis 8).
To obtain the THz dependence, we introduce ρ2F(ω) in the so-obtained DC transition, and then fit a new Fermi-Dirac-like transition with frequency dependent parameters (mid-transition temperature, transition width and minimum resistivity, as detailed in Section 3.1) [27].

2.2. HEB Master Equations

For an HEB constriction of volume VC = L × w × θ (Figure 1b), the thermal power balance is governed by the heat diffusion equations for the electron and phonon baths, represented by Equation (4a,b), respectively, as follows:
δ δ x [ V C κ e δ T e δ x ] V C C e δ T e δ t = V C C e ( T e n T p n ) n τ ep T ec n 1 L δ P J - tot δ x ,
δ δ x [ V C κ p δ T p δ x ] V C C p δ T p δ t = V C C p ( T p m T 0 m ) m τ esc T pc m 1 V C C e ( T e n T p n ) n τ ep T ec n 1 ,
where indices e and p pertain to the electrons and phonons of the superconducting film, respectively (Figure 1a). κe and κp are the corresponding thermal conductivities, Ce and Cp the unit volume specific heats (evaluated at temperatures Tec and Tpc, respectively). Integers n and m characterize the thermal exchanges between electrons and phonons and phonons and substrate (at reference temperature T0), respectively. The source term in Equation (4a) represents the Joule power locally injected in a constriction slice of width δx. This power results from both DC and THz currents flowing along the constriction of cross-sectional area SC = w × θ. More specifically, after having dropped the higher order terms:
δ P J - tot ρ ( 0 , T e ( x ) ) S C I 0 2 δ x + ρ ( ω LO , T e ( x ) ) S C × [ 1 2 I LO 2 + I LO I S cos ( ( ω LO ω S ) t ) ] δ x .
In this equation, I0 is the DC bias current, ILO (local oscillator) and IS (signal) are the THz current amplitudes at angular frequencies ωLO and ωS, respectively. The resistivity functions are deduced as discussed in Section 2.1. It therefore appears that (i) the current components are constant over SC (see below, hypothesis 2) and (ii) the knowledge of Te(x) is required; this implies an iterative resolution procedure for the coupled Equation (4a,b).
The hypotheses governing the resolution of Equation (4a,b) are given in the following list.
  • These equations obviously describe a phonon cooling mechanism for the electrons in YBCO, as opposed to Nb HEBs, where cooling by electron diffusion to the metal contacts is the dominant process [3,31]. In fact, using YBCO data (cf. Section 3.2, Table 1), we evaluate the electron diffusion length le ≈ π(τepκe/Ce)1/2 ≅ 25 nm, a significantly smaller value than the constriction length (100 to 400 nm) considered here, so that phonon cooling will prevail [22].
  • It also appears from Equation (5) that the DC and THz current densities are assumed to be constant across the area SC, a point discussed at length in Reference [32]. Due to the resistivity vs. Te(x) dependence, DC and THz powers are non-uniformly absorbed along the constriction, as noticed in Reference [33].
  • The constriction ends (at coordinates x = 0 and x = L) are at reference temperature T0, which is the cryostat/cryogenerator cold finger temperature (gold antenna contacts). So that: Te(0) = Tp(0) = Te(L) = Tp(L) = T0.
  • The solutions Te(x) and Tp(x) follow the geometrical symmetry of the constriction with respect to its center: Te(x) = Te(Lx) and Tp(x) = Tp(Lx). Consequently, the equation solving will be performed in a half-constriction, e.g., x ∈ [L/2, L].
  • The dissipated power in the constriction tends to raise Te(x) and Tp(x), which will reach their maximum values at the constriction center, so that: Temax = Te(L/2), and Tpmax = Tp(L/2), with dTe(L/2)/dx = dTp(L/2)/dx = 0.
  • The energy exchange exponent n = 3 was taken for YBCO, as extracted from electron-phonon interaction time measurements (see, e.g., Reference [1]). The YBCO to substrate phonon mismatch index m = 3 resulted from the number of available phonon modes, proportional to T3 (Debye’s model).
  • The YBCO to MgO substrate phonon escape time was deduced as τesc (ps) = 75 × θ (nm) [17]. In the case of PBCO/YBCO/PBCO trilayer constriction [30], where PBCO lies for PrBa2Cu3O7–y, a correction should be applied to θ, as considered in Reference [13].
  • For numerical computation convenience, the resistive superconductive transition has been approximated by a Fermi-Dirac function of the form ρ = ρ0 [1 + exp((TCTe)/ΔT)]−1, to fit the variation given by Equation (3).
  • Due to thermal effects, TC is sensitive to the constriction DC bias current IDC; this effect has been taken into account by writing TC(IDC) = TC(0)[1 − (IDC/(JrefSC))2/3], where Jref is a reference current density. This point is mentioned in Reference [25] and discussed in Reference [34], in the context of YBCO thin film electrical transport vs. microstructure relationship.
  • Because we can have access to the constriction RF impedance, it was possible to handle the effective power dissipated in the constriction as α × PLO, with α = αimp × α’, where αimp is the impedance matching factor between the constriction and the antenna, and α’ represents all the losses from other origins (focusing optics, diplexer, etc.).

2.3. Solving HEB Master Equations

In the two aforementioned Equation (4a,b), we set certain parameter values to limit to two the number of variables to find by the iterative solving method. The known parameters are VC, κe, Ce, n, τep, α’, SC, κp, Cp, m, τesc. Formally, the fixed (input) variables are PLO, I0 and T0, because in a measurement procedure, these are the parameters controlled by the operator. The resolution of the equations should therefore be performed on the (output) variables Te(x) and Tp(x), and their second derivatives d2Te/dx2 and d2Tp/dx2 as well.
All the above considerations give enough information that the second derivatives can be deduced at each point by our program when x varies from L/2 to L. Similarly, Te(x) and Tp(x) are deduced at each point provided that we define Temax and Tpmax, where the relation TemaxTpmax will help us to converge. The basic resolution approach is therefore to test (Temax, Tpmax) pairs until the value found at the extremity of the constriction is: Te(L) = Tp(L) = T0, which represents our criterion of success for the calculation.
Furthermore, we have made a significant improvement of this basic resolution process, by exchanging the roles of Temax and I0, such that Temax is now a fixed parameter and I0 a computational variable. This “trick” facilitates the calculations, because, on the one hand, the Tpmax solution search is carried out with a fixed upper bound (Temax is known). On the other hand, difficult cases where different constriction resistance (R) values can be associated with the same I0 value (because I0 is surjective towards R), can be simplified: In fact, Temax is strongly correlated with the resistance (because Temax is bijective with R in all our calculations).
Unlike the formal procedure just described, Equation (5) indicates that ILO (and not PLO) is the THz input, PLO being deduced at the end of the calculation. Therefore, it is not possible to perform a calculation that will accurately give a result ascribed to a given PLO value. What we obtained by our method was a “cloud of points” where each of these points is a configuration of the HEB operating at a certain power PLO, at a certain bias current I0 and having a certain resistance R (as well as all the other characteristics that we wish to deduce).
The goal of our approach is to obtain a “cloud of points” as homogeneous as possible, in order to extrapolate the operating configurations we are interested in. Typically, we recreated, by interpolation, a series of points that are the I-V responses to the different PLO values we wished to investigate (e.g., PLO ∈ [0, 50 μW]). With these new series of points, we were able to reuse the processing tools for the calculation of the conversion gain and the noise temperature, as further commented in Section 2.4.
The global solving loop process is summarized as follows. Starting with Equation (5), the local heat power dissipation leads to the global heterodyne power mixing expression:
P J - tot P DC + α P LO + 2 α [ P LO P S cos ( ( ω LO ω S ) t ) ] ,
where appears the intermediate frequency (IF) ω IF = | ω LO ω S | . As ILO is homogeneous along the constriction, the RF heat dissipation αPLO is then deduced from the ZRF = RRF + jXRF THz impedance. The result is then fed back into Equation (5), according to:
I LO = 2 α P LO R RF ,
the constriction resistance/reactance contributions being
R DC = 0 L ρ ( 0 , T e ( x ) ) S C d x , R RF = 0 L ρ 1 ( ω LO , T e ( x ) ) S C d x , and X RF = 0 L ρ 2 ( ω LO , T e ( x ) ) S C d x .
It should be noticed that in Equation (7), α = αimp × α’ (hypothesis 10 above) is a function of ZRF, through the impedance matching coefficient
α imp = ( 1 | Z RF Z a * Z RF + Z a | 2 ) × Re ( Z RF ) | Z RF | ,
where Za is the antenna impedance. The left hand term is the regular impedance matching expression, whereas the right hand term is a correction ratio between the RF power dissipated in the constriction and the total input power.
For a self-complementary antenna, Za = Ra + jXa is purely real, with
Z a = R a = 60 π ( 1 + ε r ) / 2 ,
where εr is the substrate material dielectric constant [35,36].

