The growing interest in hybrid devices that combine two-dimensional materials with a superconductor presents new challenges in material deposition. In this study, we demonstrate that achieving excellent superconducting properties by RF (radio frequency) sputtering does not require access to a high-end system but rather depends on the precise control of sputtering parameters and the selection of an appropriate lithographic process. We highlight the challenges and present practical solutions to deposit high-quality niobium thin films for the lithographic production of superconducting hybrid nanostructures. The influence of various deposition parameters, such as power, argon pressure, and film thickness, on the resultant superconducting characteristics can already be deduced at liquid nitrogen temperatures. Furthermore, niobium films tend to degrade when a PMMA [poly(methylmethacrylate)] resist is employed in the fabrication of superconducting nanostructures. We propose alternative and simple strategies to address this issue, which ultimately result in the restoration of the thin-film quality.

In basic research as well as in applied physics or engineering, there exists an increasing demand to interface low-dimensional systems, such as graphene with superconducting materials for nanostructures or superconducting electrical electrodes.1–5 A typical superconducting material candidate is aluminum. Aluminum has a melting point of only 660 °C and can be deposited with relative ease by physical vapor deposition (PVD) from a heated crucible in a final processing step onto a two-dimensional material. However, the choice of a superconductor depends on the operating conditions of the fabricated, final device, such as experimental temperature T and magnetic field B. While a critical temperature T c close to T can offer high photodetection sensitivity in transition-edge sensors that operate at the boundary between the superconducting and normal states,6,7 for the experimental study of a Josephson junction, T c should exceed T. Superconducting materials with experimentally more convenient critical fields and critical temperatures, such as niobium (Nb), niobium nitride (NbN), niobium titanium nitride (NbTiN), or molybdenum nitride (MoN), can be deposited via molecular beam epitaxy (MBE),8,9 chemical vapor deposition (CVD),10,11 atomic layer deposition (ALD),12,13 or sputtering.14–18 Here, the growth parameters determine the superconducting properties, such as the critical temperature T c and critical magnetic field B c.

Superconducting thin films are widely used in the development of novel nanostructured quantum devices. In many cases, the superconducting material must be deposited as one of several fabrication steps. Sputter deposition, a subtype of PVD, has more compatible growth parameters and is less disruptive to the thin van der Waals material or the resist than most deposition methods due to much lower process temperatures and shorter deposition times. Instead of using precursors or evaporating the base material in a crucible, sputter deposition is based on the ejection of the material from a source target by ion bombardment.19–21 The substrate is often heated to high temperatures to obtain higher adatom mobility and subsequently larger grains and higher thin-film quality.22–24 Analogous to CVD and ALD processes, these temperatures are incompatible with typical lithographic fabrication processes.

Here, we give practical and simple solutions for the sputter deposition of high-quality nanostructured superconducting Nb thin films. Our systematic study highlights that the quality of thin films and their superconducting properties are dictated by the sputter parameters and lithographic preparation, rather than by a high-end deposition system. The introduction of a lithographically defined mask structure by a PMMA resist generally tends to deteriorate the superconducting film quality due to contamination from outgassing. However, we show that relatively simple modifications to the lithographic process can restore thin-film quality. Our approach is aimed at fundamental research with limited access to advanced and high-end growth and characterization facilities. We provide guidelines to produce and optimize high-quality thin films suitable for nanostructuring, even with limited resources, that are transferable to other sputtered superconducting materials and applicable to most 2D materials. Further optimization is required for designing specific nanostructures, such as junctions, interfaces, or electrical contacts.

Nb films are prepared by RF sputtering in a modified Torr International Compact Research Coater System (CRC 600) as schematically shown in the inset of Fig. 1(b). For all sputtering processes, small (0.5 cm 2) pieces of an n-doped silicon wafer with native oxide are ultrasonically cleaned in acetone and isopropanol, followed by mild oxygen plasma treatment to remove any organic contaminants. The plasma treatment was conducted a few days before the deposition to prevent the introduction of reactive oxygen species into the deposition process. The substrates are placed on a rotating metallic sample holder inside the sputtering chamber, which is pumped below 5 × 10 5 Torr prior to the sputter deposition. At ambient temperature, films are deposited from a 2 in. pure Nb target (99.95  %) in an Argon (Ar) atmosphere provided by the flow control unit. After plasma ignition, the chamber pressure and the plasma power are brought to deposition conditions and stabilized for several minutes before the actual deposition. This period serves to clean the Nb target and condition the chamber, which is important to reduce run-to-run variations.28,29

FIG. 1.

(a) Exemplary measurement of the sheet resistance as a function of temperature to determine the critical temperature T c of a 73 nm thick niobium thin film. The inset in shows a schematic of the four-terminal measurement. (b) For fixed sputter parameters (200 W and 6.5 mTorr Ar pressure), the dependence of the T c of the Nb thin films on the layer thickness is displayed. The maximum T c of 7.6 K at 180 nm is lower than for single crystal bulk Nb, which has a T c of 9.25 K.25–27 The green shaded area marks our desired thickness range. The inset shows schematically the configuration of the sputter chamber.

