You are on page 1of 18

Version of Record: https://www.sciencedirect.

com/science/article/pii/S0040609019301105
Manuscript_62f906f6114009a7fda2d5e37ca6e659

Characterization of superconducting Sn thin films and their application to

ferromagnet-superconductor hybrids

Wonbae Bang a,1, Tyler D. Morrison a, K. D. D. Rathnayaka a, Igor F. Lyuksyutov a, Donald G.

Naugle a,b, Winfried Teizer a,b,c,*


a
Department of Physics and Astronomy, Texas A&M University, College Station, TX 77843, USA
b
Department of Materials Sciences and Engineering, Texas A&M University, College Station, TX

77843, USA
c
WPI-Advanced Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan

Abstract

We characterize thermally quench condensed (~80 K) Sn thin films of two different thicknesses to

explore their usefulness for vortex studies. The coherence length and the London penetration depth

near the critical temperature indicate that the thin films are in the dirty limit and behave as type II

superconductors. Furthermore, when we add ferromagnetic nanostripes on top of the Sn thin films,

the ferromagnet-superconductor hybrids show hysteric and anisotropic behavior. A ferromagnet-

superconductor hybrid with a thinner Sn film exhibits a stronger effect.

Keywords: Sn thin films, Type II superconductor, Ferromagnet-Superconductor Hybrids.

Highlights

• Thermally quench condensed superconducting Sn thin films in ferromagnet-superconductor

hybrids.

* Corresponding Author at: Department of Physics and Astronomy, 4242 TAMU, College Station, TX 77843-4242,
USA. E-mail: teizer@physics.tamu.edu
1
Present address: Department of Physics and Astronomy, Northwestern University, Evanston, IL 60208

© 2019 published by Elsevier. This manuscript is made available under the Elsevier user license
https://www.elsevier.com/open-access/userlicense/1.0/
• The Sn thin films behave as Type II superconductors in the dirty limit.

• The ferromagnet-superconductor hybrids with thinner Sn films exhibit stronger interactions.

1. Introduction

A fundamental property of an equilibrium state in superconductors is the magnetic structure

which occurs when external magnetic fields are applied. The equilibrium states of type I and type II

superconductors are different, i.e. the intermediate and vortex states, respectively. This equilibrium

state of the superconductor depends on the material and its geometry. A thin film of a bulk type I

superconductor can become a type II superconductor if the thickness of the thin film is sufficiently

reduced below the critical thickness [1-4].

Tin (Sn) is a well-known material with numerous applications in industry and science. The

properties of Sn have been explored thoroughly in the literature [5-9]. However, applications of thin

Sn films to Ferromagnet-Superconductor Hybrids (FSH) have received much less attention.

Typically, the FSH consists of magnetic dots over a superconducting thin film to study the interplay

between the superconducting thin film and magnetic structures. Though superconductivity and

ferromagnetism seem to be mutually exclusive states, they show fascinating properties when they

coexist in a nanoscale system, such as hysteresis and enhanced superconductivity [10]. In bulk, Sn

is a Type I superconductor [11], i.e. κ < 1 / √2, where κ = λ / ξ is the Ginzburg-Landau parameter

[12], ξ is the coherence length, and λ is the London penetration depth. Reduction of dimensionality

of the films results in a switch of the parameter to κ > 1 / √2 and subsequent behavior of a Type II

superconductor [2]. In a Type II superconductor, the vortex or mixed state can be explored [12].

Strong anisotropy in critical currents has been observed in FSH with 3.5 µm width of Co [13], and

25 µm width of Fe [14]. Furthermore, submicron-sized magnetic stripes in FSHs showed some

interesting properties such as hysteretic behavior and enhanced flux pinning [15, 16]. In this paper,

2
we discuss our experimental results, using Type II superconducting Sn films. In addition, we

investigate the applicability of Sn thin films for use in vortex dynamic studies and as the

superconductor in FSH systems. Specifically, we characterize the surface morphology, electrical

transport, and magnetic properties of the films near their TC. Additionally, we discuss changes of

superconducting properties with varying thin film thickness in FSHs.

