Next Article in Journal
Application of Infrared Spectroscopy in Prediction of Asphalt Aging Time History and Fatigue Life
Next Article in Special Issue
Improved Mechanical Properties and Corrosion Resistance of Mg-Based Bulk Metallic Glass Composite by Coating with Zr-Based Metallic Glass Thin Film
Previous Article in Journal
High-Porosity Thermal Barrier Coatings from High-Power Plasma Spray Equipment—Processing, Performance and Economics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Stoichiometry on Structural, Morphological and Nanomechanical Properties of Bi2Se3 Thin Films Deposited on InP(111) Substrates by Pulsed Laser Deposition

1
Department of Mechanical and Electro-Mechanical Engineering, National Sun Yat-Sen University, Kaohsiung 804, Taiwan
2
Department of Physics and Biophysics, Faculty of Basic Sciences, Can Tho University of Medicine and Pharmacy, 179 Nguyen Van Cu Street, Can Tho City 94000, Vietnam
3
Faculty of Basic Sciences, Nam Can Tho University, 168 Nguyen Van Cu (Ext) Street, Can Tho City 94000, Vietnam
4
Department of Electrical and Electronic Engineering, Faculty of Automotive Engineering, Ngo Quyen University, Thu Dau Mot City 820000, Vietnam
5
Department of Electrophysics, National Chiao Tung University, Hsinchu 300, Taiwan
6
Department of Materials Science and Engineering, National Chiao Tung University, Hsinchu 300, Taiwan
7
Department of Materials Science and Engineering, I-Shou University, Kaohsiung 840, Taiwan
8
Department of Fragrance and Cosmetic Science, Kaohsiung Medical University, 100 Shin-Chuan 1st Road, Kaohsiung 80782, Taiwan
*
Authors to whom correspondence should be addressed.
Coatings 2020, 10(10), 958; https://doi.org/10.3390/coatings10100958
Submission received: 7 August 2020 / Revised: 14 September 2020 / Accepted: 30 September 2020 / Published: 5 October 2020
(This article belongs to the Special Issue Recent Advances in the Growth and Characterizations of Thin Films)

Abstract

:
In the present study, the structural, morphological, compositional, nanomechanical, and surface wetting properties of Bi2Se3 thin films prepared using a stoichiometric Bi2Se3 target and a Se-rich Bi2Se5 target are investigated. The Bi2Se3 films were grown on InP(111) substrates by using pulsed laser deposition. X-ray diffraction results revealed that all the as-grown thin films exhibited were highly c-axis-oriented Bi2Se3 phase with slight shift in diffraction angles, presumably due to slight stoichiometry changes. The energy dispersive X-ray spectroscopy analyses indicated that the Se-rich target gives rise to a nearly stoichiometric Bi2Se3 films, while the stoichiometric target only resulted in Se-deficient and Bi-rich films. Atomic force microscopy images showed that the films’ surfaces mainly consist of triangular pyramids with step-and-terrace structures with average roughness, Ra, being ~2.41 nm and ~1.65 nm for films grown with Bi2Se3 and Bi2Se5 targets, respectively. The hardness (Young’s modulus) of the Bi2Se3 thin films grown from the Bi2Se3 and Bi2Se5 targets were 5.4 GPa (110.2 GPa) and 10.3 GPa (186.5 GPa), respectively. The contact angle measurements of water droplets gave the results that the contact angle (surface energy) of the Bi2Se3 films obtained from the Bi2Se3 and Bi2Se5 targets were 80° (21.4 mJ/m2) and 110° (11.9 mJ/m2), respectively.

1. Introduction

Bismuth selenide (Bi2Se3) is of great interest owing to its intriguing physical properties as a three-dimensional topological insulator [1,2,3,4,5], and potential applications in spintronics [6], optoelectronics [7] and quantum computation [8]. In addition, Bi2Se3 possesses excellent thermoelectric properties at room-temperature [9,10] and low temperature regimes [11]. For fundamental studies and application purposes, it is essential to grow Bi2Se3 thin films with high-quality and to have comprehensive characterizations of their physical properties, including the mechanical properties [12,13,14].
Nanoindentation is a versatile technique ubiquitously used to obtain the basic mechanical parameters, such as the hardness and elastic modulus, as well as to delineate the deformation mechanisms, creep and fracture behaviors of various nanostructured materials [15,16,17,18] and thin films [19,20,21,22,23] with very high sensitivity and excellent resolution. On the other hand, wettability is an important property of a solid surface, which is intimately related to the chemical compositions and morphology of the surface [24]. The peculiar wetting behaviors exhibited on the surface of two-dimensional and van der Waals layered materials have been receiving dramatically increased interest in recent years [25,26,27]. It implies that specific water–substrate interaction features are relevant to the atomic and electronic structures of the layered materials. In particular, the hydrophobic surface (water contact angle, θCA > 90°) can be used in many applications of self-cleaning surfaces and antifogging [28,29]. Consequently, how to control the behavior of hydrophobicity or hydrophilicity of films’ surfaces is also of great importance in realizing the designed functionality for device applications.
Because of the high volatility of selenium (Se), Bi2Se3 tends to form Se vacancies or antisites that serve as donors to result in a sufficiently high carrier concentration and low carrier mobility [30,31]. When severe loss of Se-atoms occurs during the thin-film growth at elevated substrate temperatures, pure phase Bi2Se3 film is usually not achieved, and the obtained films may present impurity phases or even turn into another phase [32]. Thus, to overcome this problem and obtain high-quality stoichiometric Bi2Se3 thin-films, a Se-rich environment is necessary during films’ growth. Indeed, this strategy has been employed to grow high-quality Bi2Se3 thin films by creating a Se-rich environment with a Se:Bi flux ratio ranging from 10:1 to 20:1 using molecular beam epitaxy (MBE) [33,34]. Pulsed laser deposition (PLD) offers a high instantaneous deposition rate, relatively high reproducibility, and low costs. The PLD has been used for growing epitaxial and polycrystalline Bi2Se3 thin films [9,30,35,36,37]. In 2011, Onose et al. [35] successfully grew epitaxial Bi2Se3 thin-films on InP(111) substrates using a designed target with an atomic ratio of Bi:Se of 2:8. Yet, systematic investigations on the effects of target composition, and hence the resultant films’ stoichiometry, on the properties of Bi2Se3 thin films have been relatively scarce.
Herein, we conducted comprehensive characterizations of the structural, compositional, morphological, nanomechanical, and wetting properties of Bi2Se3 thin films grown on InP(111) substrates by PLD. In particular, two different targets (i.e., a stoichiometric target of Bi2Se3 and a Se-rich target of Bi2Se5) were deliberately used to tune the stoichiometry of the resultant Bi2Se3 films and to unveil its effects on the surface wettability and nanomechanical properties, since both characteristics are of pivotal importance for their practical applications in Bi2Se3 thin film-based microelectronic and spintronic devices.