2.4. HEB Mixer Performance

2.4.1. General Considerations and Conversion Gain

For the calculation of the mixer conversion gain and noise temperature, it is necessary to discriminate between: (i) The contribution originating from the DC bias power PDC, and (ii) the contribution originating from the RF power (PLO and PS). To do this, we first define two parameters KDC and KRF as:
K DC = δ R δ P DC ( δ P LO = 0 ) ,
K RF = δ R δ P LO ( δ P DC = 0 ) ,
where R is the constriction resistance (evaluated according to the adopted model), and assuming PS << PLO.
KDC can be deduced from the device IV relationship at given PLO, whereas KRF cannot be deduced directly. Indeed, the bolometer electrothermal feedback implies that a PLO variation introduces a variation of R, and thus a change of the dissipated PDC. This feedback, which involves power variations ΔPDC and ΔPLO, can be taken into account in KRF from the resistance variation
Δ R = Δ P LO K RF + Δ P DC K DC , whence K RF = Δ R Δ P LO K DC Δ P DC Δ P LO .
The expression of the conversion gain G = (IF power delivered to the load RL)/(incident RF signal power) = PL/PS can be deduced from the small signal IF current I ˜ , as (see, e.g., Reference [24]):
G = R L I ˜ 2 / 2 P S = 2 I 0 2 R L K RF 2 P LO ( R 0 + R L ) 2 [ 1 K DC I 0 2 ( R L R 0 ) ( R L + R 0 ) ] 2
where R0 = R − ΔR and I0 is the constriction current when PS = 0.
It is to be noticed that the gain is also involved in the noise expression. Let us consider the noise measured at the mixer input (index “in”) and the noise measured at the mixer output (index “out”); the corresponding relationship for the noise temperatures is Tin = Tout/G.

2.4.2. Noise Temperature Contributions

The two main noise contributions of internal origin in a superconducting HEB are the Johnson (Jn) noise and the thermal fluctuation (TF) noise [3,37].
Johnson noise is considered as an additional voltage source vJn in series with the bolometer, such that v Jn 2 = 4 k B T 0 R 0 Δ f , where Δf is the output signal bandwidth. In this context, vJn introduces variations ΔRJn and I ˜ Jn , hence the noise power P Jn out = k B T Jn out Δ f dissipated in the mixer load resistance RL. Using the conversion gain expression (Equation (13)), one readily obtains the dual sideband (DSB) expression of the mixer input Johnson noise:
T Jn , DSB in = R 0 T 0 I 0 2 K RF P LO .
The noise spectral density originating from thermal fluctuations of the electrons involves the electron-electron interaction time τee and is expressed as follows [24]:
T e = 4 k B T e 2 τ ee C e V C .
A current I ˜ TF results from these fluctuations hence the noise power P TF out = k B T TF out Δ f dissipated in the load resistance RL. In the same manner as previously, one obtains the DSB expression of the mixer input thermal fluctuation noise:
T TF , DSB in = 1 K RF 2 P LO ( d R d T e ) 2 T e 2 C e V C τ ee .
The global expression of the mixer noise temperature is therefore
T N , DSB in = T Jn , DSB in + T TF , DSB in + T IF 2 G ,
where we have introduced the IF amplifier output noise temperature TIF, hence the approximate expression TIF/2G for the DSB input noise temperature. Besides, we have considered here TIF as independent of the working frequency and equal to the cooling temperature T0 of the HEB [38].
In practice, the mixer performance can be accessed by widespread techniques, such as the Y-factor method [39], which consists in using a blackbody (BB) source alternately adjusted at temperatures TBB = Thot or Tcold, and placed at the mixer input. Defining Y = (Phot/Pcold)out as the ratio between the associated mixer output powers (in the mixer bandwidth Δf), one obtains for the mixer input noise TNin = (ThotYTcold)/(Y − 1) in the Rayleigh-Jeans limit [39]. Moreover, the mixer gain is equal to the slope of the Pout vs. kBTBBΔf straight line plot (kB is the Boltzmann constant).

2.5. Standoff Detection Implementation with an HEB Heterodyne Detector

Passive imaging techniques rely both on collecting naturally emitting radiation from a scene and detecting the contrast between warmer and colder targets. The human being exhibits an emissivity of about 95% at frequencies above 600 GHz, which makes such a target appear warm relatively to a metal object of low emissivity.
We have checked the feasibility of standoff target detection operating in the passive mode with a YBCO HEB THz mixer. We have considered the simplified schematic in Figure 2, which illustrates losses of various origins. The noise equivalent power (NEP) requirement for the detector was deduced from the blackbody power PBB emitted by a target of area AT and collected by a detector of area AD (the HEB focusing lens effective area). For a receiver of bandwidth 2Δν at center frequency ν0 and for a temperature difference ΔT to be resolved at Top, the power difference is
Δ P BB ( Δ ν , Δ T ) = ν 0 Δ ν ν 0 + Δ ν [ B ν ( T op + Δ T ) B ν ( T op ) ] d ν ε T A D A T d T 2 ,
where Bν(T) is the blackbody spectral radiance (in W · m 2 · sr 1 · Hz 1 ), εT the target emissivity, and dT the target to detector distance (possibly corrected to include imaging system primary and secondary mirrors).
The NEP was then obtained by assuming the resolution to be diffraction limited for the detector (i.e., at the focusing lens focal plane). Consequently, the resolved distance at the target can be related to the Airy disk diameter at the detector focal plane by a factor dT/fL, approximately, where fL is the focal length of the focusing lens. We finally work out at wavelength λ:
N E P = ν 0 Δ ν ν 0 + Δ ν [ B ν ( T op + Δ T ) B ν ( T op ) ] d ν δ ν int ε T ( π 1.22 λ 2 ) 2 t opt t atm t obs ,
where δνint = 1/(2πτint), τint being the (e.g., post-detection) integration time; topt, tatm, and tobs are the transmission factors related to optics, atmosphere and obstacles, respectively (Figure 2). These factors are considered in the following.
  • topt includes the HEB planar antenna main lobe efficiency, the focusing lens and the detector cryostat window losses, and the LO injection losses (diplexer or beam splitter).
  • tatm was extracted from the atmospheric attenuation factor αatm (in dB/m) in various conditions, using the Atmospheric Effects Model (AEM) described in Reference [40]. Such extracted values were also in line with values deduced from data in Reference [41].
  • tobs represents the transmission through some obstacle existing in front of the target (cloth, cardboard, etc.).
The double sideband noise temperature required for the HEB was deduced from the NEP in the Rayleigh-Jeans limit as [39]:
T DSB = N E P 2 k B 2 Δ ν .

3. Simulation Results

3.1. YBCO Superconducting Transition in the THz Range

As shown in Section 2.3, we have been able to link the experimental DC transition with a two-fluid model representation of the resistive transition when ω → 0. From this result, we perform at terahertz frequencies a similar convolution as in Equation (3), which will therefore modify the modeled resistive transition accordingly. We approach this modification as the combination of three phenomena: The decrease of the average mid-transition critical temperature (TC), the widening of the resistive transition (ΔT) and the appearance of a minimum resistivity (ρmin). To the first order, we write:
TC(fTHz) = TC(0) × (1 − AT fTHz),
ΔT(fTHz) = ΔT(0) × (1 + BT fTHz),
ρmin(fTHz,tr = 0.8) ≅ 2ρmin(fTHz,tr = 0) = CT fTHz.
In practice, the AT, BT and CT parameters are introduced in the expression of the resistivity according to a Fermi-Dirac function (for convenience in our simulation). Their values are calculated for a given material and approximated for the 0 to 5 THz frequency range. Thus, in a first step, it is necessary to calculate CT from the expression of ρmin deduced from ρ2F(ω,tr → 0), as presented in Equation (21c). The pair of parameters (AT, BT) is then calculated by determining the pair which minimizes the difference function between the gaussian fit and the current resistivity function |ρG(T, ω) − ρ (T, ω)|, in both the 50 to 100 K and 0 to 5 THz ranges.
From an experimental point of view, our current technological process allows to achieve YBCO constrictions of length and width L = w = 400 nm and thickness θ = 35 nm, with a mid-transition critical temperature TC = 89 K [25,30], a transition width of 1.2 K and a conductivity σN = 3.15 × 105 Ω−1·m−1. Such constrictions were made from YBCO films sputtered on MgO (001) single crystalline substrates. X-ray diffraction confirmed the mainly c-axis orientation of the films, with rocking curves exhibiting full width at half maximum values of ~ 0.18 degree for the (005) YBCO line. Weak parasitic lines were observed however, which indexation could also confirm the presence of a-axis growth [16]. The measured c-axis lattice parameter values ranged from 1.169 to 1.170 nm, which are close to the value of fully oxygenated Y1Ba2Cu3O7 films [42]. Besides, in-plane twinning of the films is expected [43]. The film surface morphology was uniform with a typical roughness of about 4.5 nm rms independent of the YBCO thickness [44].
For a YBCO constriction exhibiting such dimensions, our simulation results lead to: AT = 0.011 THz−1, BT = 0.186 THz−1 and CT = 6.82 × 10−8 Ω·m·THz−1. Those results are illustrated in Figure 3.

3.2. DC Characteristics

For our simulations, we have considered two HEB devices (data gathered in Table 1). The constriction dimensions of device A are close to the limits achievable with e-beam lithography; these dimensions are those considered in the "0D" HEB model [22]. The constriction dimensions of device B are those achievable with our currently available technological process, combining optical and e-beam lithography [25]. Device B is also the constriction example taken in Section 3.1.