2. Experimental

Sequential thermal evaporations were carried out in order to produce samples of Sn thin

films with two different thicknesses. Firstly, Au contact pads for 4-probe measurement were

fabricated on a Si wafer with a 300 nm SiO2 layer, using a combination of photolithography and

thermal evaporation. A 3 nm Ge interlayer, which provides a uniform morphology, was deposited at

room temperature via thermal evaporation as an adhesion layer between the SiO2 layer and the Sn

thin film [17]. A liquid nitrogen (LN2) cooled sample mount in the evaporation chamber was used

for quench condensation of Sn. This quench condensation technique provides that conductive thin

films at a lower thickness than that of the same materials deposited at higher temperatures [18].

Beutel et al. reported that the percolation occurs below 93 nm of thermally evaporated Sn films

without any adhesive layer at room temperature. Unlike the previous study, our 50 nm of Sn thin

films is conductive due to cryogenic temperatures and a wetting layer [19]. The thickness of the thin

film was monitored by a Quartz Crystal Microbalance (QCM) in real time. The final thickness was

varied by changing the deposition time. The deposition rate was ~0.11 nm/sec, and the base

pressure in the evaporator chamber was ~1.16 × 10-5 Pa [20]. Two different thicknesses (t) of Sn

thin films were fabricated: 50 nm and 100 nm, which are less than the critical thickness of Sn (180

nm) [2]. The films had dimensions of 84.33 µm length (L) and 15.50 µm width (W) on the

substrates. In order to create FSH samples, a 20 nm insulating film of Ge was thermally evaporated

onto the Sn films. Nanostripes (100 nm width × 120 nm height) on top of the Ge film were

3
thermally evaporated and patterned by electron-beam lithography, either parallel or perpendicular to

the applied current. The period (center to center) of the Ni nanostripes was 500 nm. As a result, the

FSH samples consist of thermally evaporated Ni nanostripes and 50 nm or 100 nm of Sn thin films.

The films in the FSHs and the control samples were simultaneously fabricated on the same substrate.

For surface observation, a Scanning Electron Microscope (JSM-7500F, JEOL) and an Atomic Force

Microscope (Bruker Dimension Icon AFM) were utilized to examine the films. Also, a Quantum

Design Physical Property Measurement System (PPMS, Model 6000) was used for characterization

of the films.

3. Results and discussion

3-1. Surface and critical temperature of the Sn thin films

Scanning electron microscope (SEM) images (80,000x) and two-dimensional atomic force

microscope (AFM) images (scan size 1.0 µm × 1.0 µm) for the 50 nm and 100 nm Sn thin films are

shown in Fig. 1. The SEM images indicate that the thin films grow in layer-plus-island mode

(Stranski-Krastanov mode) [21]. The films were grown layer-by-layer up to a stable wetting layer.

Subsequent observation revealed three dimensional islands growing on the Sn rich layer [21-23].

Two-dimensional AFM images indicate that the surface roughness increases with the film thickness.

Morphological grain sizes in the AFM images [24] measured using Nanoscope Analysis by

Bruker, increase monotonically with the thin films' thickness. The surface roughness and the

morphological grain size are given in Table 1. The diameter of the grain size increases with

increasing thickness of the thin film, and results in the decrease of the surface density of islands

[22]. Thus, surface roughness and grain size of the thin film depend on the thickness of the thin film

[25-27].

The physical properties of the thin films were measured in the PPMS. The thin films were

electrically connected to the PPMS through Au wires with cold-pressed In joints to Au contact pads.

4
An external magnetic field was applied perpendicularly to the thin films. In order to find the

resistance (R) of the thin films, a 1.0 µA alternating current was applied to each sample. The

voltage at zero field was then measured. The temperature was decreased during the measurement of

R, leading to resistivity (ρ) versus temperature curves (not shown here).

ρ decreases with increasing film thickness which is given in Table 1. In this case, scattering

on the surface of the Sn thin films is considered to play a key role in the increase of ρ [28, 29]. In

accordance with the Fuchs and Sondheimer theory [30, 31], the resistivity dependence on the

thickness is due to electron scattering at the thin film surface and interface [29, 32]. From their

theory, the simplified resistivity ρs of the thin film can be written as

= 1+ 1− , (1)

where ρ0 is the bulk metal resistivity and p is the specularity coefficient. k is defined as k = t / l,

where t is the film thickness, and l is the electron mean free path. The specularity coefficient p is

determined to be p ≈ 0 for polycrystalline films and p ≈ 1 for single crystal films [33]. Therefore, it

is clear that the resistivity of the thin films is inversely proportional to their thickness.