2. Materials and Methods

In order to study the effects of film stoichiometry, two targets with different composition effects were used. One is stoichiometric Bi2Se3 and another is a Se-rich target with a nominal composition of Bi2Se5. The targets were purchased from Ultimate Materials Technology Co., Ltd. (Ping-Tung City, Taiwan). Noticeably, though having differences in Se/Bi atomic ratios of 3/2 and 5/2, both Bi2Se3 and Bi2Se5 targets were polycrystalline with the right Bi2Se3 phase. Bi2Se3 thin films were deposited on InP(111) substrates using PLD at a substrate temperature of 350 °C in vacuum at a base pressure of 4 × 10−6 Torr (~0.53 mPa). For the PLD process, ultraviolet (UV) pulses (20-ns duration) from a KrF excimer laser (λ = 248 nm, repetition: 1 Hz) were focused on the polycrystalline Bi2Se3 or Bi2Se5 target at a fluence of 5.5 J/cm2. The target-to-substrate distance was 40 mm. The target was ablated for approximately 5 min in order to clean its surface before every deposition. The deposition time was 25 min, which resulted in an average Bi2Se3 film thickness of approximately 191 nm (the growth rate of approximately 1.27 Å/pulse).
The crystal structure and surface morphology of the Bi2Se3 thin films were characterized by X-ray diffraction (XRD; Bruker D8, CuKα radiation, λ = 1.5406 Å, Bruker, Billerica, MA, USA) and field emission scanning electron microscopy (SEM, JEOL JSM-6500, JEOL, Pleasanton, CA, USA) operated at an accelerating voltage of 15 kV, respectively. Film compositions were analyzed through Oxford energy-dispersive X-ray spectroscopy (EDS, Inca X-sight 7558, Oxford Instruments plc., Oxfordshire, UK) equipped with the SEM instrument at an accelerating voltage of 15 kV, dead time of 22–30%, and collection time of 60 s. The atomic percentage of each film was determined by averaging the values measured in 5 or more distinct 14 × 20 µm2 areas on the surface of films. Moreover, the surface morphology and roughness of the thin films were examined using atomic force microscopy (AFM; Veeco Escope, Veeco, New York, USA).
The nanoindentation was performed on a Nanoindenter MTS NanoXP® system (MTS Cooperation, Nano Instruments Innovation Center, Oak Ridge, TN, USA) with a pyramid-shaped Berkovich diamond tip. The nanomechanical properties of the Bi2Se3 thin films were measured by nanoindentation with a continuous contact stiffness mode (CSM) [38]. At least 20 indentations were performed on each sample and the distance between the adjacent indents was kept at least 10 μm apart to avoid mutual interferences. We also followed the analytic method proposed by Oliver and Pharr [39] to determine the hardness and Young’s modulus of measured materials from the load–displacement results. Thus, the hardness (H) and Young’s modulus (E) of the Bi2Se3 thin films are obtained and the results are listed in Table 1. Moreover, the surface wettability of the Bi2Se3 thin films under ambient conditions was monitored using a Ramehart Model 200 contact angle goniometer (Ramé-hart, Succasunna, NJ, USA) with deionized water as the liquid.