FIG. 1.

(a) Exemplary measurement of the sheet resistance as a function of temperature to determine the critical temperature T c of a 73 nm thick niobium thin film. The inset in shows a schematic of the four-terminal measurement. (b) For fixed sputter parameters (200 W and 6.5 mTorr Ar pressure), the dependence of the T c of the Nb thin films on the layer thickness is displayed. The maximum T c of 7.6 K at 180 nm is lower than for single crystal bulk Nb, which has a T c of 9.25 K.25–27 The green shaded area marks our desired thickness range. The inset shows schematically the configuration of the sputter chamber.

Close modal

The deposition rate is regulated by setting power and argon flow and ranges from 30 to 70 nm/min. Lower deposition rates facilitate manual operation of the shutter to achieve the desired thickness. Deposition rates have been determined by measuring film thickness on test runs with a Bruker Dektak surface profilometer at a step between the bare substrate and the film. No lithographic preprocessing is needed; an area of the substrate is simply covered with a marker pen, lifted off post-deposition to create a film edge, and the thickness is measured through multiple height profiles along this step.

For the fabrication of superconducting Nb nanostructures, we use standard electron beam lithography (EBL) with a PMMA resist (AR-P 672.045, Allresist). In an alternative lithographic process that prevents the contamination by organic polymers through the resist, we use aluminum (Al) instead of a PMMA mask. First, we cover the entire sample with approximately 800 nm of Al using thermal evaporation. The sample is then coated with PMMA, and the desired structure is written by EBL. After development, the alkaline TMAH-based photoresist developer (MF-319, microresist technology) is used to remove Al and impose the pattern of the PMMA into the Al layer. After removing the PMMA, Nb sputter deposition can proceed without any organic polymers on the sample. The finished superconducting Nb structures are obtained after lift-off in the same alkaline developer, which selectively etches the Al.

Nb thin films are electrically characterized in a physical property measurement system (PPMS) DynaCool by Quantum Design from 300 K down to 2 K and up to 9 T using a standard four-terminal measurement scheme [the inset in Fig. 1(a)] to determine the critical temperature, T c, and an upper critical magnetic field, B c, that serve as the figures of merit for the film quality. An exemplary measurement to evaluate the critical temperature of a deposited thin film is presented in Fig. 1(a).

When the thickness of a superconducting film is comparable or less than the coherence length and the superconducting penetration depth, quantum size effects affect the superconducting properties because Cooper pairs cannot form effectively in confined dimensions. The thickness dependence shown in Fig. 1(b) is a combination of increasing grain sizes that exhibit properties closer to the bulk value25,26,30,31 and of stress and strain effects when niobium thin films are grown on substrates such as silicon.32–34 While most reports on sputtered Nb (and NbN films) use 200 nm or larger35,36 to reduce these effects, for our studies, we consider niobium films with thicknesses larger than 60 nm, which is a common thickness for electrical contacts.4 For the Nb film with the largest thickness, the T c of 7.6 K is closest to that of single crystal bulk Nb, which has a T c of 9.25 K.25–27 The deviation from the bulk value could be the result of residual oxygen content in the sputter chamber, which is known to impact the superconducting properties of sputtered Nb films.37 

Each data point in Fig. 2 represents the best result obtained for a fixed P but within a parameter range for the layer thickness (60–100 nm) and argon pressure (3.4–6.6 mTorr). These slight re-adjustments for each P are required due to a complex interdependence of the growth parameters to optimize the films. We found that lower argon pressures (3.4–4.3 mTorr) are required for a lower power to achieve good Nb thin films. The quality and reproducibility of most deposition techniques, including sputter deposition of superconducting Nb, are also significantly influenced by the chamber’s condition and growth history. We will discuss the relevance of these factors on the obtainable T c in more detail later.

FIG. 2.

Power dependence of (a) T c, (b) R 300 K R 15 K, and (c) R 300 K R 77 K for thin films between 60 and 100 nm thickness. Best results were obtained for a power of 200 W (at 6.5 mTorr Ar pressure).

FIG. 2.

Power dependence of (a) T c, (b) R 300 K R 15 K, and (c) R 300 K R 77 K for thin films between 60 and 100 nm thickness. Best results were obtained for a power of 200 W (at 6.5 mTorr Ar pressure).

Close modal

The increase of the RF power in Fig. 2 results in a steady rise of the critical temperature, indicating a continuous improvement in film quality. The cleanest Nb thin films and, thus, the highest T c are achieved at 200 W, followed by a slight decrease at higher powers. The power dependence can be understood via the growth kinetics: at ambient temperature, adatom mobility is rather low and their binding affinity is comparably high, which generally promotes island growth.38 Increasing the kinetic energy of the sputtered particles leads to more uniform films as it allows for a redistribution of adsorbed atoms, thus effectively reducing the likelihood of island growth. At very high powers, however, detrimental effects, such as re-sputtering and amorphization,39,40 begin to deteriorate the quality of the films.