Each TC of the samples was defined by linear extrapolation as shown in Fig. 2. TC for the 50

nm and 100 nm films correspond to 3.71 K and 3.70 K, respectively. Dependence of TC on the thin

film thickness has been reported in the literature [5, 34, 35]. Lock explained that the increase in TC is

related to stresses in the films [7]. Strain effects on TC have been quantitatively confirmed by

experiments involving deposition of thin films on ferroelectric substrates [36]. When TC is higher,

the coefficient of expansion for the substrate material is lower than that of Sn. On the other hand,

when TC in the film is lower, the coefficient of expansion of the substrate material is higher than that

of Sn. This leads to the conclusion that the changes in TC of the thin films were due to strain. This

strain develops from the differential contraction of the thin film versus the substrate upon cooling

from room temperature to the transition temperature [7]. The variation of TC due to the effects of

stress can be written as

5
!"
∆ = 1.673 + ∙ #
$, (2)

where σ0 is an intrinsic critical shear stress, α is a constant of approximately 0.5, G is the shear

modulus, t is the thickness of the film, and b is the appropriate Burgers vector (a lattice constant in

the a-a plane in the tetragonal structure of Sn) [5, 37, 38]. ∆TC is defined as the change between the

film’s TC and the bulk TC (3.722K). Therefore, it seems that the thickness of the thin film affects TC

as 1 / t.

3-2. Characteristic lengths of the Sn thin films

The extracted HC2 of the films is plotted as a function of the reduced temperature (T / Tc),

as shown in Fig. 3(a). Dots are the experimental data, and the lines are fitting curves. The estimated

HC2(0) of the 50 nm and 100 nm films are 0.0796 T, and 0.0470 T, respectively. Fig. 3(b) indicates
3 2
HC as a function of T/Tc for bulk Sn, which is expressed as HC(T) = 1.225T – 26.8T + 307.3,

where T is temperature. HC(0) is 0.03073 T [39].

The coherence length ξ, which is one of the characteristic lengths in superconductors, was

calculated from the experimental results of the R-H curves of the thin films. By combining two

equations in Eq. (3) [12],

.
)* + / 3
% = 0.855 $ and 0 1 ≅ 14)* / , (3)
,-#

HC2 becomes

3* -
0 1 = 14 ∙ , (4)
. 55 / )* +

and the derivative with respect to T yields

67 / -3* ,
6
= 14 . 55 / )* +
∙ , (5)

6
where dHC2 / dT was determined from Fig. 3(a) in a range between 0.97TC and TC. Finally, the BCS

ξ0 of the thin films could be calculated. The estimated ξ0 is 12.78 nm for the 50 nm film and 21.64

nm for the 100 nm film, as shown in Table 2. Clearly, the ξ0 of the films is less than their thickness.

<
In order to estimate mean free paths (l) of the films, 8 = 9: ; [40] and 8 = 8′, where ρ′

and l′ are resistivity and mean free path at 3.80 K, [41, 42] are used, where νF is the Fermi velocity,

τ is the relaxation time. For bulk Sn, ρ = 11.5 µΩ ·cm at 273 K and l = 3.976 nm. Since ρ, ρ', and l

are estimated, l' can be computed for the Sn films. The resistivity of the thin films was determined

at 3.80 K, when the thin films were in the normal state and slightly above their TC. The estimated l'

at 3.80 K can be seen in Table 2.

The London penetration depth at zero temperature is expressed in the BCS theory by [12]

3*
>? 0 = . ,@54/ )* 7
. (6)

All the parameters needed to calculate λL(0) are available from our experimental results. The

calculations are summarized in Table 2. The ratio of l'/ξ0 is smaller than 1 for the films. This result

demonstrates that the use of the dirty limit is appropriate.

κ is obtained as

BCDD +, BF
A= = 0.715 , (7)
)* +

where λeff (l, T) = λL(T)(ξ0 / J(0, T)l)1/2, l denotes the l' of the films. The J(0, T) is defined as 1.33

near TC [12].