3. Results and Discussion

3.1. Structural and Morphological Properties

Bi2Se3 has a rhombohedral structure with a space group D 3 d 5 ( R 3 ¯ m ) that can be described by a hexagonal primitive cell with three five-atomic-layer thick lamellae of –(Se(1)–Bi–Se(2)–Bi–Se(1))–, in which the atomic layers are stacked in sequence along the c-axis [9]. The XRD patterns of the Bi2Se3 thin films obtained from the Bi2Se3 and Bi2Se5 targets are shown in Figure 1. As is evident from Figure 1, besides the diffraction peaks of InP substrates at 26.3° and 54.1° (JCPDS PDF#00-032-0452), the films exhibited highly c-axis-preferred orientation with (006), (0015), and (0021) diffraction peaks of the Bi2Se3 phase (JCPDS PDF#33-0214). However, minor diffraction peaks belonging to the BiSe phase (PDF#29-0246) can be identified. It is noticed that, although both of the as-grown films exhibit highly c-axis preferred orientation of the Bi2Se3 phase, a slight relative shift in diffraction angles indicative of modification of the c-axis parameter is observed. Indeed, by using the dominant Bi2Se3 (006) and Bi2Se3 (0015) peaks and the hexagonal unit cell relationship [32], the average c-axis lattice constant of the Bi2Se3 thin films prepared using Bi2Se3 and Bi2Se5 targets were 28.39 Å and 28.25 Å, respectively, whose values were slightly smaller the c-axis lattice constant of 28.63 Å from the database of Bi2Se3 powder (JCPDS PDF#33-0214). This could be due to the difference in the internal stress built up during the deposition.
The grain sizes (D) of the Bi2Se3 films were estimated using the Scherrer equation D = 0.9λ/βcosθ, where λ, β, and θ are the X-ray wavelength, full width at half maximum of the Bi2Se3 (006)-oriented peak, and Bragg diffraction angle, respectively. The estimated D values of the Bi2Se3 thin films prepared using Bi2Se3 target and Bi2Se5 target were 29.7 nm and 26.0 nm, respectively.
Figure 2 shows the AFM and SEM-EDS results of Bi2Se3 thin films prepared using the Bi2Se3 and Bi2Se5 target, respectively. As shown in Figure 2a,b, the films mainly consist of triangular pyramids with features of step-and-terrace structures. This is a clear indication that the films are growing along the [0001] direction, which is consistent with XRD results displayed in Figure 1. The films also exhibit highly smooth surfaces with the centerline average roughness Ra being ~2.41 nm and ~1.65 nm for films grown from the Bi2Se3 target and from the Bi2Se5 target, respectively. In addition, the films grown from the Bi2Se5 target also show clearer step-and-terrace structures with fewer large particle-like outgrowth defects on the surface as compared to the film grown from the Bi2Se3 target (see 3D images), indicating that these films are closer to the stoichiometric composition and, thus, are less defective.
The top-view SEM images displayed in Figure 3a,b further confirmed the aforementioned surface morphology. The cross-sectional view images shown at the bottom of Figure 3a,b indicate that the films are rather uniform with their thickness being in the range of 185~197 nm. Furthermore, as is evident from the EDS results displayed in the insets of Figure 3a,b and the typical EDS spectra of the corresponding thin films shown in Figure 3c, the composition of the film prepared from the Bi2Se3 target clearly showed a substantial Se-deficiency of about 4.4 at.%, while the film prepared from the Bi2Se5 target is nearly stoichiometric, which is consistent with the conjectures discussed above. Intuitively, it is rather straightforward to explain why the Bi2Se3 target would lead to Bi-rich (or Se-deficient) film by recognizing that the re-evaporation of Se from the heated substrate (~350 °C) is much faster than Bi owing the much higher vapor pressure of Se [9,41]. The present results also suggest that to obtain stoichiometric Bi2Se3 films, a Se-excessive target is essential. We note that stoichiometric Bi2Se3 and Bi2Te3 films have been shown to exhibit reduced carrier concentration and increased carrier mobility, which led to the enhanced thermoelectric properties and provided suitable conditions for investigating the topological surface states [9,30,42].

3.2. Nanomechanical Properties

The typical nanoindentation load–displacement curves of Bi2Se3 thin film deposited on InP(111) substrates are shown in Figure 4a. The hardness and Young’s modulus of Bi2Se3 thin films were calculated from the load–displacement curves [39]; the Poisson’s ratio of Bi2Se3 films is set to 0.25 in this study. Figure 4b,c present the penetration depth dependence of hardness and Young’s modulus are obtained using the CSM method. In 2004, Li et al. [15] indicated that nanoindentation depth should never exceed 30% of the film’s thickness. In this work, the CSM technique system is applied to record stiffness data along with load and displacement data dynamically, making it possible to calculate the hardness and Young’s modulus at every data point and get their average values during the indentation experiment [15,39]. The mechanical properties obtained under nanoindentation exhibit a convergent manner and are steady with a rational tolerance around penetrating depths of 40~60nm, reflecting that the material properties obtained are intrinsic and the substrate effect on the present thin films for hardness and modulus tests is negligible. The obtained values of hardness (H) and Young’s modulus (E) are listed in Table 1 together with those reported in the literature for Bi2Se3 single crystals and thin films deposited on sapphire substrates.
From Table 1, it is somewhat surprising to observe that the values of hardness and Young’s modulus of the Bi2Se3 thin films are much larger than those of single crystals. The reason for this peculiar observation, especially the very low values for single crystals, is not clear at present. However, by comparing the results for films, the two prominent mechanical property parameters appear to have intimate correlations with the grain size (D) and surface roughness (Ra). For films grown on InP(111) substrate, as in the present case, the lattice mismatch between the Bi2Se3 thin films and substrate is about 0.2% [35], which, in turn, consistently resulted in films with better crystallinity, as indicated by the narrower full width at half maximum of the diffraction peaks, namely ~0.3° for films grown on InP(111) as compared to that of ~0.5° for the films grown on sapphire substrate [14]. Moreover, when comparing the results for the films grown with different targets, it further indicates that stoichiometry of the film can play an even more prominent role in determining the mechanical properties. Namely, the hardness and Young’s modulus of the stoichiometric Bi2Se3 thin films are both about two times larger than that of Se-deficient films, which are again about two times larger than that grown on sapphire substrate. The enhancement of H and E values can be explained by considering the film crystallinity and surface roughness. It has been reported that the crystallinity of Bi2Se3 thin films deposited on InP(111) substrate was better than those deposited on Al2O3 and Si substrates [35]. In general, better film crystallinity often results in superior nanomechanical properties [43,44]. Therefore, compared with those reported in [14], the larger values of hardness and Young’s modulus of the present Bi2Se3 thin films could be attributed to their better crystallinity. Furthermore, the film surface roughness can also be an important factor. Jian et al. [45] reported that the nanomechanical properties of ZnO thin films were significantly enhanced as the film surfaces became smoother. Even for AISI 316L stainless steel, the mechanical properties were found to decrease with increasing surface roughness [46]. Since the surface roughness of the present films are all below 2.41 nm, it is reasonable to account, at least partially, for the enhanced H and E values.
Turning to the deformation behaviors during nanoindentation, it is evident that there are several pop-ins occurring along the loading segment for both load–displacement curves shown in Figure 4a. It is noted that similar phenomena were found in the previous studies [13,14], where the pop-ins were also observed in nanoindented Bi2Se3 single-crystal and thin films, despite the fact that the loads at which the pop-ins took place varied in each individual measurement. Moreover, it is noted that there is no sign of reverse discontinuity in the unloading portion of the load–displacement curves (the so-called “pop-out” event) being observed. The reverse discontinuity is commonly ascribed to the pressure-induced phase transformation that has been observed in Si or Ge single crystals [47,48]. The absence of these incidences indicates that pressure-induced phase transition did not occur for the Bi2Se3 films in the pressure range applied in this study. In fact, Yu et al. [49] have reported that the pressure-induced phase transition in Bi2Se3 occurred at pressures of 35.6 and 81.2 GPa as revealed, respectively, by Raman spectroscopy and synchrotron XRD experiments conducted in a diamond anvil cell. These values are much higher than the room-temperature hardness of the present hexagonal Bi2Se3 thin films. On the other hand, the pop-in behaviors during nanoindentation have been reported previously in other hexagonal structured materials, such as sapphire [50] and ZnO single crystals [51], as well as GaN thin films [52,53,54] by using the Berkovich indenter tip. It is generally conceived that the nanoindentation-induced deformation mechanism in these hexagonal-structured materials were primarily dominated by the nucleation and/or propagation of dislocations. Thus, it is plausible to believe that similar mechanisms must have been prevailing in the present Bi2Se3 thin films. Reasonably, it can be seen from Table 1 that the hardness of Bi2Se3 thin films increases when D value decreases, partially due to grain boundary hardening.
Within the context of the dislocation-mediated deformation scenarios, the first pop-in event may reflect the transition from perfectly elastic to plastic deformation. Namely, it is the onset of plasticity in Bi2Se3 thin films. Under this circumstance, the corresponding critical shear stress ( τ max ) under the Berkovich indenter at an indentation load, P c , where the load–displacement discontinuity occurs, can be determined by using the following relation [55]:
τ m a x = 0.31 ( 6 P c E 2 π 3 R 2 ) 1 / 3
where R is the radius of the tip of nanoindenter. The obtained τ max values are 1.8 and 3.4 GPa for Bi2Se3 thin films grown using Bi2Se3 and Bi2Se5 targets, respectively. The τ max is responsible for the homogeneous dislocation nucleation within the deformation region underneath the indenter tip.