The room temperature resistance R 300 K of the sputtered films is dominated by electron–phonon scattering, while the resistance at low temperatures above T c is dominated by impurity scattering that affects the superconducting properties. The residual resistance ratio (RRR), defined as R 300 K R 15 K [Fig. 2(b)], is commonly used and provides a standardized method to compare different films.14,27 While bulk Nb can have a RRR above 2000, Nb films often display a lower RRR < 10.25–27, Figure 2(a) demonstrates that RRR shares a similar dependence on power with the T c, indicating that R 300 K R 15 K can indeed serve as a quality figure of merit for the thin superconducting films. However, the experimental determination of this ratio requires a cryogenic system, offering no real advantage over directly determining T c. While previous reports on the effect of residual oxygen on Nb films observed a clear dependence between T c and R 300 K,26,37 we found that for our study of RF power dependence, R 300 K alone is not a reliable quality indicator, which could result from other effects beyond residual oxygen, such as grain size. However, its ratio with R 77 K (measured simply in liquid nitrogen) is already sufficient to evaluate the superconducting properties, as shown by Fig. 2(c).14 Hence, in scientific environments where access to liquid helium or cryogenic systems is limited, characterizing the resistance at 77 K can already screen suitable thin films. In the following, we will not focus on RRR but on the more commonly used T c.

Our study is driven by the need to achieve high-quality superconducting thin films in hybrid nanostructures, such as electrical contacts for graphene devices encapsulated in hexagonal boron nitride (hBN).2 Self-aligned one-dimensional edge contacts41 to these devices can be nanofabricated by EBL to define contact areas within a layer of a PMMA resist, followed by reactive ion etching (RIE) to locally expose the graphene layer,42 and subsequent sputter deposition of Nb. Typically, the thickness of metallic contact materials to graphene devices is below 100  nm,41,43,44 thereby also setting our superconducting film thickness range between 60 and 100 nm, as indicated in Fig. 1(b).

With the standard parameters (200 W RF power and 6.5 mTorr Argon), we consistently obtain a critical temperature of approximately 6 K for large lateral thin films, which we now define as pristine. Repeatedly, we also encountered a significantly lower T c for samples that were sputtered with the same parameters as shown in Fig. 3(a). We can pinpoint the deterioration of the superconducting properties to the lithographic process and the history of the chamber. In Fig. 3, we compare the superconducting properties and the resistive behavior of pristine samples to reference samples.

FIG. 3.

(a) Best T c for the same set of deposition parameters and at lowest base pressures p base for frequent sputtering (orange, multiple times per day to daily), less frequent sputtering (red, daily to several times per week), intermittent sputtering (purple, weekly, or less), and reference samples that experienced contamination by a PMMA resist (open circles). (b) Film resistances as a function of temperature for different magnetic fields from 0 to 1.5 T for a pristine Nb thin film compared to a film prepared together with PMMA resist-coated substrates. (c) Measurement of the upper critical field at 2 K for the samples used in subfigure (b).

FIG. 3.

(a) Best T c for the same set of deposition parameters and at lowest base pressures p base for frequent sputtering (orange, multiple times per day to daily), less frequent sputtering (red, daily to several times per week), intermittent sputtering (purple, weekly, or less), and reference samples that experienced contamination by a PMMA resist (open circles). (b) Film resistances as a function of temperature for different magnetic fields from 0 to 1.5 T for a pristine Nb thin film compared to a film prepared together with PMMA resist-coated substrates. (c) Measurement of the upper critical field at 2 K for the samples used in subfigure (b).

Close modal

In materials science, impurities and contamination generally lead to issues, such as reduced mobility in the MBE growth of III/V semiconductors or high resistive, non-linear metallic electrical contacts in PVD. Notably, the observed significant sensitivity of Nb to impurities31 from a PMMA resist during sputtering dramatically impacts the critical temperature, limiting the temperature range for superconducting graphene devices to below the liquid helium temperature of 4.2 K. The reproducibility also depends on the history of the chamber and growth frequency. Despite a cleaning and conditioning procedure described in Section II prior to deposition to remove oxides and most other contaminants from the target surface, Fig. 3(a) illustrates a clear dependence on the frequency of deposition. Best and most consistent results are generally obtained at frequent depositions. Less frequent and intermittent depositions yield significantly lower critical temperatures. As discussed earlier, the niobium films are also very sensitive to residual oxygen levels in the sputter chamber.37 We, thus, suspect that frequent deposition keeps the target conditioned, with minimal oxidation, leading to more consistent results and higher film quality. For less frequent and intermittent depositions, the conducted initial cleaning procedure is not sufficient to yield ideal starting conditions. Our results highlight the importance of a well-conditioned chamber. Conditioning protocols prior to sputter deposition can significantly reduce run-to-run variations.28,29 Heating of the sample stage, which is often included in such conditioning protocols,29 as well as the application of a sample bias voltage to reduce oxygen content in the Nb films37 are not available in our experiments.