As is evident, κ is much greater than 1 / √2 and therefore all the Sn thin films in this work

indeed behave as type II superconductors. In general, κ decreases with increasing film thickness. It

can be expected that thicker films or bulk samples would have κ lower than 1, leading to type I

behavior. Therefore, it is reasonable to conclude that thin films are behaving as type II

superconductors as opposed to the type I behavior of bulk Sn.

7
3-3. Ferromagnet-Superconductor Hybrids below TC

The 50 nm and 100 nm Sn thin films were then used to create four different FSHs: a

parallel FSH of a 50 nm Sn film, a perpendicular FSH of a 50 nm Sn film, a parallel FSH of a 100

nm Sn film, and a perpendicular FSH of a 100 nm Sn film. Fig. 4(a) shows a schematic of the FSHs

setup. Both FSHs were measured at the same time. SEM images at high magnification in Fig. 4(c)

and (d) show parallel and perpendicular FSHs, respectively. We obtained magnetoresistance of the

FSHs after magnetizing Ni nanostripes along a vertical direction below TC.

Figures 5(a) and (b) show R-H curves for the FSHs at 3.60 K. R is normalized by dividing

through Rmax, the resistance in the normal state slightly above TC. The green arrows in Fig. 5

indicate the direction of the magnetic field sweep. All the FSHs show hysteretic behavior, however

the parallel FSH of the 50 nm film indicates a stronger hysteretic behavior in the R-H curves than

that of the 100 nm film, as shown in Fig. 5(a). The parallel FSH of the 50 nm film remains

superconducting in |H| < 0.01 T at 3.60 K; on the other hand, the parallel FSH of the 100 nm film

remains superconducting in |H| < 0.003 T at the same temperature. A similar phenomenon occurs in

the perpendicular FSHs of the 50 nm and 100 nm films, as shown in Fig. 5(b). Phase diagrams of

HC2 as a function of temperature near TC are shown in Fig. 5(c) for parallel and (d) for

perpendicular FSHs. The resulting phase diagrams demonstrate strong hysteresis due to the

hysteresis of the nanostripes, and the phase diagrams are similar to those predicted by mean field

theory [43]. The magnetic field distribution is created by the magnetic nanostripe array. The array

on the superconducting thin film is exposed to this inhomogeneous magnetic field. The magnetic

field distribution depends on the properties of the nanostripes and the thickness of the

superconducting thin films [44, 45]. The ratio of the thin film’s thickness (t) and the width of the

nanostripes (d) is significant. If t / d is smaller than 1, the direction of the magnetic field inside the

superconducting thin film is normal to the film surface. Also, the magnetic field inside the

superconducting thin film is close to the field at the surface of the nanostripes. Therefore, the

8
direction of the magnetic field between the nanostripes will be changed. In this case, a field

compensation effect can be expected [45, 46]. The ratio of t / d for 50 nm FSHs is less than the ratio

for 100 nm FSHs. Therefore, both the parallel and perpendicular FSHs of the 50 nm film show

about twice as large a hysteretic behavior than that of the 100 nm film. In addition, the FSHs exhibit

anisotropy in HC2 as a function of temperature. The parallel FSHs show a larger HC2 at the same

temperature. This anisotropic behavior can be explained by vortices in the Sn thin films interacting

magnetically with the magnetized Ni nanostripes. Therefore, the vortices have different barriers for

directions of motion parallel and perpendicular to the nanostripes [15]. The superconducting

properties of the FSHs depend on the thickness of the thin films. The R-H curves at lower

temperatures shows the same hysteresis, however the samples keep their superconducting state in

higher external magnetic fields. Since the R-H curves are not significantly different in their

temperature dependence, we do not discuss the R-H curves at lower temperatures in this paper.