3.3. Wettability Behavior

The surface wettability of the Bi2Se3 thin films was examined by water contact angle measurements. If the contact angle (θCA) is greater than 90°, it is said to be hydrophobic, otherwise it is hydrophilic. In Figure 5, the values of θCA for films are 80° and 110° for films grown using the Bi2Se3 target and the Bi2Se5 target, respectively.
As described above, the surface roughness measured by the AFM indicated that the Bi2Se3 thin film grown using the Bi2Se5 target have smaller surface roughness, suggesting that the wettability behavior of the surface was significantly affected by the surface morphology of the films [56]. Alternatively, the atomic arrangements and existence of surface defects might also play a role in the eventual surface energy. In general, the surface wettability is a measurement of surface energy and is most commonly quantified by θCA [57]. The surface energy for Bi2Se3 thin films was calculated by means of the Fowkes–Girifalco–Good (FGG) theory [58]. According to the analysis of the FGG method, the considered critical interaction is the dispersive force or the van der Waals force across the interface existing between the water droplet and the solid surface. The FGG equation is given as:
γ l s = γ s + γ l 2 ( γ d ) s + ( γ d ) l
where ( γ d ) s and ( γ d ) l are the dispersive portions of surface tension for the solid and liquid surfaces, respectively. By combining Young’s equation [56] with Equation (2) and taking the nonpolar liquid deionized water as the testing liquid and set ( γ d ) l = γ l , the Girifalco–Good–Fowkes–Young equation becomes as: ( γ d ) s = γ l ( c o s θ C A + 1 ) / 4 , where ( γ d ) s is the surface energy of measured materials. Using γ l = 72.8 mJ/m2, the values of surface energy obtained were 21.4 mJ/m2 and 11.9 mJ/m2 for films grown with the Bi2Se3 target and Bi2Se5 target, respectively. The lower surface energy gives rise to higher hydrophobicity. It is noted that the θCA of 110° for the present stoichiometric Bi2Se3 thin films deposited on InP(111) substrates using PLD is even larger than that (θCA~98.4°) of Bi2Se3 thin films deposited on SrTiO3(111) substrate by MBE [59]. In any case, the present study suggests that both the hydrophobic/hydrophilic transition behavior and nanomechanical properties of the Bi2Se3 thin films can be manipulated by controlling the target compositions.

4. Conclusions

The present study evidently illustrated that stoichiometry, which can be manipulated by tuning the target composition, can give rise to significant effects on the microstructural, morphological, compositional, nanomechanical and surface wetting properties of the Bi2Se3/InP (111) thin films. The Bi2Se3 thin films were grown using PLD from a stoichiometric Bi2Se3 target and a Se-rich Bi2Se5 target at a substrate temperature of 350 °C in a vacuum with a base pressure of ~4 × 10−6 Torr. The films were highly (00l)-oriented with smooth surfaces consisting mainly of triangular step-and-terrace structures, which is the common feature of epitaxial Bi2Se3 thin films. Compared to the films grown from the Bi2Se3 target, using the Bi2Se5 target is more favorable for obtaining stoichiometric films with larger hardness and Young’s modulus. In addition, the contact angle (surface energy) of the Bi2Se3 films deposited from the Bi2Se3 and Bi2Se5 targets were 80° (21.4 mJ/m2) and 110° (11.9 mJ/m2), respectively. These results suggest that, in addition to the usual factors such as surface roughness and grain morphology, stoichiometry as well as defect chemistry originated from Se-deficiency may also play important roles in determining the eventual nanomechanical and wettability properties of Bi2Se3 thin films.