Increasing the growth frequency can indeed enhance the superconducting properties. However, addressing issues associated with lithographic processes, which are required for nanostructured graphene devices, for example, are much more challenging. When we compare the results of the pristine sputter depositions to that of PMMA reference samples, we find that those samples always show inferior properties, independent of the growth frequency. Note that the reference samples PMMA were neither coated with PMMA nor nanostructured in any way. The PMMA-labeled samples are, thus, niobium films on clean silicon substrates that were sputtered together with other PMMA-coated substrates.

In Fig. 3(b), the critical temperatures for a pristine (thin dashed lines) and a PMMA sample (thick solid lines) are determined at different magnetic fields. The two blue traces representing 0 T show a significant decrease of approximately 30 % in T c from 5.5 K (pristine) to roughly 3.8 K for the PMMA. While the PMMA resist retains its integrity during the sputter deposition and can be easily removed in a lift-off process, outgassing of PMMA or an interaction of Nb species having high kinetic energies in the plasma with the PMMA leads to enhanced incorporation of impurities in the superconducting thin film. The impurities not only decrease T c but also affect the sensitivity to magnetic fields. The T c of the pristine sample in Fig. 3(b) decreases more in response to the applied magnetic field compared to the T c of the PMMA sample. Nevertheless, the magnetic field dependence in Fig. 3(c) shows a significantly higher upper critical field for the pristine sample. The underlying physical details are complex and not the scope of this paper. However, the different magnetic field dependencies in Fig. 3 are consistent with previous reports and relate to impurity scattering that affects the Cooper pair formation and the interplay of flux pinning at defects under external magnetic fields.45,46

To mitigate the degradation of Nb thin-film quality, in Fig. 4(a), we assess the feasibility of different nanostructuring approaches by comparing the averaged critical temperatures T c , avg and film stress, determined from an ensemble of samples. On average, a pronounced drop of almost 40% is observed in the T c , avg between a pristine sample and sputter processes involving a PMMA resist, displayed as blue and red data points, respectively. However, T c , avg significantly increases when the PMMA resist is hard-baked (PMMA+HB) previous to the Nb deposition (orange data point). This suggests that resist outgassing is responsible for the degradation as additional baking reduces the amount of solvent residues and hardens the resist. We could not observe any further improvement of the superconducting properties when using alternative EBL resists that are advertised as resilient against higher temperatures and plasma processes.

FIG. 4.

(a) Influence of the nanofabrication method on the average critical temperature (closed circles) and the estimated stress from XRD measurements (open circles). (b) XRD results of Nb thin films deposited under different conditions. The curves are vertically shifted for clarity. (c) and (d) Topographic 5 × 5 μm 2 AFM images of sputtered niobium thin films. (c) Pristine Nb film with a RMS roughness of 0.644 nm. (d) PMMA sample with a roughness of 0.491 nm. We interpret the larger surface roughness of the pristine film as a higher level of crystallinity. RMS values for the surface roughness are determined by using the open-source software Gwyddion.

FIG. 4.

(a) Influence of the nanofabrication method on the average critical temperature (closed circles) and the estimated stress from XRD measurements (open circles). (b) XRD results of Nb thin films deposited under different conditions. The curves are vertically shifted for clarity. (c) and (d) Topographic 5 × 5 μm 2 AFM images of sputtered niobium thin films. (c) Pristine Nb film with a RMS roughness of 0.644 nm. (d) PMMA sample with a roughness of 0.491 nm. We interpret the larger surface roughness of the pristine film as a higher level of crystallinity. RMS values for the surface roughness are determined by using the open-source software Gwyddion.