The discussion of HC2 dependence on the current direction assumes that the film

temperature is reasonably below TC . Near TC, new effects become important. The onset of

resistivity for the case of current perpendicular to the stripes occurs at lower temperature than that

for current parallel to the stripes. On the phase diagram, it is indicated as a lower TC . This is direct

indication that the magnetic field from stripes destroys superconductivity under the stripe and

nearby. Thus, the superconducting film near TC is actually a set of parallel superconducting and

insulating stripes. As a result, for the current parallel to the stripes, the system appears to be

superconducting, but for current perpendicular to the stripes it appears to be resistive. This can

result in interesting physical behavior which includes both “bulk” superconductivity and “surface”

superconductivity [47-51], where “surface” means the interface across the film between region of

destroyed superconductivity and the remaining superconducting region. The study of

superconductivity in this temperature region goes well beyond the goal of this paper and is a matter

of future studies.

9
The superconducting properties of the FSHs depend on the thickness of the thin films.

When the dimension of the Ni nanostripes are the same, but different superconductor thicknesses

are used in the in the FSH, the thinner Sn film FSH indicates enhanced superconductivity and

stronger hysteretic behavior below their TCs. Also, both the parallel and perpendicular FSHs

showed anisotropy when the Ni nanostripes are magnetized.

4. Conclusion

We have studied the correlation between thickness and the superconducting properties of

Sn thin films. We show that the type of superconductivity exhibited depends on the thickness of the

thin film. Specifically, we demonstrate that our thin films behave as Type II superconductors.

Difference in the thickness of the Sn thin film in the FSHs shows hysteric and anisotropic behavior

by parallel and perpendicular Ni nanostripes, after the nanostripes are magnetized. Thinner films in

the FSHs show stronger interaction when the Ni nanostripe structures are the same. Since our Sn

thin films function as type II superconductors, they present an excellent candidate to study vortex

dynamics by adding ferromagnetic nanostructures.

Acknowledgements

This work was supported by the Robert A. Welch Foundation, Houston, Texas (Grant A-

0514), the U.S. Department of Energy (Grant DE-FG02-07ER46450) and the WPI program of

MEXT, Japan.

References

10
[1] H. Boersch, U. Kunze, B. Lischke, W. Rodewald, Observation of mixed state in films of Type I

superconductors (Pb), Phys. Lett. A, 44A(4) (1973) 273-274. https://doi.org/10.1016/0375-

9601(73)90917-1

[2] G. J. Dolan, J. Silcox, Critical Thicknesses in superconducting thin films, Phys. Rev. Lett. 30(13) (1973)

603-606. https://doi.org/10.1103/PhysRevLett.30.603

[3] W. Rodewald, Triangular vortex lattice in thin films of Type I superconductors, Phys. Lett. A, 55A(2)

(1975) 135-136. https://doi.org/10.1016/0375-9601(75)90156-5

[4] M. Tinkham, Effect of fluxoid quantization on transitions of superconducting films, Phys. Rev. 129(6)

(1963) 2413-2422. https://doi.org/10.1103/PhysRev.129.2413

[5] R. H. Blumberg, D. P. Seraphim, Effect of elastic strain on superconducting critical temperature of

evaporated tin films, J. Appl. Phys. 33(1) (1962) 163-168. https://doi.org/10.1063/1.1728478

[6] D. H. Douglass, R. H. Blumberg, Precise critical field measurements of superconducting Sn films in

London limit, Phys. Rev. 127(6) (1962) 2038-2043. https://doi.org/10.1103/PhysRev.127.2038.

[7] J. M. Lock, Penetration of magnetic fields into superconductors III. Measurements on thin films of Tin,

Lead and Indium, Proc. R. Soc. Lond. A Math. Phys. Sci. 208(1094) (1951) 391-408.

http://doi.org/10.1098/rspa.1951.0169

[8] B. G. Orr, H. M. Jaeger, A. M. Goldman, Local superconductivity in ultrathin Sn films, Phys. Rev. B

32(11) (1985) 7586-7589. https://doi.org/10.1103/PhysRevB.32.7586.

[9] M. Strongin, R. S. Thompson, O. F. Kammerer, J. E. Crow, Destruction of superconductivity in

disordered near-monolayer films, Phys. Rev. B 1(3) (1970) 1078-1091.

https://doi.org/10.1103/PhysRevB.1.1078

[10] I. F. Lyuksyutov, V.L. Pokrovsky, Ferromagnet-superconductor hybrids. Adv. Phys. 54(1) (2005) 67-

136. https://doi.org/10.1080/00018730500057536

[11] C. Kittel, Introduction to solid state physics, 8th ed., Wiley, Hoboken, New Jersey, 2005.