Author Contributions

Data curation, Y.-M.H., C.-T.P., B.-S.C., P.H.L., L.T.C.T., N.N.U. and V.N.; Formal analysis, Y.-M.H., C.-T.P., B.-S.C., L.T.C.T., N.N.U. and V.N.; Funding acquisition, J.-Y.J.; Resources, P.H.L., C.-W.L., J.-Y.J., J.L. and S.-R.J.; Writing—original draft, P.H.L. and S.-R.J.; Writing—review & editing, J.-Y.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology, Taiwan under Contract Nos. MOST 109-2221-E-214-016.

Acknowledgments

The authors would like to thank T.-C. Lin for her technical support in the nanoindentation experiments.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, H.; Liu, C.X.; Qi, X.L.; Dai, X.; Fang, Z.; Zhang, S.C. Topological insulators in Bi2Se3, Bi2Te3 and Sb2Te3 with a single Dirac cone on the surface. Nat. Phys. 2009, 5, 438–442. [Google Scholar] [CrossRef]
  2. Moore, J.E. The birth of topological insulators. Nature 2010, 464, 194–198. [Google Scholar] [CrossRef]
  3. Xia, Y.; Qian, D.; Hsieh, D.; Wray, L.; Pal, A.; Lin, H.; Bansil, A.; Grauer, D.; Hor, Y.S.; Cava, R.J.; et al. Observation of a large-gap topological-insulator class with a single Dirac cone on the surface. Nat. Phys. 2009, 5, 398–402. [Google Scholar] [CrossRef] [Green Version]
  4. Wiedmann, S.; Jost, A.; Fauqué, B.; van Dijk, J.; Meijer, M.J.; Khouri, T.; Pezzini, S.; Grauer, S.; Schreyeck, S.; Brüne, C.; et al. Anisotropic and strong negative magnetoresistance in the three-dimensional topological insulator Bi2Se3. Phys. Rev. B 2016, 94, 081302. [Google Scholar] [CrossRef] [Green Version]
  5. Le, P.H.; Luo, C.W. Ultrafast Dynamics in Topological Insulators; InTech: London, UK, 2018. [Google Scholar]
  6. Yazyev, O.V.; Moore, J.E.; Louie, S.G. Spin polarization and transport of surface states in the topological insulators Bi2Se3 and Bi2Te3 from first principles. Phys. Rev. Lett. 2010, 105, 266806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Min, W.L.; Betancourt, A.P.; Jiang, P.; Jiang, B. Bioinspired broadband antireflection coatings on GaSb. Appl. Phys. Lett. 2008, 92, 141109. [Google Scholar] [CrossRef]
  8. Qi, X.L.; Zhang, S.C. Topological insulators and superconductors. Rev. Mod. Phys. 2011, 83, 1057–1110. [Google Scholar] [CrossRef] [Green Version]
  9. Le, P.H.; Liao, C.N.; Luo, C.W.; Lin, J.Y.; Leu, J. Thermoelectric properties of bismuth-selenide films with controlled morphology and texture grown using pulsed laser deposition. Appl. Surf. Sci. 2013, 285, 657. [Google Scholar] [CrossRef]
  10. Le, P.H.; Luo, C.W. Thermoelectric and Topological Insulator Bismuth Chalcogenide Thin Films Grown Using Pulsed Laser Deposition; InTech: London, UK, 2016. [Google Scholar]
  11. Hor, Y.; Richardella, A.; Roushan, P.; Xia, Y.; Checkelsky, J.; Yazdani, A.; Hasan, M.; Ong, N.; Cava, R. p-type Bi2Se3 for topological insulator and low-temperature thermoelectric applications. Phys. Rev. B 2009, 79, 195208. [Google Scholar] [CrossRef] [Green Version]
  12. Wang, E.; Ding, H.; Fedorov, A.V.; Yao, W.; Li, Z.; Lv, Y.F.; Zhao, K.; Zhang, L.G.; Xu, Z.; Schneeloch, J.; et al. Fully gapped topological surface states in Bi2Se3 films induced by a d-wave high-temperature superconductor. Nat. Phys. 2013, 9, 621–625. [Google Scholar] [CrossRef] [Green Version]
  13. Gupta, S.; Vijayan, N.; Krishna, A.; Thukral, K.; Maurya, K.K.; Muthiah, S.; Dhar, A.; Singh, B.; Bhagavannarayana, G. Enhancement of thermoelectric figure of merit in Bi2Se3 crystals through a necking process. J. Appl. Crystallogr. 2015, 48, 533–541. [Google Scholar] [CrossRef] [Green Version]
  14. Lai, H.D.; Jian, S.R.; Tuyen, L.T.C.; Le, P.H.; Luo, C.W.; Juang, J.Y. Nanoindentation of Bi2Se3 Thin Films. Micromachines 2018, 9, 518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Li, X.D.; Gao, H.; Murphy, C.J.; Gou, L. Nanoindentation of Cu2O nanocubes. Nano Lett. 2004, 4, 1903–1907. [Google Scholar] [CrossRef]
  16. Bao, L.; Xu, Z.H.; Li, R.; Li, X. Catalyst-free synthesis and structural and mechanical characterization of single crystalline Ca2B2O5.H2O nanobelts and stacking faulted Ca2B2O5 nanogrooves. Nano Lett. 2010, 10, 255–262. [Google Scholar] [CrossRef]
  17. Jiang, Y.; Hor, J.L.; Lee, D.; Turner, K.T. Toughening nanoparticle films via polymer infiltration and confinement. ACS Appl. Mater. Interfaces 2018, 10, 44011–44017. [Google Scholar] [CrossRef]
  18. Ma, Y.; Huang, X.; Song, Y.; Hang, W.; Zhang, T. Room-temperature creep behavior and activation volume of dislocation nucleation in a LiTaO3 single crystal by nanoindentation. Materials 2019, 12, 1683. [Google Scholar] [CrossRef] [Green Version]
  19. Zaman, A.; Meletis, E.I. Microstructure and mechanical properties of TaN thin films prepared by reactive magnetron sputtering. Coatings 2017, 7, 209. [Google Scholar] [CrossRef] [Green Version]
  20. Azizpour, A.; Hahn, R.; Klimashin, F.F.; Wojcik, T.; Poursaeidi, E.; Mayrhofer, P.H. Deformation and cracking mechanisms in CrN/TiN multilayer coatings. Coatings 2019, 9, 363. [Google Scholar] [CrossRef] [Green Version]
  21. Wiatrowski, A.; Obstarczyk, A.; Mazur, M.; Kaczmarek, D.; Wojcieszark, D. Characterization of HfO2 optical coatings deposited by MF magnetron sputtering. Coatings 2019, 9, 106. [Google Scholar] [CrossRef] [Green Version]
  22. Tuyen, L.T.C.; Jian, S.R.; Tien, N.T.; Le, P.H. Nanomechanical and material properties of Fluorine-doped tin oxide thin films prepared by ultrasonic spray pyrolysis: Effects of F-doping. Materials 2019, 12, 1665. [Google Scholar] [CrossRef] [Green Version]