Close modal
An alternative nanostructuring method that avoids a polymer-based resist and contamination introduced during the sputter deposition process is based on a thick layer of aluminum ( 800 nm) as described in Sec. II. The resulting critical temperature (green data point) is closer to that of pristine Nb thin films. Hence, Al presents a promising alternative to polymer-based masks to obtain RF-sputtered nanostructures. Due to the novelty and the complex wet etching behavior resulting from natural oxide formation on the Al surface, this method requires further process optimization tailored to the specific feature sizes and application to achieve high-precision lithography results. While this is beyond the scope of our study, we would like to mention that from first preliminary results on Nb structures obtained by the unoptimized Al-mask approach, we can deduce a lateral resolution of < 350 nm. The edge shape roughness of these structures was twice as rough compared to structures patterned by a PMMA resist. The thin film quality changes can also be assessed through x-ray diffraction (XRD) scans of the Nb thin films, which are plotted in Fig. 4(b). The continuous decrease of the XRD peak amplitude for the Nb (110) direction from pristine to PMMA suggests a reduction of the film crystallinity. This is accompanied by a shift of the peak toward lower diffraction angles, which indicates a change in the lattice constant d. With the calculated lattice constant, it is possible to estimate the strain ε and stress σ of the niobium thin films by using
(1)
and
(2)
with the lattice constant of stress-free bulk niobium d 0 3.3 Å, Young’s modulus E = 104.9 GPa, and Poisson’s ratio ν = 0.397 of niobium.47 From these equations, we estimate values for d and σ and compare them with the averaged T c , avg values for our different nanostructuring approaches in Table I. We observe that even the pristine thin films with the best superconducting properties exhibit tensile stress. Although the stress in sputtered niobium films can be adjusted by modifying the Ar pressure,48 our setup does not maintain stable plasma in that Ar pressure range. The decrease in the critical temperature for PMMA samples in Fig. 4(a) is in agreement with a significant increase in stress due to a larger lattice constant. This effect is reduced for the PMMA+HB and Al-mask samples. The enhanced tensile stress is a clear indicator of reduced thin-film quality, caused by multiple factors, with contamination from a PMMA resist as a dominant one. However, a more detailed and precise analysis of the local surface deformation and the stress distribution in niobium is complex and also requires chambers with enhanced vacuum conditions and a larger Ar pressure range.
TABLE I.

Evaluation of different nanostructuring approaches: Averaged Tc,avg, estimated lattice constant d, and stress σ based on x-ray diffraction patterns of niobium thin films.

SampleTc,avg (K)d (Å)σ (GPa)
Pristine 5.84 ± 0.2 3.33 1.33 
PMMA 3.58 ± 0.26 3.38 3.35 
PMMA+HB 4.21 ± 0.02 3.37 2.88 
Al-mask 5.12 ± 1.06a 3.36 2.38 
SampleTc,avg (K)d (Å)σ (GPa)
Pristine 5.84 ± 0.2 3.33 1.33 
PMMA 3.58 ± 0.26 3.38 3.35 
PMMA+HB 4.21 ± 0.02 3.37 2.88 
Al-mask 5.12 ± 1.06a 3.36 2.38 
a

The error in Fig. 4(a) and Table I represents the range of obtained Tc values. For the Al-mask approach, one of the measured samples exhibited greater divergence, leading to an increased error.

The differences in the crystallinity and morphology of the Nb films can directly be visualized by atomic force microscopy (AFM). Figures 4(c) and 4(d) compare 5 × 5 μm 2 AFM images from a pristine film and a film that was sputtered together with a PMMA-coated substrate. While the topographic images reveal dense and homogeneous films for both samples, visibly larger grains and a higher root mean square (RMS) are found in the pristine sample. We attribute the larger surface roughness of the pristine film to larger crystallites and, thus, a higher level of crystallinity,32 which is in agreement with our resistance and XRD measurements.

Achieving high-quality superconducting Nb thin films through sputter deposition can be challenging, especially when incorporated into larger nanostructuring schemes for two-dimensional materials. Our study focused on preserving the quality and superconducting properties of RF-sputtered Nb thin films, intended for the application in nanostructuring processes that involve 2D materials such as graphene. The feature sizes and demands for specific superconducting nanostructures, such as electrical contacts or small junctions, can vary considerably and will require further adjustments of the process parameters. However, we demonstrated that advanced growth and characterization facilities are not required to obtain competitive films, and we provided guidelines that are transferable to other superconducting materials and 2D material systems and hybrid superconducting devices.

Here, we investigated the electrical and structural properties of RF-sputtered Nb thin films using electrical transport, AFM, and XRD measurements, which demonstrate their sensitivity to small changes in growth parameters, chamber history, and lithographic processes. However, very simple and practical solutions can be employed to restore or maintain high-quality films. A standard lift-off technique based on a PMMA resist leads to contamination and dramatic deterioration of film quality; therefore, we propose an aluminum mask-based approach to address this issue. By comparing the T c, the R 300 K R 15 K ratio and the resistance ratio R 300 K R 77 K, we also demonstrate that preliminary screening of Nb films at liquid nitrogen temperatures is already sufficient to identify samples for a more detailed characterization in a cryogenic environment. Reproducibly fabricating high-quality superconducting films and nanostructures is not reserved for high-end sputtering facilities; rather, it depends on careful tuning of growth conditions.

The authors would like to especially thank Professor Koziej and D. Derelli for their help and support with the XRD measurements and R. Venugopal for conducting the AFM measurements. We would also like to thank S. Haugg and T. Finger for their expertise in operating the sputter chamber. We acknowledge support by the Deutsche Forschungsgemeinschaft (DFG) with Grant No. BL-487/14-1 and the BMBF project “NOVALIS” under Grant No. 05K22GUD.

The authors have no conflicts to disclose.