3.2.1. Temperature Profiles

The temperature profiles (solutions of Equation (4a,b)), shown in Figure 4 for devices A and B exhibit features, typically representing the HEB phenomena, namely: (i) The hot electron condition Te(x) > Tp(x), and (ii) the hot spot (i.e., normal state) region ΔxHS such that Te(x) > TC for x ∈ ΔxHS. We notice that Te(x) is closer to Tp(x) for device B; this is due to the stronger phonon film to substrate escape efficiency, as testified by the ~3.5 times longer τesc value for device B (Table 1).
As a check of the validity of our simulation approach, we run our procedure with low-TC material data for the NbN constriction considered in Reference [24]. We obtained Te(x) profiles very close to each other (less than 5% difference). This was verified for various PLO values in the quasi-static (QS) regime, i.e., at the low ωLO frequency limit of the constriction impedance (Equation (8)).

3.2.2. Current-Voltage Plots

The Te(x) profile allows to have access to the resistivity profile along the constriction, hence to the constriction resistance/impedance by integration (Equation (8)). As explained in Section 2.3, the I-V plots can then be deduced in the form of "clouds of points". This representation illustrates, in color/gray tone levels, the device DC response as a function of e.g., PLO, as shown in Figure 5 for devices A and B.
With regard to device A, we notice that the I-V dependence can be hysteretic: For a given bias current value, there are three distinct HEB voltage operating points. Indeed, in the 5 mV to ~ 15 mV range, the differential resistance is negative, this behavior being typical of HEBs operating at low T0 and exhibiting a short film to substrate escape time. Consequently, the largest voltage variations, as PLO varies, where a larger part of the constriction is still superconducting, are situated below ~ 30 mV. This effect is not observed for device B, due to the longer escape time in that latter case.
For further verification, we also compared, for an NbN device, our I-V results with those reported in Reference [24]. As shown in Figure 6, the calculated values are close to the HEB measurements at bias currents above 20 μA. The result of our calculation for PLO = 220 nW is, however, noticeably different from that reported in Reference [24]. This is seemingly due to the difference in the computational method, but this difference does not affect the calculation of the dynamic performances (conversion gain and noise temperature) of the HEB.

3.3. Mixer Performance

The noise temperature and the conversion gain were determined at given local oscillator power PLO; they were deduced from the heat equations solving results and presented in the form of “clouds of points” (see Section 2.3). To highlight the influence of PLO for a given DC bias power PDC, we have illustrated those results as PDC vs. PLO maps. We have considered four operating frequencies: The quasi-static (QS) regime, 500 GHz, 2.5 THz and 4 THz. The results are presented in the following.

3.3.1. Noise Temperature

We observe in Figure 7 that the minimum noise PDC vs. PLO area evolves in level and that the optimal PLO value also evolves (see Table 2 below). In particular, we notice that the secondary area of low noise temperature, starting at PLO ≈ 35 μW in the QS mode, vanishes when the frequency increases.
This consideration can be furthered by separating the sources of noise for the two cases of QS regime and LO frequency at 2.5 THz (Figure 8). The contribution of each noise source informs us about the evolution of the area where the noise is minimum. We notice that in the QS regime, the fluctuation noise is minimal when PLO > 20 μW, whereas at 2.5 THz the low noise area is distributed differently, with improved noise level when 10 μW < PLO < 15 μW.

3.3.2. Conversion Gain and Summary of Mixer Results

We then resume the comparison for the conversion gain at the four working frequencies, as shown below in Figure 9. The gain decreases as PDC increases, and is maximized at given PLO. Thus, as the frequency of the local oscillator increases, the conversion gain appears to be maximum around 2.5 THz and requires a local oscillator power of 12.5 μW (see Table 2 below).
As compared to previous results [25], the present model highlights a sharp decrease in the local oscillator power needed to operate the HEB at the optimum point. Intuitively, one might be inclined to think that after taking into account αimp coupling losses, PLO should increase as 1/αimp. In fact, the present model introduces two physical properties, namely the influence of the non-uniform distribution of the dissipated LO power, and the impedance matching coefficient between the antenna and the nano-constriction, which alter the response of the simulated HEB.
Indeed, the dissipation of the radiofrequency current is concentrated in the resistive zone of the nano-constriction, where dominates the dissipation by Joule effect (hot spot). Thus, if for example a quarter of the nano-constriction is resistive, then the power will be mainly dissipated in this zone, thus reducing a prerequisite of 35 μW (calculated in the model of the "regular" hot spot adapted to YBCO [25]) down to 7.5 μW.

3.3.3. IF Bandwidth

Small signal analysis (as formulated in Reference [25]) was performed to determine the IF response in terms of conversion gain as a function of the intermediate frequency fIF = (ωLOωS)/2π, as originating from the last term in Equation (5). This is illustrated in Figure 10 for devices A and B. The main features are: (i) The low frequency regular bolometric response plateau, (ii) the regular bolometric cutoff associated with the escape time τesc, (iii) a second plateau characterizing the HEB action, (iv) the HEB cutoff associated to the YBCO intrinsic relaxation time τep. Clearly, the bandwidth is limited by the τesc value (Table 1), which is in favor of ultrathin and small volume constrictions [23,25]. It therefore appears that a bandwidth close to 1 GHz can be expected for device A.

3.4. Standoff Detection Performances Requirements

We have run the simulation for the standoff system, as presented in Section 2.5 (Equations (18)–(20)), with detector double sideband width 2Δν = 2 GHz and integration bandwidth δνint = 1 Hz. Five frequencies centered in atmospheric transmission windows were considered: 670 GHz, 1.024 THz, 1.498 THz, 2.522 THz, and 3.436 THz. Besides, the detection scenario was chosen according to the following parameters:
  • Target emissivity εT = 1.
  • Optical losses were evaluated according to the antenna main lobe efficiency (−2 dB), the focusing lens and the detector cryostat window losses (−1 dB), and the LO injection losses (beam splitter: −3 dB), amounting to topt ≅ 24%.
  • tatm was determined for various relative humidity values (RH, in the 10% to 70% range) at operating temperature Top = 295 K, at sea level and with clear atmospheric conditions (e.g., no dust).
  • tobs was calculated for light cotton clothing from data in Reference [46] for frequencies below 1 THz, and in Reference [47] for frequencies above 1 THz. Values of tobs for the five selected frequencies are collected in Table 3.
Various values of temperature resolution ΔT were taken into account, from 0.5 K to 10 K, with respect to the results obtained as a function of the pertinent parameters (frequency, target distance).
Figure 11a exhibits the required HEB mixer noise temperature for a passive mode detection system as a function of frequency and atmospheric relative humidity level for a 10 m target distance and a resolution ΔT = 1 K. The influence of relative humidity strongly occurs at frequencies higher than 2.5 THz (for instance, a low value of TDSB ≈ 85 K is required for a system operating at 2.5 THz at RH level of 40 %). In order to reach a viable standoff passive detection operation with TDSB values technologically reachable, it is therefore necessary to reduce either the target distance or the temperature resolution.
Figure 11b exhibits the required noise temperature as a function of target distance and resolution temperature at 2.5 THz and highly stringent RH level of 70%. Thus, for instance, at 5 m target distance, the detection system would require TDSB ≈ 110 K for ΔT = 1 K, a restrictive parameter which could be relaxed to TDSB approaching 1100 K for ΔT = 10 K.

4. Discussion

In this discussion, we focus on a limited number of points specific to our present modeling approach, where it differs from either our previous modeling hypotheses or previously published models. We wish also to compare between our simulations and some experimental results, unfortunately very few in the case of YBCO HEBs.

4.1. Inhomogeneous PLO Hypothesis Effect on I-V Plots at Low DC Voltage

Our model introduces the matching factor αimp between the terahertz antenna and the nano-constriction, of impedance RRF + jXRF. This impedance is very small at low PDC and PLO power levels (say < 1 μW). In that case, αimp ≅ 0 and therefore the contribution of αPLO in the heat equations seems to be negligible: If it were true, the I-V plots would be very similar at various PLO levels.
In Figure 12, however, we observe two very different distributions between the I-V plots, according to PLO values. The reason is related to the method of calculation, as considered now.
  • In the conventional hot spot method (homogeneous PLO dissipation), the LO power expression, including impedance matching, is written as:
    α P LO = α 4 R a R ( R a + R ) 2 P LO ,
    where R is the constriction resistance, Ra is the antenna resistance, PLO is the power actually applied prior to losses, and is deduced from αPLO which is the power used for the hot spot calculation. When R → 0, αPLO → 0 at all PLO values (Figure 12a).
  • In the hot spot method with RF current (inhomogeneous PLO dissipation), the terahertz power is determined from the intensity of the terahertz current, as outlined in Section 2.3 (Equations (6)–(10)). After some manipulations, still with the constriction low impedance value hypothesis, the following ILO vs. PLO relationship can be worked out [27]:
    I LO α 8 P LO R a 1 + r 2 .
    Even at low PDC and PLO levels, we observe there is a non-zero ILOPLO1/2 along the constriction in the superconducting state. If PLO is strong enough, the critical temperature TC(ILO) < T0 and RRF is no longer negligible. Consequently, the PLO dissipation causes the increase of the electron temperature and therefore a non-zero resistance even at low current (Figure 12b).
Figure 12a illustrates how impedance matching is not correctly formulated with conventional HEB models, in which powers of both DC and RF origins are confounded. The resulting DC plots (the I-V curves seem to be merging) are observed neither in the NbN nor YBCO-based HEB measurements. PDC and αimpPLO are quite distinct in the RF current hot spot model, which makes it possible to obtain results consistent with the typical HEB I-V measurements (Figure 12b).
Another verification has been possible through DC I-V plots simulated at 650 GHz and compared to measurements performed on a YBCO device (L = w = 300 nm, θ = 30 nm) made available to us [48,49], and further considered in Reference [50]. Although not entirely quantitative, the constriction pumping effect by the LO power is clearly visible in Figure 13. We also observe a better similarity between those results and the I-V curves in Figure 12b, rather than those in Figure 12a, so in favor of our complete model with RF current.