[12] M. Tinkham, Introduction to superconductivity, 2nd ed., Dover Publications, Mineola, New York, 2004.

[13] A. E. Ozmetin, M. K. Yapici, J. Zou, I. F. Lyuksyutov, D. G. Naugle, Micromagnet-superconducting

hybrid structures with directional current flow dependence for persistent current switching, Appl. Phys.

Lett. 95(2) (2009) 022506. https://doi.org/10.1063/1.3176481

11
[14] A. E. Ozmetin, K. D. D. Rathnayaka, D. G. Naugle, I. F. Lyuksyutov, Strong increase in critical field

and current in magnet-superconductor hybrids, J. Appl. Phys. 105(7) (2009) 07E324.

https://doi.org/10.1063/1.3076517

[15] K. Kim, K. K. D. Rathnayaka, I. F. Lyuksyutov, D. G. Naugle, Superconducting film with an array of

magnetic nanostripes, Int. J. Mod. Phys. B, 27(15) (2013) 1362020.

https://doi.org/10.1142/S0217979213620208

[16] Z. Ye, I. F. Lyuksyutov, W. Wu, D. G. Naugle, Strongly anisotropic flux pinning in superconducting

Pb82Bi18 thin films covered by periodic ferromagnet stripes, Supercond. Sci. Technol. 24 (2011)

024011. http://doi.org/10.1088/0953-2048/24/2/024011

[17] B. Kain, R. P. Barber, Resistive transitions in quench-condensed Bi films near a normal-metal ground

plane. Phys. Rev. B 68(13) (2003) 134502. https://doi.org/10.1103/PhysRevB.68.134502

[18] R. C. Dynes, J. P. Garno, J. M. Rowell, Two-Dimensional Electrical-Conductivity in Quench-

Condensed Metal-Films. Phys. Rev. Lett. 40(7) (1978) 479-482.

https://doi.org/10.1103/PhysRevLett.40.479

[19] M. H. Beutel, N. G. Ebensperger, M. Thiemann, G. Untereiner, V. Fritz, M. Javaheri, J. Nagele, R.

Rosslhuber, M. Dressel, M. Scheffler, Microwave study of superconducting Sn films above and below

percolation. Supercond. Sci. Technol. 29(8) (2016) 085011. https://doi.org/10.1088/0953-

2048/29/8/085011

[20] W. Bang, W. Teizer, D. G. Naugle, I. F.Lyuksyutov, Electroplated high-aspect-ratio ferromagnetic

nanopillars and their application to Ferromagnet-Superconductor Hybrids, Microelectron. Eng. 181

(2017) 55-59. https://doi.org/10.1016/j.mee.2017.08.001

[21] A. Baskaran, P. Smereka, Mechanisms of Stranski-Krastanov growth, J. Appl. Phys. 111 (2012) 044321.

https://doi.org/10.1063/1.3679068

[22] N. Bündgens, H. Lüth, M. Mattern-Klosson, A. Spitzer, A.Tulke, Sn overlayers on GaAs(110): Growth

mechanism and band bending, Surf. Sci. 160(1) (1985) 46-56. https://doi.org/10.1016/0039-

6028(85)91025-8

[23] H. Lüth, Solid surfaces, interfaces and thin films, 4th, rev. and extended ed., Springer, New York, 2001.

12
[24] L. L. Melo, A. R. Vaz, M. C. Salvadori, M. Cattani, Grain sizes and surface roughness in platinum and

gold thin films, J. Metastable Nanocrystalline Mater. 20-21 (2004) 623-628.

http://doi.org/10.4028/www.scientific.net/JMNM.20-21.623

[25] N. Jadhav, E. J. Buchovecky, L. Reinbold, S. Kumar, A. F. Bower, E. Chason, Understanding the

correlation between intermetallic growth, stress evolution, and Sn whisker nucleation, IEEE Trans.