Vincent Strenzke: Conceptualization (equal); Data curation (lead); Formal analysis (lead); Investigation (lead); Writing – original draft (lead); Writing – review & editing (lead). Annika Weber: Data curation (equal); Formal analysis (equal); Investigation (equal); Writing – review & editing (supporting). Peer Heydolph: Data curation (equal); Formal analysis (equal); Investigation (equal); Writing – review & editing (supporting). Isa Moch: Data curation (equal); Formal analysis (equal); Investigation (equal); Writing – review & editing (supporting). Isabel González Díaz-Palacio: Investigation (supporting); Writing – review & editing (equal). Wolfgang Hillert: Project administration (supporting); Supervision (supporting); Writing – review & editing (supporting). Robert Zierold: Conceptualization (supporting); Funding acquisition (equal); Supervision (supporting); Writing – review & editing (equal). Lars Tiemann: Conceptualization (lead); Funding acquisition (equal); Supervision (lead); Writing – original draft (equal); Writing – review & editing (equal). Robert H. Blick: Conceptualization (supporting); Funding acquisition (equal); Project administration (lead); Supervision (equal); Writing – review & editing (supporting).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
G. H.
Lee
,
S.
Kim
,
S. H.
Jhi
, and
H. J.
Lee
, “
Ultimately short ballistic vertical graphene Josephson junctions
,”
Nat. Commun.
6
,
6181
(
2015
).
2.
V. E.
Calado
,
S.
Goswami
,
G.
Nanda
,
M.
Diez
,
A. R.
Akhmerov
,
K.
Watanabe
,
T.
Taniguchi
,
T. M.
Klapwijk
, and
L. M. K.
Vandersypen
, “
Ballistic Josephson junctions in edge-contacted graphene
,”
Nat. Nanotechnol.
10
,
761
764
(
2015
).
3.
I. V.
Borzenets
,
F.
Amet
,
C. T.
Ke
,
A. W.
Draelos
,
M. T.
Wei
,
A.
Seredinski
,
K.
Watanabe
,
T.
Taniguchi
,
Y.
Bomze
,
M.
Yamamoto
,
S.
Tarucha
, and
G.
Finkelstein
, “
Ballistic graphene Josephson junctions from the short to the long junction regimes
,”
Phys. Rev. Lett.
117
,
237002
(
2016
).
4.
G. H.
Lee
,
D. K.
Efetov
,
W.
Jung
,
L.
Ranzani
,
E. D.
Walsh
,
T. A.
Ohki
,
T.
Taniguchi
,
K.
Watanabe
,
P.
Kim
,
D.
Englund
, and
K. C.
Fong
, “
Graphene-based Josephson junction microwave bolometer
,”
Nature
586
,
42
46
(
2020
).
5.
R.
Kokkoniemi
,
J.-P.
Girard
,
D.
Hazra
,
A.
Laitinen
,
J.
Govenius
,
R. E.
Lake
,
I.
Sallinen
,
V.
Vesterinen
,
M.
Partanen
,
J. Y.
Tan
,
K. W.
Chan
,
K. Y.
Tan
,
P.
Hakonen
, and
M.
Möttönen
, “
Bolometer operating at the threshold for circuit quantum electrodynamics
,”
Nature
586
,
47
51
(
2020
).
6.
B.
Cabrera
,
R. M.
Clarke
,
P.
Colling
,
A. J.
Miller
,
S.
Nam
, and
R. W.
Romani
, “
Detection of single infrared, optical, and ultraviolet photons using superconducting transition edge sensors
,”
Appl. Phys. Lett.
73
,
735
737
(
1998
).
7.
K. D.
Irwin
and
G. C.
Hilton
,
Transition-Edge Sensors
(
Springer
,
Berlin
,
2005
), pp.
63
150
.
8.
J.
Claassen
,
S.
Wolf
,
S.
Qadri
, and
L.
Jones
, “
Epitaxial growth of niobium thin films
,”
J. Cryst. Growth
81
,
557
561
(
1987
).
9.
J.
Wright
,
C.
Chang
,
D.
Waters
,
F.
Lüpke
,
R.
Feenstra
,
L.
Raymond
,
R.
Koscica
,
G.
Khalsa
,
D.
Muller
,
H. G.
Xing
, and
D.
Jena
, “
Unexplored MBE growth mode reveals new properties of superconducting NbN
,”
Phys. Rev. Mater.
5
,
024802
(
2021
).
10.
M.
Miyake
,
Y.
Hirooka
,
R.
Imoto
, and
T.
Sano
, “
Chemical vapor deposition of niobium on graphite
,”
Thin Solid Films
63
,
303
308
(
1979
).
11.
D.
Hazra
,
N.
Tsavdaris
,
A.
Mukhtarova