4.2. Taking the Resistivity Limit in the Superconducting Transition into Account

The limits of the THz superconducting transition model appear when the normal state resistivity is very high (because YBCO tends to deteriorate with aging [16]); in this case, the effects of the working frequency are of little significance. It is possible to calculate an order of magnitude of this resistivity limit by considering the ρ2F(ω,tr → 0)/ρN > 1% criterion; starting from Equation (1a,b), we can work out:
ρ 2 F ( ω , t r 0 ) ρ N ω μ 0 λ 0 3 .
According to the limit criterion and Equation (21c) we obtain:
ρ 2 F ( f THz , t r 0 ) ρ N f THz × 1.05 × 10 7 ρ N > 1 % .
The resistive transition model should therefore be used at frequencies such that fTHz > ρN × 9.5 × 104, which is the case of our study, where the typical normal state resistivity is ρN = 3.2 × 10−6 Ω·m and our calculation frequency fTHz ≥ 0.3 THz.
Besides, it is possible to carry out the calculation further, because our two-fluid dispersive model also makes it possible to express the imaginary part of the resistivity ρ22F (kinetic inductance effect). A calculation result of the constriction impedance RRF(Te) + jXRF (Te) is shown in Figure 14. We notice that the reactive part reaches its maximum near the middle of the resistive transition, where the detectors are considered the most sensitive, because dRRF/dTe is optimal.
However, there is no obvious function to represent the variations of ρ22F (ω,Te). It is therefore necessary to know the YBCO temperature distribution along the constriction and solve the complete convolution calculation to estimate ρ22F and to deduce the imaginary part of the constriction impedance. We therefore opted for this procedure as part of the hot spot model as presented above.

4.3. Mixer Noise: Effect of Taking the Impedance Matching Factor into Account

The best operation area at PLO ≈ 10 μW, as observed in Section 3.3.1, was not present in our previous model without impedance matching [25]; we therefore studied the influence of taking into account the impedance matching factor between the antenna and the nano-constriction on the noise temperature.
As already mentioned, our approach shows that there may be cases where the loss of coupling efficiency, due to impedance mismatch, may increase the sensitivity of the device. This influence is more evident by considering the distribution of calculated noise temperatures with non-constant αimp (shown previously in Figure 7) and compare it to the distribution of noise temperatures with constant αimp (= 1), as shown below (Figure 15). We notice a significant change in the position of the better operation area. In the quasi-static mode, the best operation area at PLO ≈ 10 μW is no longer present and, in general, the noise temperatures are higher when the impedance matching is not taken into account (see Table 4).
In this discussion, where the influence of impedance matching is removed, the noise distribution as a function of frequency is only subject to change according to the distribution of PLO along the nano-constriction. We clearly observe (Table 4): (i) A decrease in the prerequisite of PLO, and (ii) an improvement of the noise temperature when the operating frequency increases. This is due to the influence of the distribution of PLO along the nano-constriction which is in question: This distribution is more localized in the quasi-static regime than at 4 THz, to take the extreme cases.
As part of a model to optimize the manufacturing of future HEBs, this result demonstrates the importance of taking impedance matching into account. The parameters to consider for manufacturing an HEB are the dimensions of the nano-constriction and the quality of the film, but also the resistive characteristic of the antenna, whose appropriate adjustment of the impedance should allow a gain in performance.

4.4. Comparing Simulated Mixer Noise with Experimental Results

A significant number of HEB heterodyne mixer performance experimental results are gathered in Figure 16.
The state-of-the-art HEB mixers are made from epitaxially grown low-TC superconducting ultrathin films: Nb HEBs mainly below 1 THz, and (100) NbN HEBs above 1.4 THz. For example, values of TDSB = 600 K at 2.5 THz, with a large noise bandwidth in excess of 7 GHz [51], and TDSB = 815 K at 4.7 THz [52] were reported for NbN HEBs. However, TC values for these HEB devices remain below 11 K. An interesting alternative can be offered with HEBs made from c-axis oriented MgB2 ultrathin films [20,53], due to increased operation temperature (TC = 39 K) and expected instantaneous bandwidth up to 10 GHz.
We notice the small number of published data concerning YBCO devices. As already mentioned in our introduction, this situation can be attributed to YBCO nanostructuration challenges, therefore a performance degradation associated with aging effects, even on highly c-axis oriented YBCO ultrathin films. Among these, the result at 585 GHz reported in Reference [15] concerns a rather large constriction (L = 1 µm, w = 2 µm, θ = 100 nm) exhibiting TDSB = 3600 K and G = − 11 dB at PLO = 1 μW. It may be compared to our performance prediction for a device B (Table 1) under not optimized conditions at 750 GHz (TDSB = 3000 K and G = − 10.8 dB at PLO = 9 μW [27]). Besides, the very large TDSB values reported in Reference [13] were attributed to inadequate experimental conditions by the authors, who predicted an improvement of TDSB < 104 K up to 6 THz (system noise temperature), a figure compatible with our predicted TDSB = 5 × 103 K (mixer noise temperature, not optimized [27]) at 2.5 THz. Our present optimized predictions (solid curve in Figure 16) have therefore to be taken as a limit that could be reached with improved nano-structuring of c-axis oriented YBCO ultrathin films [54,55].
As a corollary remark, we notice that the authors in Reference [24] applied a factor 0.3 to KRF for an optimal fit between their noise simulations and measurements on NbN HEBs. After having applied this same correction in our model, we also checked a very good (and even better) fit with the noise temperature measurements reported in Reference [24]. In fact, our results benefited from computational software improvements: Fewer errors accumulated when solving the differential equations, interpolation between points to obtain KDC, and smoothing algorithms. The physical explanation for this overestimation of the resistance variation induced by the RF power remains, however, uncertain for the authors in Reference [24], and ourselves. These considerations nevertheless encouraged us to adapt the NbN HEB hot spot model to YBCO HEBs. Clearly the adjustment of the KRF correction coefficient should require a comparative and systematic study with YBCO HEB mixer experimental tests, which is a potential point of progress in the present approach.
We finally comment on the availability of LO sources. Below 1 THz, the solid state based electronic sources (frequency multiplier chains) offer compactness, tunability and room temperature operation with, e.g., delivered PLO ≈ 150 μW at 750 GHz (see Reference [57] for instance). Up to 2.5 THz, another room temperature alternative relies on tunable backward wave oscillators (BWO). However, the operation of BWOs requires heavy weight magnets, making the BWOs bulky; they deliver, e.g., PLO ≈ 25 μW at 2 THz (see Reference [58] for instance). At and above 2.5 THz, the LO power can be obtained from optoelectronic sources, such as compact mm-size THz quantum cascade lasers (QCL) operating at 70 K, with PLO in the 50 μW to 1 mW range [59,60]. The temperature operation of 70 K, matching the HEB operation, makes it possible to integrate both components within similar cryogenic systems.

4.5. Standoff Detection Limits with YBCO HEB Mixers

We are discussing here the results on standoff detection requirements (Section 3.4) in the light of achievable noise performance (Section 3.3 and Section 4.3). Figure 17a summarizes results presented in Section 3.4 by showing the dependence on the target distance of the standoff detection system in terms of required mixer noise temperature. Constraints of a high relative humidity level and transmission through a cotton tissue have been considered to match a realistic scenario, thus leading to reduce temperature resolution (i.e., increasing ΔT values) at higher frequencies. The higher dT is expected, the lower TDSB floor of the receiver system is required.
This latter point is emphasized in Figure 17b, which exhibits the dT vs. ΔT relationship illustrating a constant TDSB value. TDSB values were chosen at 1000 K and at 2000 K, corresponding to an optimal operating point (in terms of local oscillator power in particular) for YBCO device B, as shown in Table 2. For instance, with a 2.5 THz receiver exhibiting TDSB = 2000 K, we could expect dT ≈ 3.5 m at ΔT = 10 K.
However, if the HEB mixer operates at a non-optimal point, e.g., due to a low available LO power (e.g., PLO = 9 µW), the mixer noise temperature increases (see values of TDBS-MIN in Table 5, extracted from Reference [27]). In this context of non-optimal operating point, Table 5 presents various potential scenarios for accommodating the HEB mixer performance at three frequencies. Calculations were also performed by relaxing the obstacle attenuation constraint. As an example, for a 2.5 THz receiver exhibiting TDSB = 4150 K, we should expect dT ≈ 1.6 m at ΔT = 10 K and RH = 70%. If the obstacle transmission parameter is relaxed (tobs = 1), the achievable target distance increases: For the same receiver, we could expect dT ≈ 11.5 m at ΔT = 10 K and RH = 70%.
As a final comment, we compare the chosen instantaneous bandwidth Δν = 1 GHz to the expected HEB IF cutoff related to the phonon escape time (Figure 10, Section 3.3.3). This choice is close to the simulated bandwidth for device A. For device B, however, the expected bandwidth approaches ~200 MHz, which falls short of the standoff mixer bandwidth parameter. However, it should be emphasized that measurements of YBCO-based devices have exhibited several GHz bandwidth values (e.g., 7 GHz [15]). This discrepancy could be ascribed to a vortex flow mechanism, also invoked to interpret YBCO detectors response, which would superimpose to the regular HEB action, leading to an improved bandwidth value [12].