Electron. Packag. Manuf. 33(3) (2010) 183-192. http://doi.org/10.1109/TEPM.2010.2043847

[26] H. Kim, J. S. Horwitz, G. Kushto, A. Piqué, Z. H. Kafafi, C. M. Gilmore, D. B. Chrisey, Effect of film

thickness on the properties of indium tin oxide thin films, J. Appl. Phys. 88(10) (2000) 6021-6025.

https://doi.org/10.1063/1.1318368

[27] A. Kumar, D. Singh, D. Kaura, Grain size effect on structural, electrical and mechanical properties of

NiTi thin films deposited by magnetron co-sputtering, Surf. Coat. Technol. 203(12) (2009) 1596-1603.

https://doi.org/10.1016/j.surfcoat.2008.12.005

[28] Y. Ke, F. Zahid, V. Timoshevskii, K. Xia, D. Gall, H. Guo, Resistivity of thin Cu films with surface

roughness, Phys. Rev. B 79 (2009) 155406. https://doi.org/10.1103/PhysRevB.79.155406

[29] J. J. Plombon, E. Andideh, V. M. Dubin, J. Maiz, Influence of phonon, geometry, impurity, and grain

size on Copper line resistivity, Appl. Phys. Lett. 89 (2006) 113124. https://doi.org/10.1063/1.2355435

[30] K. Fuchs, The conductivity of thin metallic films according to the electron theory of metals, Proc.

Cambridge Phil. Soc. 34 (1938) 100-108. https://doi.org/10.1017/S0305004100019952

[31] E. H. Sondheimer, The Mean Free Path of Electrons in Metals, Adv. Phys. 1(1) (1952) 1-42.

https://doi.org/10.1080/00018735200101151

[32] J. W. Lim, K. Mimura, M. Isshiki, Thickness dependence of resistivity for Cu films deposited by ion

beam deposition, Appl. Surf. Sci. 217(1-4) (2003) 95-99. https://doi.org/10.1016/S0169-

4332(03)00522-1

[33] J. C. Anderson, The use of thin films in physical investigations; a NATO Advanced Study Institute held

at the Imperial College of Science and Technology, University of London, 19-24 July, 1965, Academic

Press, New York, 1966.

[34] A. I. Gubin, K. S. Il’in, S. A. Vitusevich, M. Siegel, N. Klein, Dependence of magnetic penetration

depth on the thickness of superconducting Nb thin films, Phys. Rev. B 72 (2005) 064503.

https://doi.org/10.1103/PhysRevB.72.064503

13
[35] A. M. Toxen, Size effects in thin superconducting Indium films. Phys. Rev. 123(2) (1961) 442-446.

https://doi.org/10.1103/PhysRev.123.442

[36] D. Stamopoulos, M. Zeibekis, S. J. Zhang, Control of superconductivity by means of electric-field-

induced strain in superconductor/piezoelectric hybrids. J. Appl. Phys.123(2) (2018) 023903.

https://doi.org/10.1063/1.5005045

[37] A. Cottrell, Dislocations and plastic flow in crystals, The International series of monographs on physics,

Clarendon Press, Oxford, 1953.

[38] D. R. Lide, CRC Handbook of Chemistry and Physics. 87th ed., Taylor & Francis Group: Florida, 2006.

[39] N. L. Muench, Effects of Stress on Superconducting Sn, in, Tl, and Al, Phys. Rev. 99(6) (1955) 1814-

1820. https://doi.org/10.1103/PhysRev.99.1814.

[40] N. W. Ashcroft, N. D. Mermin, Solid state physics, Brooks Cole, New York, 1976.

[41] E. R. Andrew, The size-variation of resistivity for mercury and Tin, Proc. Phys. Soc. London A 62(2)

(1949) 77-88. https://doi.org/10.1088/0370-1298/62/2/301

[42] A. C. B. Lovell, The electrical conductivity of thin metallic films I-Rubidium on pyrex glass surfaces,

Proc. R. Soc. Lond. A Math. Phys. Sci. 157(A891) (1936) 0311-0330.

http://doi.org/10.1098/rspa.1936.0197

[43] I. F. Lyuksyutov, Magnetic nanorod-superconductor hybrids, J. Supercon. Nov. Magn. 23(6) (2010)

1047-1049. https://doi.org/10.1007/s10948-010-0678-z.