,
M.
Jacquemin
,
F.
Blanchet
,
R.
Albert
,
S.
Jebari
,
A.
Grimm
,
A.
Konar
,
E.
Blanquet
,
F.
Mercier
,
C.
Chapelier
, and
M.
Hofheinz
, “
Superconducting properties of NbTiN thin films deposited by high-temperature chemical vapor deposition
,”
Phys. Rev. B
97
,
144518
(
2018
).
12.
S.
Linzen
,
M.
Ziegler
,
O. V.
Astafiev
,
M.
Schmelz
,
U.
Hübner
,
M.
Diegel
,
E.
Il’ichev
, and
H.-G.
Meyer
, “
Structural and electrical properties of ultrathin niobium nitride films grown by atomic layer deposition
,”
Supercond. Sci. Technol.
30
,
035010
(
2017
).
13.
I.
González Díaz-Palacio
,
M.
Wenskat
,
G. K.
Deyu
,
W.
Hillert
,
R. H.
Blick
, and
R.
Zierold
, “
Thermal annealing of superconducting niobium titanium nitride thin films deposited by plasma-enhanced atomic layer deposition
,”
J. Appl. Phys.
134
,
035301
(
2023
).
14.
R.
Frerichs
and
C. J.
Kircher
, “
Properties of superconducting niobium films made by asymmetric AC sputtering
,”
J. Appl. Phys.
34
,
3541
3543
(
1963
).
15.
H. C.
Theuerer
and
J. J.
Hauser
, “
Getter sputtering for the preparation of thin films of superconducting elements and compounds
,”
J. Appl. Phys.
35
,
554
555
(
1964
).
16.
D.
Gerstenberg
and
P. M.
Hall
, “
Superconducting thin films of niobium, tantalum, tantalum nitride, tantalum carbide, and niobium nitride
,”
J. Electrochem. Soc.
111
,
936
(
1964
).
17.
J.
Gavaler
,
A.
Santhanam
,
A.
Braginski
,
M.
Ashkin
, and
M.
Janocko
, “
Dimensional effects on current and field properties in NbN films
,”
IEEE Trans. Magn.
17
,
573
576
(
1981
).
18.
L.
Hallett
,
I.
Charaev
,
A.
Agarwal
,
A.
Dane
,
M.
Colangelo
,
D.
Zhu
, and
K. K.
Berggren
, “
Superconducting MoN thin films prepared by DC reactive magnetron sputtering for nanowire single-photon detectors
,”
Supercond. Sci. Technol.
34
,
035012
(
2021
).
19.
J. T.
Gudmundsson
, “
Physics and technology of magnetron sputtering discharges
,”
Plasma Sources Sci. Technol.
29
,
113001
(
2020
).
20.
R.
Behrisch
and
W.
Eckstein
,
Sputtering by Particle Bombardment
(
Springer
,
Berlin
,
2007
).
21.
M.
Konuma
,
Film Deposition by Plasma Techniques
(
Springer
,
Berlin
,
1992
).
22.
J. A.
Thornton
and
D. W.
Hoffman
, “
Stress-related effects in thin films
,”
Thin Solid Films
171
,
5
31
(
1989
).
23.
A.
Anders
, “
A structure zone diagram including plasma-based deposition and ion etching
,”
Thin Solid Films
518
,
4087
4090
(
2010
).
24.
O. V.
Dobrovolskiy
and
M.
Huth
, “
Crossover from dirty to clean superconducting limit in DC magnetron-sputtered thin Nb films
,”
Thin Solid Films
520
,
5985
5990
(
2012
).
25.
D. K.
Finnemore
,
T. F.
Stromberg
, and
C. A.
Swenson
, “
Superconducting properties of high-purity niobium
,”
Phys. Rev.
149
,
231
243
(
1966
).
26.
Y.
Asada
and
H.
Nosé
, “
Superconductivity of niobium films
,”
J. Phys. Soc. Jpn.
26
,
347
354
(
1969
).
27.
G. W.
Webb
, “
Low-temperature electrical resistivity of pure niobium
,”
Phys. Rev.
181
,
1127
1135
(
1969
).
28.
J.
Sosniak
,
J.
Hull
, and
G.
W.
, “
Superconductivity of niobium thin films deposited by DC diode sputtering
,”
J. Appl. Phys.
38
,
4390
4392
(
1967
).
29.
A.
Anferov
,
K.-H.
Lee
,
F.
Zhao
,
J.
Simon
, and
D. I.
Schuster
, “
Improved coherence in optically defined niobium trilayer-junction qubits
,”
Phys. Rev. Appl.
21
,
024047
(
2024
).
30.
G.
von Minnigerode
, “
Der Einfluß von Gitterfehlern auf die Übergangstemperaturen verschiedener Supraleiter
,”
Z. Phys.
154
,
442
459
(
1959
).
31.
W.
DeSorbo
, “