5. Conclusions and Future Plans

Our first contribution in this paper was to provide a physical description of the superconducting transition enabling to introduce the YBCO THz conductivity in the HEB equations solving loop. This description allowed us to have access to the full impedance of the nano-constriction as a function of temperature, highlighting in particular: (i) The shift and broadening of the superconducting transition as the THz frequency increases, (ii) a residual resistance in the low temperature limit. The knowledge of this impedance also allowed to introduce the THz power matching factor with the antenna, thus leading to a relevant discussion on the influence of the LO power level to optimize the HEB performance.
Our second contribution was to approach in a realistic way the LO power dissipation in the constriction, by making the assumption of a uniform current along this constriction, as delivered by the antenna. The resulting non-uniform LO power dissipation along the constriction, in the frame of a hot spot model, has a clear influence on the shape of the DC I-V plots that gave us the opportunity to check our simulations with published measurement results.
Our third contribution was to simulate the gain, noise temperature and IF bandwidth of an HEB THz mixer. The presentation of the results in the form of level maps in PDC vs. PLO coordinates made it possible to clearly illustrate the optimal operating conditions as a function of the THz frequency for a YBCO HEB. For instance, it was shown that TDSB ≈ 1000 K could ultimately be achieved at PLO ≈ 13 μW for a constriction of dimensions compatible with our technological process. The IF instantaneous bandwidth was shown to lie below 1 GHz for a 10 nm thick YBCO constriction, and to be roughly inversely proportional to the thickness. Some comparisons with the few available published results could be made. It has been noted that some flux flow effect could be involved in the detection mechanism and increase the bandwidth, which is a point to be further considered. These models will also provide a useful guide to refine our ongoing HEB fabrication process.
Our fourth contribution consisted in evaluating the performances required to implement a YBCO HEB in a standoff detection system in passive mode. The knowledge of the optimized performances of the HEB made it possible to discuss the conditions of detection (at the diffraction limit) according to the parameters of the standoff system (distance to the target, absorption and atmospheric humidity, temperature resolution, obstacles). The impact of non-optimized conditions (with respect to PLO level) was also discussed. Typically, detection at 3 m through cotton cloth could be readily achieved in moderate humidity conditions with 10 K target temperature resolution.
This set of modeling studies will support ongoing new imager design and evaluation. Future work will include the development of a 2.5 THz coherent detection system prototype, based on a YBCO HEB mixer as a heterodyne detector and a QCL source as a local oscillator, each device being implemented in its own cryostat or, even better, in its own miniaturized cryo-generator [61]. The main challenges concern: (i) The production of high quality c-axis oriented YBCO ultrathin films and antenna-coupled nano-constrictions made from these films, which are particularly sensitive to technological steps; (ii) the implementation of a THz LO source of sufficient power, while limiting its thermal dissipation; (iii) the development of a low noise/broadband cooled IF amplifier chain; (iv) the implementation of optical components in the cryostats; (v) the consideration of an adequate strategy for the imaging system with YBCO HEB mixers, e.g., scanning vs. staring array options.

Author Contributions

Conceptualization, R.L., A.D., V.J. and A.K.; methodology, R.L.; formal analysis, R.L. and V.J.; investigation, R.L., A.D., V.J. and A.K.; visualization, R.L., A.K. and A.D.; writing—original draft preparation, A.K.; writing—review and editing, A.K., A.D. and V.J.; supervision, A.K.; project administration, A.D. and A.K.; funding acquisition, A.D.

Funding

This research was part of the MASTHER project, funded by the French National Research Agency under grant # 2011-BS03-008-01.