  23. Hwang, Y.M.; Pan, C.T.; Lu, Y.X.; Jian, S.R.; Chang, H.W.; Juang, J.Y. Influence of post-annealing on the structural and nanomechanical properties of Co thin films. Micromachines 2020, 11, 180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Quere, D. Wetting and Roughness. Ann. Rev. Mater. Res. 2008, 38, 71–99. [Google Scholar] [CrossRef]
  25. Rafiee, J.; Mi, X.; Gullapalli, H.; Thomas, A.V.; Yavari, F.; Shi, Y.; Ajayan, P.M.; Koratkar, N.A. Wetting transparency of graphene. Nat. Mater. 2012, 11, 217–222. [Google Scholar] [CrossRef] [PubMed]
  26. Gaur, A.P.S.; Sahoo, S.; Ahmadi, M.; Dash, S.P.; Guinel, M.J.-F.; Katiyar, R.S. Surface energy engineering for tunable wettability through controlled synthesis of MoS2. Nano Lett. 2014, 14, 4314–4321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Chow, P.K.; Singh, E.; Viana, B.C.; Gao, J.; Luo, J.; Li, J.; Lin, Z.; Arriaga, L.E.; Shi, Y.; Wang, Z.; et al. Wetting of mono and few-layered WS2 and MoS2 films supported on Si/SiO2 substrates. ACS Nano 2015, 9, 30023–30031. [Google Scholar] [CrossRef] [PubMed]
  28. Fujishima, A.; Rao, T.N.; Tryk, D.A. Titanium dioxide photocatalysis. J. Photochem. Photobiol. C Photochem. Rev. 2000, 1, 1–21. [Google Scholar] [CrossRef]
  29. Guo, Z.; Liu, W.; Su, B.L. Superhydrophobic surfaces: From natural to biomimetic to functional. J. Colloid Interface Sci. 2011, 353, 335–355. [Google Scholar] [CrossRef]
  30. Le, P.H.; Wu, K.H.; Luo, C.W.; Leu, J. Growth and characterization of topological insulator Bi2Se3 thin films on SrTiO3 using pulsed laser deposition. Thin Solid Film. 2013, 534, 659–665. [Google Scholar] [CrossRef]
  31. Richardella, A.; Zhang, D.M.; Lee, J.S.; Koser, A.; Rench, D.W.; Yeats, A.L.; Buckley, B.B.; Awschalom, D.D.; Samarth, N. Coherent heteroepitaxy of Bi2Se3 on GaAs (111)B. Appl. Phys. Lett. 2010, 97, 262104. [Google Scholar] [CrossRef]
  32. Tuyen, L.T.C.; Le, P.H.; Luo, C.W.; Leu, J. Thermoelectric properties of nanocrystalline Bi3Se2Te thin films grown using pulsed laser deposition. J. Alloy. Compd. 2016, 673, 107–114. [Google Scholar] [CrossRef]
  33. Wang, Z.Y.; Guo, X.; Li, H.D.; Wong, T.L.; Wang, N.; Xie, M.H. Superlattices of Bi2Se3/In2Se3: Growth characteristics and structural properties. Appl. Phys. Lett. 2011, 99, 023112. [Google Scholar] [CrossRef] [Green Version]
  34. Taskin, A.A.; Sasaki, S.; Segawa, K.; Ando, Y. Manifestation of topological protection in transport properties of epitaxial Bi2Se3 thin films. Phys. Rev. Lett. 2012, 109, 066803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Onose, Y.; Yoshimi, R.; Tsukazaki, A.; Yuan, H.; Hidaka, T.; Iwasa, Y.; Kawasaki, M.; Tokura, Y. Pulsed laser deposition and ionic liquid gate control of epitaxial Bi2Se3 thin films. Appl. Phys. Express. 2011, 4, 083001. [Google Scholar] [CrossRef]
  36. Lee, Y.F.; Punugupati, S.; Wu, F.; Jin, Z.; Narayan, J.; Schwartz, J. Evidence for topological surface states in epitaxial Bi2Se3 thin film grown by pulsed laser deposition through magneto-transport measurements. Curr. Opin. Solid State Mater. Sci. 2014, 18, 279–285. [Google Scholar] [CrossRef]
  37. Orgiani, P.; Bigi, C.; Das, P.K.; Fujii, J.; Ciancio, R.; Gobaut, B.; Galdi, A.; Sacco, C.; Maritato, L.; Torelli, P.; et al. Structural and electronic properties of Bi2Se3 topological insulator thin films grown by pulsed laser deposition. Appl. Phys. Lett. 2017, 110, 171601. [Google Scholar] [CrossRef] [Green Version]
  38. Li, X.; Bhushan, B. A review of nanoindentation continuous stiffness measurement technique and its applications. Mater. Charact. 2002, 48, 11–36. [Google Scholar] [CrossRef]
  39. Oliver, W.C.; Pharr, G.M. An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments. J. Mater. Res. 1992, 7, 1564–1583. [Google Scholar] [CrossRef]
  40. Chrobak, D.; Chrobak, A.; Nowak, R. Effect of doping on nanoindentation induced incipient plasticity in InP crystal. AIP Adv. 2019, 9, 125323. [Google Scholar] [CrossRef] [Green Version]
  41. Noro, H.; Sato, K.; Kagechika, H. The thermoelectric properties and crystallography of Bi-Sb-Te-Se thin films grown by ion beam sputtering. J. Appl. Phys. 1993, 73, 1252–1260. [Google Scholar] [CrossRef]
  42. Le, P.H.; Liao, C.N.; Luo, C.W.; Lin, J.Y.; Leu, J. Thermoelectric properties of nanostructured bismuth–telluride thin films grown using pulsed laser deposition. J. Alloy. Compd. 2014, 615, 546–552. [Google Scholar] [CrossRef]
  43. Wang, S.H.; Jian, S.R.; Chen, G.J.; Cheng, H.Z.; Juang, J.Y. Annealing-driven microstructural evoluation and its effects on the surface and nanomechanical properties of Cu-doped NiO thin films. Coatings 2019, 9, 107. [Google Scholar] [CrossRef] [Green Version]
  44. Lou, B.S.; Moirangthem, I.; Lee, J.W. Fabrication of tungsten nitride thin films by superimposed HiPIMS and MF system: Effects of nitrogen flow rate. Surf. Coat. Technol. 2020, 393, 125743. [Google Scholar] [CrossRef]
  45. Jian, S.R.; Teng, I.J.; Yang, P.F.; Lai, Y.S.; Lu, J.M.; Chang, J.G.; Ju, S.P. Surface morphological and nanomechanical properties of PLD-derived ZnO thin films. Nanoscale Res. Lett. 2008, 3, 186–193. [Google Scholar] [CrossRef] [Green Version]
  46. Walter, C.; Antretter, T.; Daniel, R.; Mitterer, C. Finite element simulation of the effect of surface roughness on nanoindentation of thin films with spherical indenters. Surf. Coat. Technol. 2007, 202, 1103–1107. [Google Scholar] [CrossRef]