[44] I. F. Lyuksyutov, Nanoscale magnetic traps, Modern Physics Letters B, 16(15-16) (2002) 569-576.

https://doi.org/10.1142/S0217984902004081

[45] Z. Ye, I. F. Lyuksyutov, W. Wu, D. G. Naugle, Superconducting properties of Pb82Bi18 films

controlled by ferromagnetic nanowire arrays, Supercond. Sci. Technol. 24(2) (2011) 024019.

http://doi.org/doi:10.1088/0953-2048/24/2/024019

[46] M. Lange, M. J. Van Bael, Y. Bruynseraede, V. V. Moshchalkov, Nanoengineered magnetic-field-

induced superconductivity, Phys. Rev. Lett. 90(19) (2003) 197006.

https://doi.org/10.1103/PhysRevLett.90.197006

[47] A. Y. Aladyshkin, A. I. Buzdin, A. A. Fraerman, A. S. Mel'nikov, D. A. Ryzhov, A. V. Sokolov,

Domain-wall superconductivity in hybrid superconductor-ferromagnet structures. Phys. Rev. B 68(18)

(2003) 184508. https://doi.org/10.1103/PhysRevB.68.184508

14
[48] A. Y. Aladyshkin, J. Fritzsche, R. Werner, R. B. G. Kramer, S. Guénon, R. Kleiner, D. Koelle, V. V.

Moshchalkov, Crossover between different regimes of inhomogeneous superconductivity in planar

superconductor-ferromagnet hybrids. Phys. Rev. B 84(9) (2011) 094523.

https://doi.org/10.1103/PhysRevB.84.094523

[49] D. Saint-James, P. G. Gennes, Onset of superconductivity in decreasing fields. Phys. Lett. 7(5) (1963)

306-308. https://doi.org/10.1016/0031-9163(63)90047-7

[50] Z. R. Yang, M. Lange, A. Volodin, R. Szymczak, V. V. Moshchalkov, Domain-wall superconductivity

in superconductor-ferromagnet hybrids. Nat. Mater. 3(11) (2004) 793-798.

https://doi.org/10.1038/nmat1222

[51] M. A. Zorro, T. Saraiva, C. C. D. Silva, Nucleation of superconductivity in multiply connected

superconductor-ferromagnet hybrids. Supercond. Sci. Technol. 27(5) (2014) 055002.

https://doi.org/10.1088/0953-2048/27/5/055002

The list of tables

Table 1. Data of surface observation and critical temperature of the 50 nm and 100 nm Sn thin films.

Table 2. Estimated values of the characteristic lengths of the Sn thin films.

Figure Captions

15
Fig. 1. SEM (a, b) and AFM (c, d) images of (a, c) 50-nm and (b, d) 100-nm Sn thin films. The z-

axis ranges of the AFM images are ±30 for the 50-nm and ±80 for the 100-nm film.

Fig. 2. Resistivity as a function of temperature around the critical temperature. Inset: resistivity as a

a function of temperature, with the range extended to room temperature.

16
Fig. 3. (a) HC2 as a function of T/TC of the thin films, and (b) HC as a function of T/TC of bulk Sn.

Fig. 4. (a) a schematic of the FSH hybrid sample (b) SEM images of the overall FSH setup, with a

higher magnification image of (b) the parallel FSH and (c) the perpendicular FSH.

17
Fig. 5. Comparison of R-H curves of (a) parallel FSHs with 50 nm and 100 nm films and (b)

perpendicular FSHs with 50 nm and 100 nm films. HC2-T of (c) parallel FSHs with the 50 nm and

100 nm films and (d) perpendicular FSHs with the 50 nm and 100 nm films.

Table 1.

Thickness (nm) Surface Roughness (nm) Grain size (nm) TC (K) ρ at 300K (µΩ·cm)
50 5.8 ± 0.5 110.9 ± 5.5 3.71 50.32
100 14.8 ± 0.8 240.6 ± 12 3.70 25.04

Table2.

Thickness (nm) λL (0) (nm) ξ0 (nm) l' (nm) κ l'/ξ0


50 279.37 12.78 3.61 55.33 0.28
100 262.73 21.64 14.13 13.29 0.65

18

You might also like