Acknowledgments

The authors wish to thank Dr. Juliane Raasch (KIT, Karlsruhe, Germany) to have kindly provided measurement results on YBCO nano-constrictions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gershenzon, E.M.; Gol’tsman, G.N.; Gousev, Y.P.; Elant’ev, A.I.; Semenov, A.D. Electromagnetic radiation mixer based on electron heating in resistive state of superconductive Nb and YBCO films. IEEE Trans. Magn. 1991, 27, 1317–1320. [Google Scholar] [CrossRef]
  2. Shukarov, A.; Lobanov, Y.; Gol’tsman, G. Superconducting hot-electron bolometer: From the discovery of hot-electron phenomena to practical applications. Supercond. Sci. Technol. 2016, 29, 023001. [Google Scholar] [CrossRef]
  3. Karasik, B.S.; McGrath, W.R.; Wyss, R.A. Optimal Choice of Material for HEB Superconducting Mixers. IEEE Trans. Appl. Supercond. 1999, 9, 4213–4216. [Google Scholar] [CrossRef]
  4. Karasik, B.S.; Sergeev, A.V.; Prober, D.E. Nanobolometers for THz Photon Detection. IEEE Trans. Terahertz Sci. Technol. 2011, 1, 97–111. [Google Scholar] [CrossRef]
  5. Maezawa, H. Application of Superconducting Hot-Electron Bolometer Mixers for Terahertz-Band Astronomy. IEICE Trans. Electron. 2015, E98-C, 196–206. [Google Scholar] [CrossRef]
  6. Lindgren, M.; Currie, M.; Williams, C.; Hsiang, T.Y.; Fauchet, P.M.; Sobolewski, R. Intrinsic picosecond response time of Y-Ba-Cu-O superconducting photodetectors. Appl. Phys. Lett. 1999, 74, 853–855. [Google Scholar] [CrossRef]
  7. Garcés-Chávez, V.; De Luca, A.; Gaugue, A.; Georges, P.; Kreisler, A.; Brun, A. Temporal response of high-temperature superconductor thin-film using femtosecond pulses. In Proceedings of the 4th European Conference on Applied Superconductivity, Sitges, Spain, 14–17 September 1999; Obradors, X., Ed.; Institute of Physics: Bristol, UK, 2000; pp. 93–96. [Google Scholar]
  8. Thoma, P.; Raasch, J.; Scheuring, A.; Hofherr, M.; Il’in, K.; Wünsch, S.; Semenov, A.; Hübers, H.-W.; Judin, V.; Müller, A.-S.; et al. Highly responsive Y-Ba-Cu-O Thin Film THz Detectors With Picosecond Time Resolution. IEEE. Trans. Appl. Supercond. 2013, 23, 2400206. [Google Scholar] [CrossRef]
  9. Li, C.T.; Deaver, B.S., Jr.; Lee, M.; Weikle, R.M., II; Rao, R.A.; Eom, C.B. Gain bandwidth and noise characteristics of millimeter-wave YBa2Cu3O7 hot-electron bolometer mixers. Appl. Phys. Lett. 1998, 73, 1727–1929. [Google Scholar] [CrossRef]
  10. Harnack, O.; Beuven, S.; Darula, M.; Kohlstedt, H.; Tarasov, M.; Stephanson, E.; Ivanov, Z. HTS mixers based on the Josephson effect and on the hot-electron bolometric effect. IEEE Trans. Appl. Supercond. 1999, 9, 3765–3768. [Google Scholar] [CrossRef]
  11. Uchida, T.; Yazaki, H.; Yasuoka, Y.; Suzuki, K. Slot antenna coupled YBa2Cu3O7δ hot-electron bolometers for millimeter-wave radiation. Phys. C 2001, 357–360, 1596–1599. [Google Scholar] [CrossRef]
  12. Il’in, K.S.; Siegel, M. Microwave mixing in microbridges made from YBCO thin films. J. Appl. Phys. 2002, 92, 361–369. [Google Scholar] [CrossRef]
  13. Gousev, Y.P.; Semenov, A.D.; Nebosis, R.S.; Pechen, E.V.; Varlashkin, A.V.; Renk, K.F. Broad-band coupling of THz radiation to YBa2Cu3O7δ hot-electron bolometer mixer. Supercond. Sci. Technol. 1996, 9, 779–787. [Google Scholar] [CrossRef]
  14. Cherednichenko, S.; Rönnung, F.; Gol’tsman, G.; Kollberg, E.; Winkler, D. YBa2Cu3O7−δ hot-electron bolometer mixer at 0.6 THz. Proc. Int. Symp. Space Terahertz Technol. 2000, 526–531. [Google Scholar]
  15. Lee, M.; Li, R.C.-T. Wide bandwidth far-infrared mixing using a high-Tc superconducting bolometer. In Proceedings of the 8th International Conference on Terahertz Electronics, Darmstadt, Germany, 28–29 September 2000; pp. 99–101. [Google Scholar]
  16. Kreisler, A.J.; Dégardin, A.F.; Aurino, M.; Péroz, C.; Villegier, J.-C.; Beaudin, G.; Delorme, Y.; Redon, M.; Sentz, A. New trend in terahertz detection: High-Tc superconducting hot electron bolometer technology may exhibit advantages vs. low-Tc devices. In Proceedings of the 2007 IEEE/MTT-S International Microwave Symposium, Honolulu, HI, USA, 3–8 June 2007; pp. 345–348. [Google Scholar] [CrossRef]
  17. Harrabi, K.; Cheenne, N.; Chibane, F.; Boyer, F.; Delord, P.; Ladan, F.-R.; Maneval, J.-P. Thermal boundary resistance of YBa2Cu3O7 on MgO films deduced from the transient V(I) response. Supercond. Sci. Technol. 2000, 13, 1222–1226. [Google Scholar] [CrossRef]
  18. Ladan, F.-R.; Harrabi, K.; Rosticher, M.; Mathieu, P.; Maneval, J.-P.; Villard, C. Current-temperature diagram of resistive states in long superconducting niobium filaments. J. Low Temp. Phys. 2008, 153, 103–122. [Google Scholar] [CrossRef]
  19. Harnack, O.; Il’in, K.S.; Siegel, M.; Karasik, B.S.; McGrath, W.R.; de Lange, G. Dynamics of the response to microwave radiation in YBa2Cu3O7x hot-electron bolometer mixers. Appl. Phys. Lett. 2001, 79, 1906–1908. [Google Scholar] [CrossRef]
  20. Cunnane, D.; Kawamura, J.H.; Wolak, M.A.; Acharya, N.; Xi, X.X.; Karasik, B.S. Optimization of Parameters of MgB2 Hot-Electron Bolometers. IEEE Trans. Appl. Supercond. 2017, 27, 2300405. [Google Scholar] [CrossRef]
  21. Novoselov, E.; Cherednichenko, S. Broadband MgB2 Hot-Electron Bolometer Mixers Operating up to 20 K. IEEE Trans. Appl. Supercond. 2017, 27, 2300504. [Google Scholar] [CrossRef]
  22. Karasik, B.S.; McGrath, W.R.; Gaidis, M.C. Analysis of a high-Tc hot-electron superconducting mixer for terahertz applications. J. Appl. Phys. 1997, 81, 1581–1589. [Google Scholar] [CrossRef]
  23. Adam, A.; Gaugue, A.; Ulysse, C.; Kreisler, A.; Boulanger, C. Three-temperature model for hot electron superconducting bolometers based on high-Tc superconductor for terahertz applications. IEEE Trans. Appl. Supercond. 2003, 13, 155–159. [Google Scholar] [CrossRef]
  24. Khosropanah, P.; Merkel, H.; Yngvesson, S.; Adam, A.; Cherednichenko, S.; Kollberg, E. A distributed device model for phonon-cooled HEB mixers predicting IV characteristics, gain, noise and IF bandwidth. In Proceedings of the 11th International Symposium on Space Terahertz Technology, University of Michigan, Ann Arbor, MI, USA, 1–3 May 2000; pp. 474–488. [Google Scholar]
  25. Ladret, R.G.; Dégardin, A.F.; Kreisler, A.J. Nanopatterning and hot spot modeling of YBCO ultrathin film constrictions for THz mixers. IEEE Trans. Appl. Supercond. 2013, 23, 2300305. [Google Scholar] [CrossRef]
  26. Ladret, R.G.; Kreisler, A.J.; Dégardin, A.F. High-TC micro and nano-constrictions modeling: Hot spot approach for DC characteristics and HEB THz mixer performance. J. Phys. Conf. Ser. 2014, 507, 042020. [Google Scholar] [CrossRef]
  27. Ladret, R.G.; Kreisler, A.J.; Dégardin, A.F. YBCO-Constriction Hot Spot Modeling: DC and RF Descriptions for HEB THz Mixer Noise Temperature and Conversion Gain. IEEE Trans. Appl. Supercond. 2015, 25, 2300505. [Google Scholar] [CrossRef]
  28. Ma, J.-G.; Wolff, I. Modeling the microwave properties of superconductors. IEEE Trans. Microw. Theory Technol. 1995, 43, 1053–1059. [Google Scholar] [CrossRef]
  29. Ma, J.-G.; Wolff, I. Electromagnetics in high-Tc superconductors. IEEE Trans. Microw. Theory Technol. 1996, 44, 537–542. [Google Scholar] [CrossRef]
  30. Péroz, C.; Villégier, J.-C.; Dégardin, A.F.; Guillet, B.; Kreisler, A.J. High critical current densities observed in PrBa2Cu3O7−δ/YBa2Cu3O7−δ/PrBa2Cu3O7−δ ultrathin film constrictions. Appl. Phys. Lett. 2006, 89, 142502. [Google Scholar] [CrossRef]
  31. Gao, J.-R.; Yagoubov, P.A.; Klapwijk, T.M.; de Korte, P.A.J.; Ganzevles, W.F.; Floet, D.W. Development of THz Nb diffusion-cooled hot electron bolometer mixers. Proc. SPIE Conf. 2003, 4855, 371–382. [Google Scholar] [CrossRef]
  32. Kollberg, E.L.; Yngvesson, K.S.; Ren, Y.; Zhang, W.; Gao, J.-R. Impedance of hot-electron bolometer mixers at terahertz frequencies. IEEE Trans. Terahertz Sci. Technol. 2011, 1, 383–389. [Google Scholar] [CrossRef]
  33. Miao, W.; Zhang, W.; Zhong, J.Q.; Shi, S.C.; Delorme, Y.; Lefevre, R.; Feret, A.; Vacelet, T. Non-uniform absorption of terahertz radiation on superconducting hot electron bolometer microbridges. Appl. Phys. Lett. 2014, 104, 052605. [Google Scholar] [CrossRef]
  34. Abbott, F.; Dégardin, A.F.; Kreisler, A.J. YBCO thin film sputtering: An efficient way to promote microwave properties. IEEE Trans. Appl. Supercond. 2005, 15, 2907–2910. [Google Scholar] [CrossRef]
  35. Grossman, E.-N. Lithographic antennas for submillimeter and infrared frequencies. In Proceedings of the IEEE International Symposium on Electromagnetic Compatibility, Atlanta, GA, USA, 14–18 August 1995; Volume 14–18, pp. 102–107. [Google Scholar]
  36. Türer, I.; Dégardin, A.F.; Kreisler, A.J. UWB Antennas for CW Terahertz Imaging: Geometry Choice Criteria. In Ultra-Wideband, Short-Pulse Electromagnetics 10; Sabath, F., Mokole, E.L., Eds.; Springer: New York, NY, USA, 2014; pp. 463–472. ISBN 978-1-4614-9499-7. [Google Scholar]
  37. Karasik, B.S.; Elantiev, A.I. Noise temperature limit of a superconducting hot-electron bolometer mixer. Appl. Phys. Lett. 1996, 68, 853–855. [Google Scholar] [CrossRef]
  38. Semenov, A.D.; Hubers, H.-W. Bandwidth of a hot-electron bolometer mixer according to the hot spot model. IEEE Trans. Appl. Supercond. 2001, 11, 196–199. [Google Scholar] [CrossRef]
  39. Rieke, G.H. Visible and infrared coherent receivers. In Detection of Light: From the Ultraviolet to the Submillimeter, 2nd ed.; Cambridge University Press: New York, NY, USA, 2003; pp. 275–301, ISBN-13: 978-0-511-06518-7. [Google Scholar]
  40. Wallace, B. Analysis of RF imaging applications at frequencies over 100 GHz. Appl. Opt. 2010, 49, E38–E47. [Google Scholar] [CrossRef] [PubMed]