  47. Jian, S.R.; Chen, G.J.; Juang, J.Y. Nanoindentation-induced phase transformation in (11 0)-oriented Si single-crystals. Curr. Opin. Solid State Mater. Sci. 2010, 14, 69–74. [Google Scholar] [CrossRef]
  48. Jang, J.; Lance, M.J.; Wen, S.; Pharr, G.M. Evidence for nanoindentation-induced phase transformations in germanium. Appl. Phys. Lett. 2009, 86, 131907. [Google Scholar] [CrossRef]
  49. Yu, Z.; Wang, L.; Hu, Q.; Zhao, J.; Yan, S.; Yang, K.; Sinogeikin, S.; Gu, G.; Mao, H.-K. Structural phase transitions in Bi2Se3 under high pressure. Sci. Rep. 2015, 5, 1–9. [Google Scholar] [CrossRef] [Green Version]
  50. Mao, W.G.; Shen, Y.G.; Lu, C. Nanoscale elastic-plastic deformation and stress distributions of the C plane of sapphire single crystal during nanoindentation. J. Eur. Ceram. Soc. 2011, 31, 1865–1871. [Google Scholar] [CrossRef] [Green Version]
  51. Jian, S.R. Pop-in effects and dislocation nucleation of c-plane single-crystal ZnO by Berkovich nanoindentation. J. Alloy. Compd. 2015, 644, 54–58. [Google Scholar] [CrossRef]
  52. Chien, C.H.; Jian, S.R.; Wang, C.T.; Juang, J.Y.; Huang, J.C.; Lai, Y.S. Cross-sectional transmission electron microscopy observations on the Berkovich indentation-induced deformation microstructures in GaN thin films. J. Phys. D Appl. Phys. 2007, 40, 3985. [Google Scholar] [CrossRef]
  53. Jian, S.R. Mechanical deformation induced in Si and GaN under Berkovich nanoindentation. Nanoscale Res. Lett. 2008, 3, 6–13. [Google Scholar] [CrossRef] [Green Version]
  54. Jian, S.R. Cathodoluminescence rosettes in c-plane GaN films under Berkovich nanoindentation. Opt. Mater. 2013, 35, 2707–2709. [Google Scholar] [CrossRef]
  55. Johnson, K.L. Contact Mechanics; Cambridge University Press: Cambridge, UK, 1985. [Google Scholar]
  56. Ottone, C.; Lamberti, A.; Fontana, M.; Cauda, V. Wetting Behavior of Hierarchical Oxide Nanostructures: TiO2 Nanotubes from Anodic Oxidation Decorated with ZnO Nanostructures. J. Electrochem. Soc. 2014, 164, D484. [Google Scholar] [CrossRef]
  57. Angelo, M.S.; McCandless, B.E.; Birkmire, R.W.; Rykov, S.A.; Chen, J.G. Contact wetting angle as a characterization for processing CdTe/CdS solar cells. Pro. Photovolt. 2007, 15, 93–111. [Google Scholar] [CrossRef]
  58. Mahadik, D.B.; Rao, A.V.; Parale, V.G.; Kavale, M.S.; Wagh, P.B.; Ingale, S.V.; Gupta, S.C. Effect of surface composition and roughness on the apparent surface free energy of silica aerogel materials. Appl. Phys. Lett. 2011, 99, 104104. [Google Scholar] [CrossRef]
  59. Zhao, P.; Huang, Y.; Shen, Y.; Yang, S.; Chen, L.; Wu, K.; Li, H.; Meng, S. A modified Wenzel model for water wetting on van der Waals layered materials with topographic surfaces. Nanoscale 2017, 9, 3843–3849. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD patterns of Bi2Se3 thin films grown on InP (111) substrates from two different targets of Bi2Se3 (a) and Bi2Se5 (b) using pulsed laser deposition.
Figure 1. XRD patterns of Bi2Se3 thin films grown on InP (111) substrates from two different targets of Bi2Se3 (a) and Bi2Se5 (b) using pulsed laser deposition.
Coatings 10 00958 g001
Figure 2. 2D and 3D AFM images of the Bi2Se3 thin films deposited from (a) Bi2Se3 target and (b) Bi2Se5 target.
Figure 2. 2D and 3D AFM images of the Bi2Se3 thin films deposited from (a) Bi2Se3 target and (b) Bi2Se5 target.
Coatings 10 00958 g002
Figure 3. Top-view and cross-sectional SEM images of the Bi2Se3 thin films deposited from (a) Bi2Se3 target and (b) Bi2Se5 target. (c) EDS spectra of the corresponding Bi2Se3 thin films.
Figure 3. Top-view and cross-sectional SEM images of the Bi2Se3 thin films deposited from (a) Bi2Se3 target and (b) Bi2Se5 target. (c) EDS spectra of the corresponding Bi2Se3 thin films.
Coatings 10 00958 g003
Figure 4. (a) The load–displacement curves of Bi2Se3 thin films deposited on InP(111) substrates using two different target compositions. A clear single “pop-in” behavior is displayed in both curves during loading. (b) A hardness—displacement curve and (c) a Young’s modulus–displacement curve for a Bi2Se3 thin films deposited using Bi2Se3 and Bi2Se5 targets.
Figure 4. (a) The load–displacement curves of Bi2Se3 thin films deposited on InP(111) substrates using two different target compositions. A clear single “pop-in” behavior is displayed in both curves during loading. (b) A hardness—displacement curve and (c) a Young’s modulus–displacement curve for a Bi2Se3 thin films deposited using Bi2Se3 and Bi2Se5 targets.
Coatings 10 00958 g004
Figure 5. Contact angle test: the images of water droplets on the Bi2Se3 thin film surfaces.
Figure 5. Contact angle test: the images of water droplets on the Bi2Se3 thin film surfaces.
Coatings 10 00958 g005
Table 1. The microstructural parameters, nanomechanical properties, contact angle and surface energy of Bi2Se3 thin films. The mechanical properties of InP(111) are also listed.
Table 1. The microstructural parameters, nanomechanical properties, contact angle and surface energy of Bi2Se3 thin films. The mechanical properties of InP(111) are also listed.
SampleD
(nm)
Ra
(nm)
H
(GPa)
E
(GPa)
τ m a x
(GPa)
θCA ( γ d ) s
(mJ/m2)
Bi2Se3 thin film on InP(111) substrate (Bi2Se3 target)29.72.415.4110.21.880°21.4
Bi2Se3 thin film on InP(111) substrate (Bi2Se5 target)26.01.6510.3186.53.4110°11.9
Bi2Se3 thin film on sapphire substrate [14]34.28.5~2.1~58.6~0.7
Single-crystal Bi2Se3 [13]~0.4–0.9~2–9
Single-crystal InP(111) [40]~572.4–76.21.96