Effect of dissolved gases on some superconducting properties of niobium
,”
Phys. Rev.
132
,
107
121
(
1963
).
32.
B.
Okolo
,
P.
Lamparter
,
U.
Welzel
, and
E. J.
Mittemeijer
, “
Stress, texture, and microstructure in niobium thin films sputter deposited onto amorphous substrates
,”
J. Appl. Phys.
95
,
466
476
(
2004
).
33.
C.
Clavero
,
D. B.
Beringer
,
W. M.
Roach
,
J. R.
Skuza
,
K. C.
Wong
,
A. D.
Batchelor
,
C. E.
Reece
, and
R. A.
Lukaszew
, “
Strain effects on the crystal growth and superconducting properties of epitaxial niobium ultrathin films
,”
Cryst. Growth Des.
12
,
2588
2593
(
2012
).
34.
C.
Todt
,
S.
Telkamp
,
F.
Krizek
,
C.
Reichl
,
M.
Gabureac
,
R.
Schott
,
E.
Cheah
,
P.
Zeng
,
T.
Weber
,
A.
Müller
,
C.
Vockenhuber
,
M. B.
Panah
, and
W.
Wegscheider
, “
Development of Nb-GaAs based superconductor-semiconductor hybrid platform by combining in situ dc magnetron sputtering and molecular beam epitaxy
,”
Phys. Rev. Mater.
7
,
076201
(
2023
).
35.
C.
Wu
, “
Intrinsic stress of magnetron-sputtered niobium films
,”
Thin Solid Films
64
,
103
110
(
1979
).
36.
R.
Jha
and
V. P. S.
Awana
, “
Control of sputtering parameters for deposition of NbN thick films
,”
Nov. Supercond. Mater.
1
,
7
12
(
2013
).
37.
M.
Manzo-Perez
,
M.
Jamalzadeh
,
Z.
Huang
,
X.
Tong
,
K.
Kisslinger
,
D.
Nykypanchuk
, and
D.
Shahrjerdi
, “
Effects of residual oxygen on superconducting niobium films
,”
Appl. Phys. Lett.
125
,
112602
(
2024
).
38.
K.-H.
Müller
, “
Ion-beam-induced epitaxial vapor-phase growth: A molecular-dynamics study
,”
Phys. Rev. B
35
,
7906
7913
(
1987
).
39.
O.
Auciello
and
R.
Kelly
,
Ion Bombardment Modification of Surfaces
(
Elsevier Science Publishers
,
1984
).
40.
D.
Brice
,
J.
Tsao
, and
S.
Picraux
, “
Partitioning of ion-induced surface and bulk displacements
,”
Nucl. Instrum. Methods Phys. Res. B
44
,
68
78
(
1989
).
41.
L.
Wang
,
I.
Meric
,
P. Y.
Huang
,
Q.
Gao
,
Y.
Gao
,
H.
Tran
,
T.
Taniguchi
,
K.
Watanabe
,
L. M.
Campos
,
D. A.
Muller
,
J.
Guo
,
J.
Hone
,
K. L.
Shepard
, and
C. R.
Dean
, “
One-dimensional electrical contact to a two-dimensional material
,”
Science
342
,
614
617
(
2013
).
42.
Y.
Seo
,
S.
Masubuchi
,
E.
Watanabe
,
M.
Onodera
,
R.
Moriya
,
K.
Watanabe
,
T.
Taniguchi
, and
T.
Machida
, “
Selective etching of hexagonal boron nitride by high-pressure CF 4 plasma for individual one-dimensional ohmic contacts to graphene layers
,”
Appl. Phys. Lett.
117
,
243101
(
2020
).
43.
Y.
Cao
,
V.
Fatemi
,
S.
Fang
,
K.
Watanabe
,
T.
Taniguchi
,
E.
Kaxiras
, and
P.
Jarillo-Herrero
, “
Unconventional superconductivity in magic-angle graphene superlattices
,”
Nature
556
,
43
50
(
2018
).
44.
M. S.
Choi
,
N.
Ali
,
T. D.
Ngo
,
H.
Choi
,
B.
Oh
,
H.
Yang
, and
W. J.
Yoo
, “
Recent progress in 1D contacts for 2D-material-based devices
,”
Adv. Mater.
34
,
2202408
(
2022
).
45.
U.
Welp
,
Z. L.
Xiao
,
J. S.
Jiang
,
V. K.
Vlasko-Vlasov
,
S. D.
Bader
,
G. W.
Crabtree
,
J.
Liang
,
H.
Chik
, and
J. M.
Xu
, “
Superconducting transition and vortex pinning in Nb films patterned with nanoscale hole arrays
,”
Phys. Rev. B
66
,
212507
(
2002
).
46.
U.
Welp
,
Z. L.
Xiao
,
V.
Novosad
, and
V. K.
Vlasko-Vlasov
, “
Commensurability and strong vortex pinning in nanopatterned Nb films
,”
Phys. Rev. B
71
,
014505
(
2005
).
47.
J.
Emsley
,
The Elements
(
Oxford University Press
,
1998
).
48.
H.
Gao
,
S.
Wang
,
D.
Xu
,
X.
Wang
,
Q.
Zhong
,
Y.
Zhong
,
J.
Li
, and
W.
Cao
, “
Study of DC magnetron sputtered Nb films
,”
Crystals
12
,
31
(
2022
).