  41. Slocum, D.M.; Goyette, T.M.; Slingerland, E.J.; Giles, R.H.; Nixon, W.E. Terahertz atmospheric attenuation and continuum effects. Proc. SPIE Conf. 2013, 8716, 871607. [Google Scholar] [CrossRef]
  42. Beno, M.A.; Soderholm, L.; Capone, D.W., II; Hinks, D.G.; Jorgensen, J.D.; Grace, J.D.; Schuller, I.K.; Segre, C.U.; Zhang, K. Structure of the single-phase high-temperature superconductor YBa2Cu3O7−δ. Appl. Phys. Lett. 1987, 51, 57–59. [Google Scholar] [CrossRef]
  43. Streiffer, S.K.; Zielinski, E.M.; Lairson, B.M.; Bravman, J.C. Thickness dependence of the twin density in YBa2Cu3O7−δ thin films sputtered onto MgO substrates. Appl. Phys. Lett. 1991, 58, 2171–2173. [Google Scholar] [CrossRef]
  44. Kreisler, A.; Dégardin, A.; Guillet, B.; Villégier, J-C.; Chaubet, M. High-Tc Superconducting Hot Electron Bolometers for Terahertz Mixer Applications. In Proceedings of the 9th World Multi-Conference on Systemics, Cybernetics and Informatics, Orlando, FL, USA, 10–13 July 2005; pp. 165–217. [Google Scholar]
  45. Ladret, R.G.; Kreisler, A.J.; Dégardin, A.F. Terahertz superconducting hot electron bolometers: Technological issues and predicted mixer performance for Y-Ba-Cu-O devices. Proc. SPIE Conf. 2014, 9252, 92520R. [Google Scholar] [CrossRef]
  46. Goyette, T.M.; Dickinson, J.C.; Linden, K.J.; Neal, W.R.; Joseph, C.S.; Gorveatt, W.J.; Waldman, J.; Giles, R.; Nixon, W.E. 1.56 Terahertz 2-frames per second standoff imaging. Proc. SPIE Conf. 2008, 6893, 68930J. [Google Scholar] [CrossRef]
  47. Puc, U.; Abina, A.; Rutar, M.; Zidansek, A.; Jeglic, A.; Valusis, G. Terahertz spectroscopic identification of explosive and drug simulants concealed by various hiding techniques. Appl. Opt. 2015, 54, 4495–4502. [Google Scholar] [CrossRef] [PubMed]
  48. Raash, J. Electrical-Field Sensitive YBa2Cu3O7−x Detectors for Real-Time Monitoring of Picosecond THz Pulses. Doktor-Ingenieur Thesis, Karlsruhe Institut für Technologie (KIT), Karlsruhe, Germany, August 2017. Available online: https://publikationen.bibliothek.kit.edu/1000073798 (accessed on 24 January 2019).
  49. Ladret, R. Nano-Mélangeurs Bolométriques Supraconducteurs à Electrons Chauds en Y-Ba-Cu-O Pour Récepteur Térahertz en Mode Passif. PhD Thesis, Université Pierre et Marie Curie, Paris, France, July 2016. Available online: https://tel.archives-ouvertes.fr/tel-01832519/document (accessed on 15 Ianuary 2019). (In French).
  50. Schmid, A.; Raasch, J.; Kuzmin, A.; Steinmann, J.L.; Wuensch, S.; Arndt, M.; Siegel, M.; Müller, A.-S.; Cinque, G.; Frogley, M.D. An Integrated Planar Array of Ultrafast THz Y–Ba–Cu–O Detectors for Spectroscopic Measurements. IEEE Trans. Appl. Supercond. 2017, 27, 2200105. [Google Scholar] [CrossRef]
  51. Tretyakov, I.; Ryabchun, S.; Finkel, M.; Maslennikova, A.; Kaurova, N.; Lobastova, A.; Voronov, B.; Gol’tsman, G. Low noise and wide bandwidth of NbN hot-electron bolometer mixers. Appl. Phys. Lett. 2011, 98, 033507. [Google Scholar] [CrossRef]
  52. Kloosterman, J.L.; Hayton, D.J.; Ren, Y.; Kao, T.-Y.; Hovenier, N.; Gao, J.-R.; Klapwijk, T.M.; Hu, Q.; Walker, C.K.; Reno, J.L. Hot electron bolometer heterodyne receiver with a 4.7-THz quantum cascade laser as a local oscillator. Appl. Phys. Lett. 2013, 102, 011123. [Google Scholar] [CrossRef]
  53. Wolak, M.A.; Acharya, N.; Tan, T.; Cunnane, D.; Karasik, B.S.; Xi, X. Fabrication and Characterization of Ultrathin MgB2 Films for Hot-Electron Bolometer Applications. IEEE Trans. Appl. Supercond. 2015, 25, 7500905. [Google Scholar] [CrossRef]
  54. Arpaia, R.; Ejrnaes, M.; Parlato, L.; Tafuri, F.; Cristiano, R.; Golubev, D.; Sobolewski, R.; Bauch, T.; Lombardi, F.; Pepe, G.P. High-temperature superconducting nanowires for photon detection. Phys. C 2015, 509, 16–21. [Google Scholar] [CrossRef]
  55. Lyatti, M.; Savenko, A.; Poppe, U. Ultra-thin YBa2Cu3O7−x films with high critical current density. Supercond. Sci. Technol. 2016, 29, 065017. [Google Scholar] [CrossRef]
  56. Kreisler, A.; Gaugue, A. Recent progress in HTSC bolometric detectors: From the mid-infrared to the far-infrared (THz) range. Supercond. Sci. Technol. 2000, 13, 1235–1245. [Google Scholar] [CrossRef]
  57. Vadiodes. Available online: http://www.vadiodes.com/en/frequency-multipliers (accessed on 14 January 2019).
  58. Microtech Inst. Available online: http://mtinstruments.com/THz_Generators.html (accessed on 14 January 2019).
  59. Ravarro, M.; Jagtap, V.; Santarelli, G.; Sirtori, C.; Li, L.; Khanna, S.P.; Linfield, E.; Barbieri, S. Continuous- wave coherent imaging with terahertz quantum cascade lasers using electro-optic harmonic sampling. Appl. Phys. Lett. 2013, 102, 091107. [Google Scholar] [CrossRef]
  60. Wang, T.; Liu, J.; Liu, F.; Wang, L.; Zhang, J.; Wang, Z. High-Power Single-Mode Tapered Terahertz Quantum Cascade Lasers. IEEE Photonics Technol. Lett. 2015, 27, 1492–1494. [Google Scholar] [CrossRef]
  61. Masther Research Project Funded by the French National Research Agency under Grant # 2011-BS03-008-01. Information. Available online: http://www.agence-nationale-recherche.fr/suivi-bilan/editions-2013-et-anterieures/recherches-exploratoires-et-emergentes/blanc-generalite-et-contacts/blanc-presentation-synthetique-du-projet/?tx_lwmsuivibilan_pi2%5BCODE%5D=ANR-11-BS03-0008 (accessed on 14 January 2019). (In French).
Figure 1. Superconducting hot electron bolometers (HEB): (a) Illustration of the point bolometer thermal model with two-reservoirs (electrons and phonons of the superconductor); (b) Sketch of a YBCO HEB constriction (of length L, width w, and thickness θ) connected to the arms of a THz planar antenna.
Figure 1. Superconducting hot electron bolometers (HEB): (a) Illustration of the point bolometer thermal model with two-reservoirs (electrons and phonons of the superconductor); (b) Sketch of a YBCO HEB constriction (of length L, width w, and thickness θ) connected to the arms of a THz planar antenna.
Photonics 06 00007 g001
Figure 2. Simplified schematic for HEB standoff THz detection arrangement. The target resolved area is deduced from the Airy pattern at the HEB location (Equation (19)).
Figure 2. Simplified schematic for HEB standoff THz detection arrangement. The target resolved area is deduced from the Airy pattern at the HEB location (Equation (19)).
Photonics 06 00007 g002
Figure 3. Frequency-dependent YBCO superconducting transition illustrated by the constriction resistance vs. frequency. The DC plot (f = 0) is a fit from experiment [30]. The gaussian curves describe the TC distribution as discussed in the text. Fermi-Dirac fits have been used for the device performance simulation as more appropriate to the convergence of numerical solutions. Redrawn after [27].
Figure 3. Frequency-dependent YBCO superconducting transition illustrated by the constriction resistance vs. frequency. The DC plot (f = 0) is a fit from experiment [30]. The gaussian curves describe the TC distribution as discussed in the text. Fermi-Dirac fits have been used for the device performance simulation as more appropriate to the convergence of numerical solutions. Redrawn after [27].
Photonics 06 00007 g003
Figure 4. For devices A and B, electron temperature Te(x) (solid curves) and phonon temperature Tp(x) (dashed curves) profiles: (a) For device A (PLO = 5 μW); (b) for device B (PLO = 35 μW). T0 and TC are the reference (cold finger) and mid-transition critical temperatures, respectively. Arrows on TC lines delimit the hot spot regions ΔxHS (see text). Curve labels indicate the DC bias current I0 values.
Figure 4. For devices A and B, electron temperature Te(x) (solid curves) and phonon temperature Tp(x) (dashed curves) profiles: (a) For device A (PLO = 5 μW); (b) for device B (PLO = 35 μW). T0 and TC are the reference (cold finger) and mid-transition critical temperatures, respectively. Arrows on TC lines delimit the hot spot regions ΔxHS (see text). Curve labels indicate the DC bias current I0 values.
Photonics 06 00007 g004
Figure 5. For devices A and B, DC current vs. DC voltage maps according to PLO values, at 400 GHz LO frequency: (a) For device A (T0 = 60 Κ); (b) for device B (T0 = 70 Κ). Redrawn after [45].
Figure 5. For devices A and B, DC current vs. DC voltage maps according to PLO values, at 400 GHz LO frequency: (a) For device A (T0 = 60 Κ); (b) for device B (T0 = 70 Κ). Redrawn after [45].
Photonics 06 00007 g005
Figure 6. For the NbN HEB reported in Reference [24], the comparison between I-V plots measured on the fabricated device and simulation according to the model developed in Reference [24] and our present model.
Figure 6. For the NbN HEB reported in Reference [24], the comparison between I-V plots measured on the fabricated device and simulation according to the model developed in Reference [24] and our present model.
Photonics 06 00007 g006
Figure 7. For an HEB constriction of dimensions L = w = 400 nm and θ = 35 nm (Figure 1), maps exhibiting double sideband noise temperature TDSB levels in DC bias power vs. LO power coordinates. Impedance matching coefficient αimp with the antenna was included.
Figure 7. For an HEB constriction of dimensions L = w = 400 nm and θ = 35 nm (Figure 1), maps exhibiting double sideband noise temperature TDSB levels in DC bias power vs. LO power coordinates. Impedance matching coefficient αimp with the antenna was included.
Photonics 06 00007 g007