Share and Cite

MDPI and ACS Style

Hwang, Y.-M.; Pan, C.-T.; Chen, B.-S.; Le, P.H.; Uyen, N.N.; Tuyen, L.T.C.; Nguyen, V.; Luo, C.-W.; Juang, J.-Y.; Leu, J.; et al. Effects of Stoichiometry on Structural, Morphological and Nanomechanical Properties of Bi2Se3 Thin Films Deposited on InP(111) Substrates by Pulsed Laser Deposition. Coatings 2020, 10, 958. https://doi.org/10.3390/coatings10100958

AMA Style

Hwang Y-M, Pan C-T, Chen B-S, Le PH, Uyen NN, Tuyen LTC, Nguyen V, Luo C-W, Juang J-Y, Leu J, et al. Effects of Stoichiometry on Structural, Morphological and Nanomechanical Properties of Bi2Se3 Thin Films Deposited on InP(111) Substrates by Pulsed Laser Deposition. Coatings. 2020; 10(10):958. https://doi.org/10.3390/coatings10100958

Chicago/Turabian Style

Hwang, Yeong-Maw, Cheng-Tang Pan, Bo-Syun Chen, Phuoc Huu Le, Ngo Ngoc Uyen, Le Thi Cam Tuyen, Vanthan Nguyen, Chih-Wei Luo, Jenh-Yih Juang, Jihperng Leu, and et al. 2020. "Effects of Stoichiometry on Structural, Morphological and Nanomechanical Properties of Bi2Se3 Thin Films Deposited on InP(111) Substrates by Pulsed Laser Deposition" Coatings 10, no. 10: 958. https://doi.org/10.3390/coatings10100958

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