Figure 8. For an HEB constriction of dimensions L = w = 400 nm and θ = 35 nm (Figure 1), maps exhibiting Johnson noise and thermal fluctuation noise contributions to TDSB levels in DC bias power vs. LO power coordinates. Impedance matching coefficient αimp with the antenna was included.
Figure 8. For an HEB constriction of dimensions L = w = 400 nm and θ = 35 nm (Figure 1), maps exhibiting Johnson noise and thermal fluctuation noise contributions to TDSB levels in DC bias power vs. LO power coordinates. Impedance matching coefficient αimp with the antenna was included.
Photonics 06 00007 g008
Figure 9. For an HEB constriction of dimensions L = w = 400 nm and θ = 35 nm (Figure 1), maps exhibiting G levels in DC bias power vs. LO power coordinates. Impedance matching coefficient αimp with the antenna was included.
Figure 9. For an HEB constriction of dimensions L = w = 400 nm and θ = 35 nm (Figure 1), maps exhibiting G levels in DC bias power vs. LO power coordinates. Impedance matching coefficient αimp with the antenna was included.
Photonics 06 00007 g009
Figure 10. For devices A and B, conversion gain vs. fIF, at PLO values minimizing TDSB. Data for device A were available in the QS regime [25], whereas data at 2.5 THz were used for device B [45].
Figure 10. For devices A and B, conversion gain vs. fIF, at PLO values minimizing TDSB. Data for device A were available in the QS regime [25], whereas data at 2.5 THz were used for device B [45].
Photonics 06 00007 g010
Figure 11. Standoff detection double sideband noise temperature requirements: (a) As a function of operating frequency for various atmospheric humidity contents (at fixed ΔT and dT); (b) as a function of target distance for various target temperature resolutions (at fixed ν0 and RH).
Figure 11. Standoff detection double sideband noise temperature requirements: (a) As a function of operating frequency for various atmospheric humidity contents (at fixed ΔT and dT); (b) as a function of target distance for various target temperature resolutions (at fixed ν0 and RH).
Photonics 06 00007 g011
Figure 12. Comparing low-voltage behavior of DC I-V plots taking impedance matching factor with antenna into account: (a) Regular hot-spot model with uniform PLO power dissipation along the constriction; (b) our hot-spot model approach with ILO as an input parameter, i.e., non-uniform PLO power dissipation.
Figure 12. Comparing low-voltage behavior of DC I-V plots taking impedance matching factor with antenna into account: (a) Regular hot-spot model with uniform PLO power dissipation along the constriction; (b) our hot-spot model approach with ILO as an input parameter, i.e., non-uniform PLO power dissipation.
Photonics 06 00007 g012
Figure 13. For a YBCO HEB, I-V map, with PLO levels, from our simulation with 650 GHz RF current (non-uniform LO power dissipation). Measurement points, without and with LO power applied at the same frequency, are also indicated (courtesy J. Raasch [48]).
Figure 13. For a YBCO HEB, I-V map, with PLO levels, from our simulation with 650 GHz RF current (non-uniform LO power dissipation). Measurement points, without and with LO power applied at the same frequency, are also indicated (courtesy J. Raasch [48]).
Photonics 06 00007 g013
Figure 14. Example of the real and imaginary parts of a total THz impedance of a constriction; the temperature was assumed uniform in this case.
Figure 14. Example of the real and imaginary parts of a total THz impedance of a constriction; the temperature was assumed uniform in this case.
Photonics 06 00007 g014
Figure 15. For an HEB constriction of dimensions L = w = 400 nm and θ = 35 nm (Figure 1), maps exhibiting double sideband noise temperature TDSB levels in DC bias power vs. LO power coordinates. Impedance matching coefficient with the antenna was not included (αimp = constant = 1).
Figure 15. For an HEB constriction of dimensions L = w = 400 nm and θ = 35 nm (Figure 1), maps exhibiting double sideband noise temperature TDSB levels in DC bias power vs. LO power coordinates. Impedance matching coefficient with the antenna was not included (αimp = constant = 1).
Photonics 06 00007 g015
Figure 16. For HEB heterodyne receivers or mixers, double sideband noise temperature as a function of operating frequency. DC and PC are for diffusion cooled and phonon cooled devices, respectively. Hot spot (HS) model results are those of Table 2 (optimized) and Reference [27] (fixed LO power). QL: Quantum limit hν/(2kB). Redrawn and updated after [56] and [45].
Figure 16. For HEB heterodyne receivers or mixers, double sideband noise temperature as a function of operating frequency. DC and PC are for diffusion cooled and phonon cooled devices, respectively. Hot spot (HS) model results are those of Table 2 (optimized) and Reference [27] (fixed LO power). QL: Quantum limit hν/(2kB). Redrawn and updated after [56] and [45].
Photonics 06 00007 g016
Figure 17. Standoff detection DSB requirements: (a) Required noise temperature as a function of ν0 for various distances at specified ΔT and fixed humidity; simulated TDSB values for device B are also shown at both optimal PLO conditions - solid curve (Table 2) and at fixed PLO = 9 µW - dashed curve [27]; (b) Required distance vs. temperature difference relationship to achieve TDSB = 1000 K or 2000 K, at fixed RH. Symbols: Computed values, dotted curves: Best fits (according to functions indicated).
Figure 17. Standoff detection DSB requirements: (a) Required noise temperature as a function of ν0 for various distances at specified ΔT and fixed humidity; simulated TDSB values for device B are also shown at both optimal PLO conditions - solid curve (Table 2) and at fixed PLO = 9 µW - dashed curve [27]; (b) Required distance vs. temperature difference relationship to achieve TDSB = 1000 K or 2000 K, at fixed RH. Symbols: Computed values, dotted curves: Best fits (according to functions indicated).
Photonics 06 00007 g017
Table 1. For constrictions A and B, device characteristics and YBCO physical parameters used in the simulations. Unless otherwise stated, the parameter values are those of [22].
Table 1. For constrictions A and B, device characteristics and YBCO physical parameters used in the simulations. Unless otherwise stated, the parameter values are those of [22].
DeviceL, w (nm)θ (nm)Ce (J·m−3·K−1)Cp (J·m−3·K−1)κe (W·m−1·K−1)κp(W·m−1·K−1)
A100102.5×1046.5×105110
B400352.5×1046.5×105110
Deviceτep (ps)τesc (ns)σN1–1·m–1)JC2 (A·cm–2)TC3 (K)ΔT (K)T0 (K)
A1.00.753.15×1052.2×106851.260
B1.72.63.15×1052.2×106891.270
1 Normal state conductivity at ~ 100 K. 2 Critical current density at 77 K. 3 Mid-transition critical temperature.
Table 2. For an HEB of dimensions L = w = 400 nm and θ = 35 nm at different operating frequencies, with matching coefficient αimp included, compared performances at an optimal operating point with respect to the noise temperature.
Table 2. For an HEB of dimensions L = w = 400 nm and θ = 35 nm at different operating frequencies, with matching coefficient αimp included, compared performances at an optimal operating point with respect to the noise temperature.
FrequencyIDC (μA)R (Ω) 1PDC (µW)PLO (μW)G (dB)TDSB (K)
QS45421.34.47.5−9.01781
500 GHz46421.44.67.5−7.11208
2.5 THz39414.62.312.5−6.11013
4 THz37913.01.913.5−6.51093
1 RF resistance.
Table 3. Transmission coefficient through an obstacle of light cotton cloth [46,47].
Table 3. Transmission coefficient through an obstacle of light cotton cloth [46,47].
Frequency (THz)0.671.021.492.523.44
tobs (%)75472120.1
Table 4. For an HEB of dimensions L = w = 400 nm and θ = 35 nm at different operating frequencies, with unit impedance matching coefficient (αimp = 1), compared performances at an optimal operating point with respect to the noise temperature.
Table 4. For an HEB of dimensions L = w = 400 nm and θ = 35 nm at different operating frequencies, with unit impedance matching coefficient (αimp = 1), compared performances at an optimal operating point with respect to the noise temperature.
FrequencyIDC (μA)R (Ω) 1PDC (µW)PLO (μW)G (dB)TDSB (K)
QS24127.91.635−13.82607
500 GHz27126.41.930−13.62704
2.5 THz39219.23.012−10.82332
4 THz38918.22.712.5−10.12021
1 RF resistance.
Table 5. Standoff detection scenarios (with and without obstacle) designed to accommodate YBCO HEB mixer performance (device B) in terms of minimum TDSB at PLO = 9 µW [27].
Table 5. Standoff detection scenarios (with and without obstacle) designed to accommodate YBCO HEB mixer performance (device B) in terms of minimum TDSB at PLO = 9 µW [27].
Frequency (THz)TDSB-MIN (K) [27]ObstacleΔT (K)RH (%)DT (m)NEP (fW/Hz1/2)TDSB (K)
0.672480 1Yes0.570783.12495
0.672480 1No0.57090.83.12485
1.022730Yes1.07019.03.42760
1.022730No1.07026.53.42775
2.524150Yes10303.85.14160
2.524150Yes10701.65.14150
2.524150No107011.55.24245
1 Simulated at 750 GHz.

Share and Cite

MDPI and ACS Style

Ladret, R.; Dégardin, A.; Jagtap, V.; Kreisler, A. THz Mixing with High-TC Hot Electron Bolometers: A Performance Modeling Assessment for Y-Ba-Cu-O Devices. Photonics 2019, 6, 7. https://doi.org/10.3390/photonics6010007

AMA Style

Ladret R, Dégardin A, Jagtap V, Kreisler A. THz Mixing with High-TC Hot Electron Bolometers: A Performance Modeling Assessment for Y-Ba-Cu-O Devices. Photonics. 2019; 6(1):7. https://doi.org/10.3390/photonics6010007

Chicago/Turabian Style

Ladret, Romain, Annick Dégardin, Vishal Jagtap, and Alain Kreisler. 2019. "THz Mixing with High-TC Hot Electron Bolometers: A Performance Modeling Assessment for Y-Ba-Cu-O Devices" Photonics 6, no. 1: 7. https://doi.org/10.3390/photonics6010007

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

Article Metrics

Back to TopTop